Sunteți pe pagina 1din 321

Finite Element Analysis of Composite Structures

Marcos Latorre, Francisco J. Montáns. ETSIAE. UPM.

October 14, 2014


- 2-
Contents

Contents
1 Introduction to Composite Materials 1
1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Types of composite materials . . . . . . . . . . . . . . . . . . . . . . 2
1.2.1 Unidirectionally fiber-reinforced composites . . . . . . . . . . 4
1.2.2 Laminated fiber-reinforced composites . . . . . . . . . . . . . 5
1.3 Fabrication and assembly . . . . . . . . . . . . . . . . . . . . . . . . 8
1.4 Advantages and disadvantages of laminated composites . . . . . . . 8
1.5 Use of composite materials in industry . . . . . . . . . . . . . . . . . 9

2 Mechanical Analysis of a Lamina 13


2.1 Stress and strain tensors . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.1.1 Concept of stress . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.1.2 Concept of strain . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.1.3 Tensorial nature of stresses and strains . . . . . . . . . . . . . 15
2.2 Constitutive equations for isotropic solids . . . . . . . . . . . . . . . 19
2.2.1 Engineering (Voigt) notation . . . . . . . . . . . . . . . . . . 22
2.3 The linearized stress-strain behavior for anisotropic materials . . . . 24
2.3.1 Monoclinic material . . . . . . . . . . . . . . . . . . . . . . . 24
2.3.2 Orthotropic material . . . . . . . . . . . . . . . . . . . . . . . 26
2.3.3 Transversely isotropic material . . . . . . . . . . . . . . . . . 29
2.4 Apparent material constants. Coupling coefficients. . . . . . . . . . . 31
2.4.1 Apparent material constants . . . . . . . . . . . . . . . . . . 31
2.4.2 Coupling coefficients . . . . . . . . . . . . . . . . . . . . . . . 35
2.4.3 Particular case: Specially orthotropic lamina . . . . . . . . . 38
2.4.4 Restriction on engineering constants for an orthotropic material 39
2.4.5 Reduced stress-strain relations for plane stress . . . . . . . . 41
2.5 Temperature, moisture and piezoelectric effects . . . . . . . . . . . . 51
2.5.1 Temperature effects . . . . . . . . . . . . . . . . . . . . . . . 51
2.5.2 Moisture effects . . . . . . . . . . . . . . . . . . . . . . . . . . 55
2.5.3 Piezoelectric effects . . . . . . . . . . . . . . . . . . . . . . . . 56
2.6 Failure criteria for composite laminae . . . . . . . . . . . . . . . . . . 58
2.6.1 Strength of an orthotropic lamina in principal material directions 59
2.6.2 Usual strength criteria for biaxial in-plane stresses . . . . . . 61
2.7 Determination of the overall behavior of a lamina through the behav-
ior of its constituents. Strength of materials approach. . . . . . . . . 75
2.8 Determination of the overall behavior of a lamina through the behav-
ior of its constituents. Theory of elasticity approach. . . . . . . . . . 85
2.8.1 Lower and upper bounds on preferred Young’s moduli . . . . 85
2.8.2 The Halpin-Tsai Equations . . . . . . . . . . . . . . . . . . . 91
2.9 Determination of the strength properties of a lamina from its con-
stituents. Strength of materials approach. . . . . . . . . . . . . . . . 92

- 3-
Contents

3 Constitutive Equations for Laminated Composites 99


3.1 Review of the theory of isotropic plates . . . . . . . . . . . . . . . . 101
3.2 Laminated Plate Theory . . . . . . . . . . . . . . . . . . . . . . . . . 108
3.2.1 Coupling coefficients . . . . . . . . . . . . . . . . . . . . . . . 113
3.2.2 Standard Laminate Code . . . . . . . . . . . . . . . . . . . . 114
3.3 Symmetric and antisymmetric laminates . . . . . . . . . . . . . . . . 130
3.3.1 Single-layered orthotropic plate . . . . . . . . . . . . . . . . . 130
3.3.2 Symmetric laminates . . . . . . . . . . . . . . . . . . . . . . 131
3.3.3 Antisymmetric laminates . . . . . . . . . . . . . . . . . . . . 134
3.3.4 Other specific ply layups . . . . . . . . . . . . . . . . . . . . . 137
3.4 Temperature and moisture effects in a laminate . . . . . . . . . . . . 140
3.4.1 Temperature distribution in the laminate . . . . . . . . . . . 140
3.4.2 Hygrothermal loads in the laminate . . . . . . . . . . . . . . 143
3.4.3 Recovery of stresses in the laminae . . . . . . . . . . . . . . . 144
3.5 Piezoelectric effects in the laminate . . . . . . . . . . . . . . . . . . 144

4 Mechanical analysis of composite laminates 147


4.1 Mechanical failure in composites . . . . . . . . . . . . . . . . . . . . 147
4.2 Strength of laminates . . . . . . . . . . . . . . . . . . . . . . . . . . 148
4.2.1 Failure mechanisms in a lamina . . . . . . . . . . . . . . . . . 148
4.2.2 Additional considerations on failure criteria for laminae . . . 153
4.2.3 Progressive failure analysis . . . . . . . . . . . . . . . . . . . 161
4.3 Governing equations of a laminated plate . . . . . . . . . . . . . . . 165
4.3.1 Equilibrium equations in terms of stress resultants . . . . . . 167
4.3.2 Equilibrium equations in terms of displacements . . . . . . . 177
4.3.3 Matrix-operator form of the field equation . . . . . . . . . . . 182
4.3.4 Boundary conditions for the field equations . . . . . . . . . . 182
4.3.5 Initial conditions for the field equations . . . . . . . . . . . . 188
4.3.6 The Airy stress function . . . . . . . . . . . . . . . . . . . . . 188
4.3.7 The energy approach . . . . . . . . . . . . . . . . . . . . . . . 190
4.4 Deformations in composite laminated plates: some simple cases . . . 194
4.4.1 Laminated beams . . . . . . . . . . . . . . . . . . . . . . . . . 195
4.4.2 Cylindrical bending . . . . . . . . . . . . . . . . . . . . . . . 201
4.4.3 Laminated simply supported rectangular plates: the Navier
solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 204
4.4.4 Laminated rectangular plates with two edges simply supported:
the Levy-Nadai solution . . . . . . . . . . . . . . . . . . . . . 211
4.5 Buckling analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214
4.5.1 Buckling of symmetric laminated beams . . . . . . . . . . . . 214
4.5.2 Buckling of specially orthotropic laminated plates . . . . . . 219
4.5.3 Buckling of antisymmetric laminated plates . . . . . . . . . . 229
4.6 Vibration analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235
4.6.1 Vibration of symmetric laminated beams . . . . . . . . . . . 236
4.6.2 Vibration of simply supported specially orthotropic plates . . 239
4.6.3 Vibration of antisymmetric laminates . . . . . . . . . . . . . 243
4.7 Edge interlaminar stresses . . . . . . . . . . . . . . . . . . . . . . . . 254

- 4-
Contents

4.8 Considerations on fatigue and fracture in anisotropic laminates. . . . 266


4.8.1 Fracture mechanics . . . . . . . . . . . . . . . . . . . . . . . . 266
4.8.2 Fatigue . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 284
4.9 Introduction to the design of composite structures . . . . . . . . . . 285
4.9.1 The approach . . . . . . . . . . . . . . . . . . . . . . . . . . . 286
4.9.2 The details . . . . . . . . . . . . . . . . . . . . . . . . . . . . 287
4.9.3 Responsibility of the structure . . . . . . . . . . . . . . . . . 287
4.9.4 Mathematical tools for optimal design . . . . . . . . . . . . . 288

A Cauchy’s Tetrahedron 289

B Transformation Rules and Rotations of Vectors and Second-Order


Tensors 291
B.1 Transformation Rules for Change of Coordinates . . . . . . . . . . . 291
B.2 Rotations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 293
B.3 Relation between Change of Basis Rules and Rotations . . . . . . . . 295

C Thermodynamics of solids 296


C.1 Temperature effects . . . . . . . . . . . . . . . . . . . . . . . . . . . 296
C.2 Hygrometric effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . 305
C.3 Piezoelectric and pyroelectric effects . . . . . . . . . . . . . . . . . . 307

- 5-
1 Introduction to Composite Materials
1.1 Introduction
According to the meaning of the word “composite”, i.e. “consisting of two or more
distinct parts”, a composite material consists simply of two or more distinct mate-
rials which are joined together to form a unit material. Attending strictly to this
general definition, homogeneous materials such as the metal alloys may be consid-
ered composite materials. However, in this text we are devoted to the study of other
specific type of composites. We will refer to a composite material when the “distinct
parts” which constitute the specimen are entirely distinguishable from each other,
or more specifically, when there exist one or more discontinuous phases, which we
call the reinforcement, embedded in a continuous phase, which we call the matrix.
The resulting “unit material” will be then heterogeneous.
A composite material is not just another new material. As we will see below,
composite materials require unconventional approaches in the design process. Ide-
ally, the main purpose is to join together several media to obtain composite materials
which take the advantages of each and every one of their constituents. Of course,
the accomplishment of this synergistic requirement implies a previous thorough un-
derstanding of the properties of the constituents, of the role of the interface through
which they interact and of the consequences of joining them together. Although this
is certainly a huge and difficult task, many important advances have been achieved
for the last decades in this field.
Aside, due primarily to the high stiffness-to-density and strength-to-density ra-
tios that composites usually present, they have found their way in a wide variety
of industrial applications in which the reduction of weight, without an associated
decrease of stiffness and strength, is an important design requirement. Modern air-
craft industry, where weight reduction is extremely important for cost savings and
increase of the maneuverability, irremediably comes to mind as an ideal example.

A historical review
Composite materials have been present since the beginning of time. Living beings
such as plants or animals are made up of materials containing several constituents.
Just to enumerate a few examples, the wood of trees contains cellulose fibers within
a lignin matrix, the coconut palm leaf can be seen as a cantilever making use of
the concept of fiber reinforcement, the skin and muscles are living tissues including
collagen fibers, the bones also contain short and soft collagen fibers embedded in a
mineral medium, and so on.
Aside, historical examples of man-made composites are abundant in the litera-
ture. The first fiber-composite material seems to be the papyrus paper made by the
Egyptians around 4000 B.C. This paper was formed by two cuttings of the fibrous
papyrus plant being laid up at a right angle relative to one another. We will see
below, Chapter 3, that such a system of fiber-reinforced layers is currently known
as an antisymmetric cross-ply laminate. There exist also evidences that bundles of
cuttings of this plant were used by early Egyptians to build, for example, boats,
sails and ropes.

- 1-
1 Introduction to Composite Materials

Other early uses of composite materials include the addition of straw to bricks of
mud (since 1300 B.C. up to the present days!), a practice initiated by the Israelites.
First, the straw assisted the evaporation of moisture from the interior of the brick,
which reduced the required drying time. Second, it helped to redistribute the cracks
evenly, thereby enhancing the overall strength of this structural element. Note
this truly constitutes a synergistic combination of materials, since the individual
constituents could not serve the function by themselves. As a matter of fact, the
Book of Exodus (Exodus 5 : 7) states “do not supply them straw for their bricks,
let them gather their own straw”. We would identify today this unit material as a
randomly oriented, short fiber composite. Other significant examples include the use
of bamboo shoots to reinforce mud walls in houses and laminated metals in forging
swords.
Finally, production of composite materials being reinforced by continuous glass
fibers (today commonly known as E-glass and marketed under the trade name Fiber-
glass) began in 1939. Other fabrics made up with continuous advanced fibers, such
as boron, carbon and aramid (Kevlar), came into commercial production in the late
1950s and early 1960s. Since then, the use of fiber-composite materials has been
growing at a rapid rate. Evidently, at the same time, the development of theoretical
and applied mechanics for the study of these composites has experienced a great,
valuable progress.

1.2 Types of composite materials


We provide in this section a classification of composite materials which fit to the
definition given in the preceding section, i.e. composites containing one or more
discontinuous reinforcing phases embedded in a continuous matrix. In other words,
they are composites whose constituents can be distinguishable to the naked eye.
In general, the reinforcement within a composite material can appear in the
form of particles to form particulate composites or in the form of discontinuous or
continuous fibers to form fibrous composites. Particles are nonfibrous in nature and
have shapes approximately equiaxed (spherical, cubic, tetragonal, etcetera). On the
contrary, fibers are characterized by having a length greater (if whiskers) or much
greater (if continuous fibers) than their cross-sectional dimensions.
Particulate composites may be also defined as low-performance composites be-
cause the particles really provide some stiffening and strengthening, but the matrix
remains as the main responsible for carrying the loads acting over the material. That
way, particles are not very effective in improving the stiffness and strength of the ma-
trix, fact which explains the denomination of low-performance composites. Instead,
particle fillers are widely used because they can modify more effectively other im-
portant properties of matrix materials. For example, they are useful for improving
thermal and electrical conductivities, wear and abrasion resistance, machinability
and high temperature performance, among other matrix properties.
Short fibers embedded within a matrix material provide, in general, greater
mechanical properties to those given by particles to the matrix. This is mainly
because these whiskers or ribbons (with length-to-diameter ratios in the hundreds)
have generally high stiffness and strength individual properties, which causes that

- 2-
1.2 Types of composite materials

they can contribute to increase the mechanical performance of the matrix material
to a large extent. This geometrical characteristic also enables them to be delib-
erately oriented along a certain direction, hence defining a preferred direction of
anisotropy. The mechanical properties of the resulting preferentially oriented, short
fiber composite are enhanced in that direction. However, in most cases the fibers are
assumed to be randomly oriented within the composite and the properties of the re-
sulting discontinuous-fiber-reinforced composite are isotropic, that is, not direction-
dependent. In some particulate composites, the inclusions can also be preferentially
dispersed in order to provide some kind of anisotropy.
Filamentary composites can be arguably considered the true high-performance
composites. The long (continuous) fibers are embedded within the matrix and are
oriented about a certain preferred direction. The main difference with its particulate
counterpart is that the continuous fibers really constitute the load carrying media
within the composite, being able to provide high stiffness and strength when these
properties are measured in the direction of the fibers. In this case, the main role of
the matrix consists of giving protection and support to the fibers, binding them to-
gether, and acting as a load splicing medium. However, as we will see in Section 2.9,
the matrix can also contribute decisively toward the strength (failure) performance
of the composite. Moreover, we will explain why its presence becomes necessary
when a fiber is broken. This type of fibrous composites are commonly manufactured
as thin laminae or plies containing fibers aligned in the same direction. Thin fibers
(boron being an exception) can also be woven in two or more directions within the
same ply to give a woven fabric.
A classification scheme for the three types of composites being considered is
shown next
 
  Randomly oriented


 Particulate

 composites

 
 Preferentially oriented





  


   Randomly oriented
Composite 

 Discontinuous
Materials  
 (short) fibers 


 
 Preferentially oriented


 Fibrous

 composites  


 
  Unidirectionally reinforced





 Continuous

 

  (long) fibers 

Bidirectionally reinforced

In Figure 1, the arrangement of the reinforcement within the matrix for the com-
posites being considered is shown.

For all the reinforcements being presented, the presence of a binder material that
gives the final form of the structural element emerges as necessary. Polymers (resins)
are the matrix materials most extensively used to form fiber-reinforced composites.
Typically, these polymeric materials present lower density, stiffness and strength
than the fibers. They are usually thermosets, which are irreversibly cross-linked in

- 3-
1 Introduction to Composite Materials

a) b)

c) d)

e) f)

Figure 1: Types of reinforcement in composites: a) Randomly oriented, particulate


composite; b) Preferentially oriented, particulate composite; c) Randomly oriented,
short fiber composite; d) Preferentially oriented, short fiber composite; e) Unidirec-
tional filamentary composite; f) Woven fabric.

a three-dimensional network (in clear contrast to thermoplastics) during a process


known as curing. After this chemical reaction is induced, the form of the matrix
containing the fibers, that is, the form of the resulting composite, cannot be changed
in form. Below, we give some few details associated to the fabrication and assembly
processes using this type of matrices. Other binder materials used to a lesser extent
include metals, ceramics and carbon.

1.2.1 Unidirectionally fiber-reinforced composites


Composite materials being reinforced by continuous fibers have become more promi-
nent because most materials are stronger and stiffer in the fibrous form than in any
other form due to the well-known size-effect. By comparison of theoretical and ex-
perimental results, it is observed that measured strengths of most materials are much
smaller (by a couple of orders of magnitude) than their theoretical strength. The
presence of imperfections within the actual material, which are not included in the
theoretical models, satisfactorily explains these differences. The imperfections can
be minimized by means of reducing the cross-sectional dimensions of the specimen,
thereby strengthen the material (see Ref. [1]). This fact motivates the use of high-
performance composites containing fibers with a very high length-to-diameter ratio.
In Table 1 it can be observed that materials in fiber form effectively acquire higher
ultimate strengths than conventional, bulk materials. An ideal situation in which
imperfections would be completely eliminated would consist of man-made filaments
with a cross-sectional diameter equal to the crystal size. A cut-edge technology
tending to this bound is currently developing the so-called nanocomposites, whose
fibers (or nanotubes) have a diameter in the scale of nanometers (10−9 m). In con-
trast, conventional fibers have typically reinforcing constituents on the microscale
(10−6 m).

- 4-
1.2 Types of composite materials

Tensile Tensile Specific Specific


modulus strength Density stiffness strength
Material
E (GP a) σu (GP a) ρ (g / cm3 ) E/ρ σu /ρ
Fibers
E-glass 72.4 3.5 2.54 28.5 1.38
S-glass 85.5 4.6 2.48 34.5 1.85
Kevlar 49 130 2.8 1.50 87 1.87
Graphite 240 2.5 1.90 126 1.3
Boron 385 2.8 2.63 146 1.1
Beryllium 240 1.3 1.83 131 0.71
Bulk materials
Mild Steel 210 0.83 7.8 26.9 0.11
Al (2024-T4) 73 0.41 2.7 27.0 0.15
Glass 70 2.1 2.5 28.0 0.84
Beryllium 300 0.7 1.83 164 0.38

Table 1: Properties of fibers and conventional monolithic materials. Adapted from


Ref. [2]. The “E” in E-glass stands for “Electrical” because it was originally designed
for electrical applications (today is used for many other purposes). The “S” in S-
Glass is due to the higher content in “Silica” of this fiber. Kevlar 49 is an aramid
fiber, product of DuPont.

Tables 2 and 3 presents the key characteristics of fibers and matrix materials
most widely used in filamentary composites. The commonly accepted abbreviations
for these constituents are also indicated.

Fibers Cost Advantages Drawbacks


Good damage tolerance Relatively high density
Glass (E-Gl, S-Gl) Low
High chemical resistance Low fatigue resistance

Good impact resistance Poor compressive properties


Kevlar (Kv) Moderate
High durability Very difficult to machine

Good fatigue resistance Low impact resistance


Graphite (Gr) High
Low thermal expansion High electrical conductivity

Table 2: Key characteristics of most used fibers in filamentary composites.

1.2.2 Laminated fiber-reinforced composites


We have already mentioned above that fibrous composites are most frequently man-
ufactured as thin laminae or plies. The fibers of a given ply are all aligned in a
certain (preferred) direction, so one could expect that the mechanical properties in
the transverse direction to the fibers are mainly dominated by the matrix behavior,
thereby the composite presenting lower strength and stiffness in the transverse direc-
tion to the fibers. This issue may be circumvented laiding up several unidirectionally

- 5-
1 Introduction to Composite Materials

Resins Cost Advantages Drawbacks


Good damage tolerance Sensitive to
Polyester (Pe) Low
High chemical resistance moisture content

Good mechanical performance Some sensitive to


Epoxy (Ep) Moderate
Good temperature resistance moisture content

Very high temperature resistance High curing


Polyimide (Pi) Moderate
Insensitive to moisture content temperature

Table 3: Key characteristics of most used matrix materials in filamentary compos-


ites.

reinforced layers, each one being characterized by its own orientation, in order to
conform what we call a laminated composite. This way, an overall reinforcement can
be obtained in the plane of the laminate. Figure 2 schematically shows an example
of a laminate built using various laminae with different orientations of the fibers.
Note a laminate is doubly composite, that is, it is a composite structure composed
by layers, each one of them being made up of a composite material.

The study and characterization of laminates (Chapters 3 and 4) is far more


reaching than the study of the mechanical behavior of a single lamina (Chapter 2).
Obviously, a comprehensive analysis of the latter case results necessary before the
former can be confronted. However, other important aspects, such as the under-
standing of the ultimate consequences of joining together several plies to obtain a
certain laminate with some optimum specifications, directly arise.
The mechanical properties of many-layered, cross-ply (with fiber directions ori-
ented at right angles relative to one another) laminates are presented in Table 4.
The properties of bidirectional laminates are used to provide a more fair comparison
with the properties of the isotropic bulk metals given in Table 1. The mechanical
advantages of this type of structural materials over metals are the usually greater
specific properties that they present due to the low density of the fiber and matrix
materials, mainly the latter. In this aspect the so-called advanced composites, formed
using Kevlar, graphite or boron, are superior in specific stiffness and strength to the
related properties of metals. Glass-fiber-epoxy composite presents still a high spe-
cific strength, but its specific elastic modulus is the lowest of all the listed materials,
including the steel and aluminium alloys. However, composites reinforced with glass
fiber are extensively used (although not in primary structures) because of their low
cost and no poor mechanical performance.
The fact that, for example, the specific stiffness may become an important design
parameter for certain structural applications is readily observed by means of a simple
calculation using a prismatic rod subjected to an axial load. From Strength of
Materials, it is well-known that the axial extension of the rod given in terms of the
load P , its length L, the cross-sectional area A and the Young’s modulus E is
PL
∆L = (1)
EA

- 6-
1.2 Types of composite materials

Figure 2: Unbonded layup of a symmetric cross-ply laminate.

Tensile Tensile Specific Specific


Material modulus strength Density modulus strength
system E (GP a) σu (GP a) ρ (g / cm3 ) E/ρ σu /ρ
E-Glass/Epoxy 21.5 0.57 1.97 10.9 0.26
Kevlar/Epoxy 40 0.65 1.40 29.0 0.46
Graphite/Epoxy 83 0.38 1.54 53.5 0.24
Boron/Epoxy 106 0.38 2.00 53.0 0.19

Table 4: Mechanical properties of cross-ply laminates. Adapted from Ref. [2].

Since the mass m of the rod is m = ρAL, where ρ is the material density, we obtain

P L2 1
m= (2)
∆L E/ρ

If the applied load P and the elastic elongation ∆L are prescribed for a given rod
(of certain length L), then the specific stiffness E/ρ emerges as the only material
parameter to be tailored to meet the design requirements. In this case, the higher
the specific stiffness, the lower the mass of the rod. Another simple calculation for
the buckling of this rod using the Euler column formula√leads to an expression for
the required mass m which is inversely proportional to E/ρ. From these results,
the importance of composite materials, whose specific properties are found to be
excellent, becomes evident for applications in which the mass of the structure is a
critical design parameter.

- 7-
1 Introduction to Composite Materials

Figure 3: Autoclave for industrial processes. ASC Process Systems. [Public Do-
main].

1.3 Fabrication and assembly


In this section, we briefly outline the usual procedure for manufacturing structural
elements made up with fibrous composite materials and enumerate the typical as-
sembly methods.
Fabrication of laminated composite elements typically include three basic oper-
ations. First, the needed quantity of plies are arranged, with the desired orientation
of the fibers, in a tooling that will control the laminate shape. Second, the set of
layers are introduced into a device known as autoclave, where pressure is applied to
compact the material and remove void-producing entrapped gases. Finally, heat is
applied to the structural element within the autoclave until the thermoset matrix
material is cured. An example of an autoclave, needed to perform this fabrication
process, is shown in Figure 3.
The final structural element may be obtained in a single autoclave operation
following the three previous steps. Alternatively, several laminated elements can be
separately precured in a first operation and subsequently bonded in a separate step
(known as secondary bonding) in order to complete the assembly of the composite
structure. Similarly to metals, although with different usage and design issues, cured
laminated elements can also be assembled using mechanical fasteners.

1.4 Advantages and disadvantages of laminated composites


The two main advantages of composite materials over conventional isotropic struc-
tural materials can already be inferred from several arguments introduced within
the foregoing sections. First, they present higher stiffness-to-density and strength-
to-density ratios than metals. We have seen that high specific properties can lead to

- 8-
1.5 Use of composite materials in industry

a substantial reduction of structural weight for a given, prescribed by designers, me-


chanical performance. Second, we have shown how an anisotropic laminate can be
conformed using several unidirectionally fiber-reinforced layers oriented in specific
directions. The advantage of this layup procedure lies in the so-called controlled
anisotropy, which let us tailor the final properties of the material to met the de-
sign requirements by merely changing the arrangement of the reinforcement. Other
additional advantages of composites are their low thermal expansion and the high
fatigue, corrosion and impact resistances they usually present.
However, there also exist some drawbacks and limitations in the use of fila-
mentary composites. These include high costs and other difficulties related to the
involved manufacturing procedures (including curing, bonding and joining), com-
plexities associated to the inherent anisotropic material behavior (which implies
additional economic and time costs related to the education of personnel), low levels
of fracture toughness and high sensitivity to moisture concentrations and to elevated
temperatures. Less possibilities of mass production than with traditional materials
is also a handicap for certain applications, specially consumer goods.

1.5 Use of composite materials in industry


The preceding, distinct advantages that fiber reinforced composites present over
other materials, along with the great advancements achieved during the last decades
in the related field of theoretical and applied mechanics, have motivated an increasing
use of this special type of materials in a wide variety of high-performance engineering
applications. They have found their place in, for example, the aircraft, aerospace
and military markets, the marine market, the sectors of appliances, consumer prod-
ucts and business equipment, the construction market, the sector of electrical and
electronic devices and, most recently, the realm of orthopedic surgery.
The 53% of the overall airframe of the new Airbus A350 XWB, Figure 4, is
comprised by composites. Other materials used include aluminium (19%), titanium
(14%) and steel (6%) alloys. These percentages dramatically compare to the ones
of the Airbus A330 (developed in the 1980s) which only comprises around a 10% of
composites in its airframe. The use of composites in the fuselage, wing and tail of
the A350 contributes to a lower aircraft weight coupled with increased resistance to
corrosion and fatigue during the jetliner’s lifetime. Even if the composite material
cost may be higher, the reduction in fuel costs (due to the reduction of weight) and
in maintenance expenses (due to the enhanced corrosion and fatigue resistances)
make it more profitable.
The restriction on the coefficient of thermal expansion for certain components
of satellites, which are subjected to extremely large ranges of temperatures, mo-
tivates the use of composite materials to build them. For example, an antenna
reflector entirely made of graphite fiber composite is shown in Figure 5. Conven-
tional engineering materials cannot really meet these strict requirements, consisting
of coefficients of thermal expansion in the order of 10−7 K−1 . For example, steel
presents typical values in the order of 10−5 K−1 .
The design of a forward-swept wing for a supersonic jet aircraft (prototype) such
as the Grumman X-29A, Figure 6, has several high-speed aerodynamic and ma-

- 9-
1 Introduction to Composite Materials

Figure 4: Low altitude pass of the first model of the A350 at the Air-
bus site in Toulouse. Photo by: Don-vip (own work) [CC-BY-SA-3.0
(http://creativecommons.org/licenses/by-sa/3.0)], via Wikimedia Commons.

Figure 5: Antenna reflector for spatial applications. Departamento de Aeronaves y


Vehı́culos Espaciales, ETSIAE, UPM, Madrid, Spain.

- 10-
1.5 Use of composite materials in industry

Figure 6: X-29 in Banked Flight. NASA Photo Number: EC87-0182-14. [Public


domain].

neuverability benefits. However, this type of wing presents important aeroelastic


problems, which have caused that no highly swept-forward jet fighters (a represen-
tative example being the Sukhoi Su-47 Berkut) have entered production yet. For
example, the so-called divergence may appear when an increment of bending at some
part of the wing (typically the tip in forward-swept wings) is accompanied by an
increment of twist (and the angle of attack) of the wing, both moments being orig-
inated by aerodynamic forces. Then, a greater angle of attack causes further lift,
further bending and further twist, and so on. This instability, more easily induced
in forward-swept wings, may increase bending up to the failure of the wing. It may
also initiate the stall at the wing tips and make the ailerons to lose control of the
airplane. In this aspect, we will see in Chapter 3 how antisymmetric laminated
composites allow the forward-swept wing twists (in the opposite sense to the sense
caused by the aerodynamic forces) as the wing bends, thereby preventing the prob-
lem of divergence and delaying the stall. Accordingly, we say that composites allow
aeroelastic tailoring of the swept-forward wing.
So varied applications as heat shields for spacecrafts and launch vehicles, wind
turbines, solar panel components, boat hulls, bicycle frames, racing and consumer
cars, pneumatic tires and implanted orthopedic components, among many others,
are partially or entirely manufactured with composites, hence taking advantage of
the special characteristics of these high-performance, wide-ranging engineering ma-
terials.

- 11-
1 Introduction to Composite Materials

- 12-
2 Mechanical Analysis of a Lamina
The materials introduced in the previous Chapter present mechanical properties
which are inherently direction-dependent. We study herein the mechanics associated
to this specific type of materials, known as anisotropic materials in the most general
case. Among all the possible material symmetries, we will pay more attention to the
special case of orthotropic materials. First, the description of the elastic behavior
of anisotropic bodies is presented using a macromechanical approach. That is, the
developed theory applies to materials with direction-dependent properties, indepen-
dently of the way in which the corresponding type of anisotropy is originated. Since
most composites are industrially produced in the shape of plies, these equations are
then particularized to the special case of thin laminae. Some macroscopic failure
criteria needed to study the behavior up to fracture of these type of solids are then
presented. In the final sections, the influence of the constituents characteristics into
the overall stiffness and strength properties of the composite material is investigated
using several micromechanical approaches.

2.1 Stress and strain tensors


2.1.1 Concept of stress
In Mechanics of Materials we study the response of solids to external actions such
as, for example, external forces as shown in Figure 7.a. To analyze the effect of these
loads acting over the solid, a generic free-body diagram, Figure 7.b, is considered,
where the shown external forces must be balanced by internal forces appearing inside
the body. These internal forces lead to the concept of stress when are associated to
a point of the solid and referred per unit area. In that sense, from Figure 7.b, the
normal stress
F int (A) dFnint
σn = lim n = (3)
A→0 A dA
with
Fnint = F int · n̂ kn̂k = 1 (4)
represents the normal force per unit area acting at the point P on its associated
surface A when the limit A → 0 is taken. In the previous expressions, n̂ is the
normal unit vector to the surface A.
Equivalently, the shear stress

Ftint (A) dFtint


σt = lim = (5)
A→0 A dA
with
Ftint = F int · t̂ t̂ = 1 (6)
stands for the shear (tangential) force per unit area acting at the point P on the
surface A → 0 along a given tangent direction t̂ to that surface (n̂ · t̂ = 0). Although
there are infinite tangent directions to the surface A, we will see in Section 2.1.3
that any shear stress acting over A can be expressed in terms of the shear values
associated to two mutually orthogonal tangent directions to that surface.

- 13-
2 Mechanical Analysis of a Lamina

a) b)
ext
F1
ext F4 F1
ext

int
ext ext P F
F2 F 2

F3
ext
F5ext F3
ext

Figure 7: a) Body subjected to several external loads F ext ext


1 , ..., F 5 . b) Free-body
diagram of the portion of the solid on one side of an imaginary cut showing a
representative internal force F int acting on a surface A associated to a given point
P.

2.1.2 Concept of strain

Within the framework of deformable solids, in clear contrast to classical mechanics,


external loads originate body deformations in addition to rigid solid motions. In
a similar way in which the stresses are defined, that is, locally, we introduce the
concept of strain to characterize the local deformation of the solid at a certain point
P . Again, this concept is better understood if longitudinal and shear strains are
separately introduced.
In Figure 8, the undeformed and (axially) deformed configurations of a variable
cross section rod are shown, with total lengths 0 L and L =0 L + ∆L, respectively.
In this case, the uniaxial deformation is non-uniform along the length of the rod.
Hence, in order to locally quantify the deformation, we introduce the dimensionless
longitudinal stretch measure

l dl
λ = lim = (7)
0 l→0 0 l d0 l

which represents the final length per unit length of the portion of the rod situated
at a certain point P . The longitudinal strain ε can be defined using the stretch λ
through
dl dl − d0 l
εl = λ − 1 = 0 − 1 = (8)
d l d0 l
which is a measure of the increment of length dl−d0 l of the longitudinal infinitesimal
element relative to its initial length d0 l. That is, εl represents the change of length
per unit length at P . For the special case in which the deformation is uniform along
the rod (just consider the rod of Figure 8 to be of constant cross section), both
the stretch measure and the strain measure take uniform values throughout the rod

- 14-
2.1 Stress and strain tensors

0
l

0
L ΔL
L
ext ext
F P F

l
Figure 8: Undeformed and deformed configurations of a variable cross section rod
under axial loading.

yielding
L
λ= 0L
(9)

and
L − 0L ∆L
εl = 0L
= 0 (10)
L
On the other hand, a measure of the shear deformation can be defined using
an infinitesimal element of area, as the ones depicted in Figure 9. Comparing the
reference (a) and the infinitesimal pure shear (b) configurations shown in that Figure,
the shear strain εt is defined as half the angular distortion undergone by the two
initially orthogonal edges of the element, namely
γ
εt = (11)
2
where γ represents the total angular distortion (or engineering shear strain) suffered
by the element. The total angular distortion γ (and hence the shear strain εt ) can be
defined in a more practical way from the configuration shown in Figure 9.c, known
as a simple shear state of deformation. Note in this last figure (c) how the right
edge “slides” with respect to the left edge. In the infinitesimal context (γ ≪ 1) the
pure shear and simple shear states only differ on a solid rigid rotation of γ/2 to a
first approximation. The shear strain εt is the same in both cases.

2.1.3 Tensorial nature of stresses and strains


Through the following chapters we are going to work with three-dimensional solids,
where stresses and strains become tensorial entities. In order to define the stresses in
the three-dimensional space, we introduce the infinitesimal volume of Figure 10, rep-
resenting a point P of a solid, along with the Cartesian right-handed reference frame
Xref = Oxyz. Following the foregoing definitions for normal and shear stresses, Eqs.

- 15-
2 Mechanical Analysis of a Lamina

a) b) c)
y y y

x x x
Figure 9: Shear deformation. a) Initial configuration. b) Pure shear state. c) Simple
shear state.

(3) and (5), we can define both type of stresses associated to each face of the ele-
ment. For instance, for the face which normal vector is defined by the x-axis, i.e.
n̂x = ex , we use the tangent vectors t̂y = ey , in the direction of Oy, and t̂z = ez , in
the direction of Oz, and introduce the notation
σxx = (σn )x σxy = (σt )y σxz = (σt )z (12)
where the first subscripts (of the left hand sides) indicate the surface in which the
stresses are being defined and the second subscript indicates the direction in which
each stress is acting. Obviously, for a normal stress both subscripts are equal and
for a shear stress they are different. Using this notation, it is easily shown from
balance of forces that
σxx = σ(−x)(−x) σxy = σ(−x)(−y) σxz = σ(−x)(−z) (13)
so only three stress values are needed to be defined for each normal direction (i.e., no
matter the sense). Analogously, we can define the stresses σyy , σyx and σyz acting
on the face with normal n̂y = ey and the stresses σzz , σzx and σzy acting on the
face with normal n̂z = ez , so a total of nine stress components are being introduced.
The enforcement of balance of moments of the unit cube subjected to these stress
components gives the additional relations
σxy = σyx σxz = σzx σyz = σzy (14)
Thus, only six stress components (of the nine being considered) acting at the point
P and defined in the system of representation Xref result to be independent.
Focusing again on the face of the infinitesimal cube which normal is ex , we note
that Eqs. (3) and (5) can be written in vectorial form as
F int
x
tx = lim (15)
Ax →0 Ax

where tx = {σxx , σxy , σxz }T (the superscript T means transpose) is called the trac-
tion vector associated to the face which normal is ex and represents the force per

- 16-
2.1 Stress and strain tensors

z
σzz σzy
σzx σyz
σxz y
σyy
σyx
σxy
σxx
x
Figure 10: Infinitesimal cube associated to a certain point P of a body and to the
reference frame Xref = Oxyz. Corresponding components of stress.

unit area acting at the point P on the surface Ax → 0. Obviously, the previous
choice of the orientation of the reference frame Xref has been completely arbitrary,
so one can generalize this last vectorial expression for any other generic direction n̂
to give
F int
n
tn = lim (16)
An →0 An
with its meaning being easily understood from the previous definitions. It follows
from Appendix A that any traction vector tn can be calculated in terms of the
six independent stress components σxx , σyy , σzz , σxy , σyz and σzx , contained in the
traction vectors tx , ty and tz , through the fundamental relation (Cauchy’s postulate)
tn = σ · n̂ (17)
where  
σxx σxy σzx
[σ]Xref =  σxy σyy σyz  (18)
σzx σyz σzz
is the representation of the second-order symmetric stress tensor σ in the axes
Xref . That way the considered six stress components at the given point (or other
six components relative to other arbitrary reference frame) completely define the
stress state at that point. The tensorial character of σ arises from the consideration
that Eq. (17), with n̂ fixed, holds for any orientation of the system of representation
Xref and its associated stress components.
With respect to the strain field at the point P , it can be deduced that uni-
axial and shear strains can be defined in a unified (tensorial) expression from the
displacement field in the continuum. In Figure 11, the position of a point P in
the undeformed, 0 xP , and deformed, xP , configurations of the solid is represented.
These two positions are related by means of the displacement uP , i.e.
xP = 0 xP + uP (19)

- 17-
2 Mechanical Analysis of a Lamina

uP
P
P

0
xP
xP
z
O
x y

Figure 11: The reference and current configurations of a body and definition of the
displacement vector uP associated to a point P .

We can write this last equation for a generic point 0 x of the reference solid as
 
x 0 x, t = 0 x + u 0 x, t (20)

where the temporal dependence is also considered.


Consider now the undeformed and deformed infinitesimal volumes associated to
a given point P being related by the displacement field and projected onto, for
example, the plane Oyz, as in Figure 12. In that figure, departing from Eqs. (8)
and (11) with d0 ly = dy and d0 lz = dz, we obtain

dly − dy uy (0, dy, 0) − uy (0, 0, 0) ∂uy


εyy ≈ ≈ = (21)
dy dy ∂y
dlz − dz uz (0, 0, dz) − uz (0, 0, 0) ∂uz
εzz ≈ ≈ = (22)
dz dz ∂z
γyz γ2 + γ1 1
εyz = = ≈ (tan γ2 + tan γ1 ) (23)
2 2 2 
1 uy (0, 0, dz) − uy (0, 0, 0) uz (0, dy, 0) − uz (0, 0, 0)
≈ + (24)
2 dz dy
 
1 ∂uy ∂uz
= + (25)
2 ∂z ∂y

which leads to the definition of the following second-order symmetric strain tensor
!    !T 
0
∂u x, t 0 0
 1 ∂u x, t ∂u x, t
ε 0 x, t = sym 0
=  0
+  (26)
∂ x 2 ∂ x ∂0x

Thus, we can formally calculate the strains at certain point P from the material
gradient of the displacement field u 0 x, t . Note the physical interpretation of

- 18-
2.2 Constitutive equations for isotropic solids

z
uy(0,0,dz)
dz+uz(0,0,dz)
z γ2
dlz

dly
dz
uz(0,0,0) γ1 uz(0,dy,0)
y
dy y uy(0,0,0) dy+uy(0,dy,0)

Figure 12: Engineering strains from the displacement field.

longitudinal and shear strains considered in Eqs. (21)–(25) is only included in Eq.
(26) when infinitesimal deformations, i.e. |u| ≪ 1, are considered. Hence, we will
refer to ε as the infinitesimal or engineering strain tensor. The representation of
this tensor in the reference frame Xref of Figure 10 in effect yields
 
εxx εxy εzx
[ε]Xref =  εxy εyy εyz  (27)
εzx εyz εzz

with components

∂ux ∂uy ∂uz


εxx = εyy = εzz = (28)
∂x ∂y ∂z
     
1 ∂ux ∂uy 1 ∂uy ∂uz 1 ∂uz ∂ux
εxy = + εyz = + εzx = + (29)
2 ∂y ∂x 2 ∂z ∂y 2 ∂x ∂z
The symmetries εyx = εxy , εzy = εyz and εxz = εzx are apparent from these last
expressions.

2.2 Constitutive equations for isotropic solids


In the previous section we have introduced the Cauchy stress tensor and the in-
finitesimal strain tensor separately. The former provides a measure of the tensional
state at each point of the solid, whereas the latter gives a measure of the deforma-
tion state at each point. Obviously, at any instant t, both fields exist at the same
time within the deformed body, so one may wonder if there is a relationship be-
tween them. A constitutive equation is simply an equation that relates the stresses
appearing in the solid to the strains and other possible derived quantities (such as
strain rates, permanent strains, etc.). From experimental observation and a deep
theoretical background, we can define constitutive models for any real material.

- 19-
2 Mechanical Analysis of a Lamina

As we have mentioned during the definition of the strain tensor ε, we restrict


the deformations to be infinitesimal. Within this context, a generic constitutive
equation for a elastic material relates stresses and strains through

σ = f (ε) (30)

where f is a non-linear tensor-valued function of the strain tensor ε. If the func-


tion f is known for a specific material model, then we can directly calculate the
corresponding stress state for each specified strain state (and vice versa, using the
inverse function of f ). From now on, we will deal with hyperelastic materials, which
are elastic materials for which there exists an stored energy function of the strains
w (ε), per unit volume, such that the stresses can be directly derived from w (ε) by
means of
∂w (ε)
σ (ε) = (31)
∂ε
Aside, it can be observed from experimental testing over composite materials that
the components of stress are linear functions of the components of strain (within a
specific range), so the preceding non-linear equations can be formally linearized to
give
σ = C : ε (32)
z }| { z }| { z }| {
3×3 3×3×3×3 3×3
where the fourth-order tensor

∂σ (ε) ∂ 2 w (ε)
C= = (33)
∂ε ∂ε∂ε

is called the stiffness tensor and does not depend on ε for linear, (hyper-)elastic
materials. The symbol : in Eq. (32) stands for the double-dot product, in this
case, of a fourth-order tensor and a second-order tensor. Although a fourth-order
tensor may have 3 × 3 × 3 × 3 = 81 independent components, the consideration of
the symmetries of σ and ε (each one with 6 independent components) in Eq. (32)
implies that C has only 36 independent stiffness components. That is, the stiffness
tensor C has the two minor symmetries Cijkl = Cjikl and Cijkl = Cijlk . In addition,
from Eq. (33)
∂2w ∂2w
Cijkl = = = Cklij (34)
∂εij ∂εkl ∂εkl ∂εij
and the stiffness tensor C has also the major symmetry Cijkl = Cklij , which reduces
the number of independent components of C to 21. The linear relation Eq. (33)
indicates that for linear elastic materials, the strain energy density w is a quadratic
function of the components of strain, namely

1 1
w (ε) = ε : C : ε = ε : σ (ε) (35)
2 2

The first linear elastic model we introduce corresponds to the case in which the
material has mechanical properties that are not direction-dependent, that is, the

- 20-
2.2 Constitutive equations for isotropic solids

material is isotropic. In that case, the relation between strains and stresses is given
by the well-known Hooke’s law for isotropic materials
1
εxx = [σxx − ν (σyy + σzz )] (36)
E
1
εyy = [σyy − ν (σzz + σxx )] (37)
E
1
εzz = [σzz − ν (σxx + σyy )] (38)
E
σxy σyz σzx
γxy = γyz = γzx = (39)
G G G
in terms of the Young’s (tensile) modulus E, the Poisson’s ratio ν (which describes
the so-called Poisson effect or transverse-longitudinal coupling effect) and the shear
modulus G = E/[2 (1 + ν)]. The elastic properties E and ν can be easily measured
from a uniaxial tension test performed over a longitudinal specimen. For example,
if σxx 6= 0 and σyy = σzz = 0, then εxx = σxx /E and εyy = εzz = −νσxx /E =
−νεxx . The slope of the uniaxial stress-strain curve gives the tensile modulus E =
σxx /εxx and the slope of the transverse-longitudinal strains curve gives the value
ν = −εyy /εxx = −εyy /εxx . The inversion of Eqs. (36)–(39) provides the following
expressions for the stresses in terms of the strains

E
σxx = [(1 − ν) εxx + νεyy + νεzz ] (40)
(1 + ν) (1 − 2ν)
E
σyy = [(1 − ν) εyy + νεzz + νεxx ] (41)
(1 + ν) (1 − 2ν)
E
σxx = [(1 − ν) εzz + νεxx + νεyy ] (42)
(1 + ν) (1 − 2ν)

σxy = Gγxy σyz = Gγyz σzx = Gγzx (43)


These last stress-strain relations can be written in the compact, tensorial form

E Eν
σ= ε+ (trε) I (44)
1+ν (1 + ν) (1 − 2ν)

where tr (ε) = εxx + εyy + εzz = εv represents the volumetric dilatation and I is the
second-order identity tensor. At this point, note that Eq. (44) is an example of the
type of linear elastic models given in Eq. (32). It is apparent that the associated
stiffness tensor C = ∂σ/∂ε is uniquely determined if the material parameters E
and ν are given. Or, in other words, only 2 of the initially possible 21 independent
components of C actually result to be independent for an isotropic linear elastic
material.
Similarly, for general hyperelastic materials, there exists a complementary strain
energy per unit volume wc (ε) which is function of the stresses, such that the strains
can be directly derived from wc (ε) by means of

∂wc (σ)
ε (σ) = (45)
∂σ

- 21-
2 Mechanical Analysis of a Lamina

A further linearization of Eq. (45) yields


ε=S:σ (46)
where the fourth-order tensor
∂ε (σ) ∂ 2 wc (σ)
S= = (47)
∂σ ∂σ∂σ
is called the compliance tensor and does not depend on σ for linear elastic ma-
terials. The complementary strain energy density wc of this type of materials is a
quadratic function of the components of stress, namely
1 1
wc (σ) = σ : S : σ = σ : ε (σ) (48)
2 2

Finally, note that Eqs. (36)–(39) are a particular case of the general linear case
given in Eq. (46) since they can be written in tensorial notation as
1+ν ν
ε= σ − (trσ) I (49)
E E
where tr (σ) /3 = p is the so-called hydrostatic pressure. Like C, the constitutive
tensor S = ∂ε/∂σ = C−1 only depends on two independent material parameters (E
and ν, or any other combination of two elastic parameters) for isotropic materials.

2.2.1 Engineering (Voigt) notation


In order to operate with equations such as Eq. (32) or Eq. (46), which combine
second-order and fourth-order tensors, they can be written in matrix form. To do
that, the 9 components of the second-order tensors σ or ε may be arranged in
a 9 × 1 matrix (vector), whereas the 81 components of C or S may be arranged
in a 9 × 9 matrix. However, attending to the aforementioned symmetries of all
these tensors, only 6 components of the second-order tensors are independent and
only 21 components of the fourth-order tensors are independent, which considerably
reduces the size of the resultant matrix equations. Hence, considering a reference
frame X = Oxyz (we will omit subscripts of this type in matrix notation when no
confusion is possible), we can write for σ and ε
       

 σ 1 
 
 σ xx 
 
 ε1 
 
 εxx 


 σ 
 
 σ 
 
 ε 
 
 ε 


 2 
 
 yy 
 
 2 
 
 yy 

       
σ3 σzz ε3 εzz
{σ} = = {ε} = = (50)

 σ4  
 σxy  
 ε4 
 
 γxy 
       
 σ     σyz   ε    γyz 
 5   5 

   
   
     

σ6 σzx ε6 γzx
where the index pairs {xx, yy, zz, xy, yz, zx} have been collapsed to {1, 2, 3, 4, 5, 6},
i.e.
xx ↔ 1
yy ↔ 2
zz ↔ 3
(51)
xy ↔ 4
yz ↔ 5
zx ↔ 6

- 22-
2.2 Constitutive equations for isotropic solids

Accordingly, C is represented in matrix notation as


 
C11 C12 C13 C14 C15 C16

 C12 C22 C23 C24 C25 C26 

 C13 C23 C33 C34 C35 C36 
[C] =   (52)

 C14 C24 C34 C44 C45 C46 

 C15 C25 C35 C45 C55 C56 
C16 C26 C36 C46 C56 C66

and S as  
S11 S12 S13 S14 S15 S16

 S12 S22 S23 S24 S25 S26 

 S13 S23 S33 S34 S35 S36 
[S] =   (53)

 S14 S24 S34 S44 S45 S46 

 S15 S25 S35 S45 S55 S56 
S16 S26 S36 S46 S56 S66
Note that the matrices [C] and [S] are symmetric. Then, we can write

{σ} = [C] {ε} {ε} = [S] {σ} [C] = [S]−1 (54)

Using this practical (engineering or Voigt) notation, it is straightforward to


rewrite Eqs. (40)–(43) as
 
  C11 C12 C12 0 0 0  

 σ1 

 C12 C11 C12 0 0 0 
 ε1 

 σ2   ε2 





 C12 C12 C11 0 0 0 
 

   C11 − C12 
 

σ3  ε3
= 0 0 0 0 0 
 σ4   2  ε4 

   C11 − C12 





 σ 5


 
 0 0 0 0 0 

 ε5 



σ6
  2  ε6

C11 − C12
0 0 0 0 0
2
(55)
with
(1 − ν) E νE
C11 = C12 = (56)
(1 + ν) (1 − 2ν) (1 + ν) (1 − 2ν)
C11 − C12 E
= =G (57)
2 2 (1 + ν)
and Eqs. (36)–(39) as
    

 ε1 
 S11 S12 S12 0 0 0 
 σ1 


 ε 
  S12 S11 S12 0 0 0  σ2 


 2 
  
   
 

ε3  S12 S12 S11 0 0 0 σ3
= 

 ε4 


 0 0 0 2 (S11 − S12 ) 0 0 
 σ4 

   


 ε5



 0 0 0 0 2 (S11 − S12 ) 0 

 σ5 


   
ε6 0 0 0 0 0 2 (S11 − S12 ) σ6
(58)

- 23-
2 Mechanical Analysis of a Lamina

with
1 ν 2 (1 + ν) 1
S11 = S12 = − 2 (S11 − S12 ) = = (59)
E E E G
As mentioned above, we observe that both [C] and [S] have only two independent
components (C11 , C12 and S11 , S12 , respectively) which are defined in terms of two
independent elastic constants, i.e. E and ν in this case. Finally, we note that the
matrix expressions for [C] and [S] given just above, which have been directly obtained
from Eqs. (44) and (49), are the same whatever the orientation of the system of
representation X in which Eqs. (44) and (49) are projected, as it should be for an
isotropic (non-direction-dependent) material.

2.3 The linearized stress-strain behavior for anisotropic materials


A material which does not behave isotropically is said to be anisotropic. The me-
chanical properties of these materials are direction-dependent so, even though con-
sidering linear and elastic assumptions, we cannot write simple constitutive models
valid for any system of representation, like the one given in Eq. (44) (or Eq. (49)).
Instead, we have necessarily to deal with specific expressions of the fourth-order
tensors C and S for each considered basis or, more practically, with their associated
matrix expressions [C] and [S], namely Eqs. (52) and (53), with 21 independent
material parameters to be determined from mechanical tests. These stress-strain
relations will be applicable to the most general anisotropic case.

2.3.1 Monoclinic material


There are special cases of anisotropic materials for which, attending to material
symmetry considerations, the number of independent material parameters may be
reduced to a large extent. The first symmetry case we consider corresponds to a
monoclinic material. This material presents behavioral symmetry with respect to
a specific plane, whereupon the stress responses associated to any two strain states
being symmetric with respect to that plane have also to be symmetric. In order
to elucidate the number of independent parameters of these type of materials, we
consider the plane Oxy to be the symmetry plane of the monoclinic material. Then,
two symmetric deformation states characterized by the strain tensors ε and ε′ are
related by
     T
ε′xx ε′xy ε′zx 1 0 0 εxx εxy εzx 1 0 0
 ε′xy ε′yy ε′yz  =  0 1 0   εxy εyy εyz   0 1 0  (60)
ε′zx ε′yz ε′zz 0 0 −1 εzx εyz εzz 0 0 −1
 
εxx εxy −εzx
=  εxy εyy −εyz  (61)
−εzx −εyz εzz

where we define  
1 0 0
 0 1 0  = M xy (62)
0 0 −1

- 24-
2.3 The linearized stress-strain behavior for anisotropic materials

to be the improper orthogonal tensor that describes the symmetric mapping with
respect to the plane Oxy (see the footnote1 ). The associated responses can be
expressed in terms of the components of C and the components of ε through {σ} =
[C] {ε} (using Eqs. (52)) and (Eq. (50)2 )) and {σ ′ } = [C] {ε′ } (using Eqs. (52)) and
(61)). Then, we relate the components of σ and σ′ enforcing the same symmetry
condition which has been regarded for ε, i.e.

σ′ = M xy σM Txy (63)

or
 ′ ′ ′
  
σxx σxy σzx σxx σxy −σzx

 σxy ′
σyy ′ 
σyz =  σxy σyy −σyz  (64)
′ ′ ′
σzx σyz σzz −σzx −σyz σzz

and we obtain six different equations, expressed in terms of the components of C


and the components of ε, to be satisfied. If we consider that the components of ε
are completely arbitrary, it is straightforward to deduce that

C15 = C16 = C25 = C26 = C35 = C36 = C45 = C46 = 0 (65)

Hence, for a monoclinic material the stiffness matrix reads


 
C11 C12 C13 C14 0 0

 C12 C22 C23 C24 0 0 

 C13 C23 C33 C34 0 0 
[C] =   (66)

 C14 C24 C34 C44 0 0 

 0 0 0 0 C55 C56 
0 0 0 0 C56 C66

with 13 independent components (material parameters) defined in the reference


frame X, which plane Oxy is the plane of material symmetry. Alternatively, since
the symmetry of stresses with respect to the plane Oxy (cause) also implies the
symmetry of strains (effect) for a monoclinic material, we deduce that the same
entries must vanish for the compliance matrix [S]. In matrix notation it reads

 
S11 S12 S13 S14 0 0

 S12 S22 S23 S24 0 0 

 S13 S23 S33 S34 0 0 
[S] =   (67)

 S14 S24 S34 S44 0 0 

 0 0 0 0 S55 S56 
0 0 0 0 S56 S66

1
If v = [vx , vy , vz ]T denotes the components of a vector in the basis Xref , note that v ′ = M xy v =
[vx , vy , −vz ]T is the symmetric vector (mirror image) of v with respect to the plane Oxy. Thus, for
a second order tensor A, its symmetric tensor with respect to the plane Oxy is A′ = M xy AM Txy .

- 25-
2 Mechanical Analysis of a Lamina

Another, more intuitive, way to deduce which components of [S] vanish for a mon-
oclinic material consists in considering, for example, the stress vector
   

 σxx 
 
 σ 


 0 
 
 0 


 
 
 

   
0 0
σ= = (68)

 0 
 
 0 

   
 0







 0 


   
0 0

to be applied to the monoclinic differential volume of Figure 13. Then, the resulting
strains are, using the general compliance matrix of Eq. (53)
   

 εxx 
 
 S 11 σ 


 εyy 
 
 S σ 


 
 
 12 

   
εzz S13 σ
= {ε} = [S] {σ} = (69)

 γ xy 
  S14 σ 
 
   

 γ
 yz





 S 15 σ 


   
γzx S16 σ

However, since the applied stresses are symmetric relative to the plane Oxy, the
response must also be symmetric. The admissible deformation (preserving the cor-
responding symmetries) is also shown in Figure 13. We deduce that angular distor-
tions γyz and γzx cannot appear upon the application of the single stress component
σxx . Then, from Eq. (69), we directly obtain

S15 = S16 = 0 (70)

Analogous arguments lead to the results S25 = S26 = 0 (for the application of a
single stress component σyy 6= 0), S35 = S36 = 0 (for σzz 6= 0) and S45 = S46 = 0
(for σxy 6= 0) and we recover Eq. (67). Note that due to the lack of symmetry with
respect to the plane Oyz, we cannot ensure that the term S56 vanishes when a single
stress σyz 6= 0 is applied to the monoclinic differential cube.

2.3.2 Orthotropic material

A material with two orthogonal planes of symmetry is called an orthotropic material.


If, for example, we choose the planes Oxy and Oyz to define these symmetry planes,
then it can be inferred from the composite transformation
    
1 0 0 −1 0 0 −1 0 0
M xy M yz =  0 1 0   0 1 0  =  0 1 0  = −M zx (71)
0 0 −1 0 0 1 0 0 −1

that if there exist two orthogonal planes of material symmetry, then a third plane
mutually orthogonal to both initial planes is also a plane of symmetry, the plane
Ozx in this case. We can prove this assertion by means of the successive application

- 26-
2.3 The linearized stress-strain behavior for anisotropic materials

z
z

σ y
y x
x
σ

Figure 13: Deformation undergone by a monoclinic material (with symmetry plane


Oxy) when a single stress component σxx is applied. Axial (εxx ), transverse (εyy ,
εzz ) and in-plane (γxy ) strains may be caused. Out-of-plane shear deformation (γyz ,
γzx ) is not possible.

of the symmetric mappings M yz and M xy to a generic strain tensor ε to obtain


that

ε′ = M xy M yz εM Tyz M Txy (72)
= (M xy M yz ) ε (M xy M yz )T (73)
T
= (−M zx ) ε (−M zx ) (74)
= M zx εM Tzx (75)

which represents a symmetry mapping of the tensor ε with respect to the plane
Ozx. By premise, the planes Oxy and Oyz are planes of symmetric behavior, so the
same successive symmetries apply on the stress response σ associated to ε. Hence
σ ′ = M zx σM Tzx and the stress tensors σ and σ′ are symmetric with respect to the
plane Ozx as well, Q.E.D. Note that −M zx also represents a rotation of π radians
around the axis Oy (i.e. −M zx = Qπy , see Appendix B), so the expression of the
symmetric mapping of ε with respect to the plane Ozx can also be obtained from
the rotation of ε an angle of π radians about the y-axis, that is

ε′ = M zx εM Tzx (76)
T
= (−M zx ) ε (−M zx ) (77)
T
= Qπy ε Qπy (78)
which is a useful interpretation. For an orthotropic material, the normal vectors
to the three mutually orthogonal planes define the so-called orthotropy preferred
(or principal) directions of the material. These directions are aligned with the axes
of the system of reference Oxyz. Using the transformation tensor M yz instead of
M xy and following an analogous procedure to that which led to the result given in
Eq. (65) for monoclinic materials, we additionally obtain for an orthotropic material
that
C14 = C24 = C34 = C56 = 0 (79)

- 27-
2 Mechanical Analysis of a Lamina

so the stiffness matrix for this type of materials reduces to


 
C11 C12 C13 0 0 0
 C12 C22 C23 0 0 0 
 
 C13 C23 C33 0 0 0 
[C] =  0
 (80)
 0 0 C44 0 0 

 0 0 0 0 C55 0 
0 0 0 0 0 C66

which contains 9 independent components (material parameters) defined in the pre-


ferred reference frame Oxyz. Again, the same entries vanish for the compliance
matrix [S] of an orthotropic material, which is
 
S11 S12 S13 0 0 0
 S12 S22 S23 0 0 0 
 
 S13 S23 S33 0 0 0 
[S] = 
  (81)
 0 0 0 S 44 0 0 

 0 0 0 0 S55 0 
0 0 0 0 0 S66

For further use, the components of [C] can be expressed in terms of the components
of [S] using Eq. (54)3
 
2
S22 S33 − S23 S13 S23 − S12 S33 S12 S23 − S13 S22
 0 0 0 
 S S 2
S 
 S 11 S 33 − S 13 S 12 S 13 − S 23 S 11 
 0 0 0 

 S S 2


 S11 S22 − S12 
 0 0 0 
[C] =  S 
 1 
 0 0 

 S44 

 1 
 sym 0 
 S55 
 1 
S66
(82)
where
 
S11 S12 S13
S = det  S12 S22 S23  (83)
S13 S23 S33
2 2 2
= S11 S22 S33 − S11 S23 − S22 S13 − S33 S12 + 2S12 S13 S23 (84)

The admissible response (preserving the corresponding symmetries) resulting from


the application of a single stress component σxx 6= 0 to an infinitesimal orthotropic
volume are shown in Figure 14. Evidently, the angular distortions γxy , γyz and γzx
are not possible, i.e. S14 = S15 = S16 = 0, which is in correspondence with Eq. (81);
and so forth. Note that, now, due to the symmetric behavior with respect to the
plane Oyz, we can ensure that the term S56 vanishes when a single stress σyz 6= 0 is
applied to the orthotropic differential cube.

- 28-
2.3 The linearized stress-strain behavior for anisotropic materials

z z

σ
y y

x σ x

Figure 14: Deformation undergone by an orthotropic material when a single stress


component σxx is applied. Axial (εxx ) and transverse (εyy , εzz ) strains are caused.
No shear deformation (γxy, γyz , γzx ) is possible.

2.3.3 Transversely isotropic material

Now we consider the specific group of transversely isotropic materials, which have
infinite planes of symmetry characterized by the set of symmetric mappings, see
Figure 15,
 
1 0 0
M α =  0 cos 2α sin 2α  (85)
0 sin 2α − cos 2α

where α ∈ [−π/2, π/2] (note M 0 ≡ M xy and M π/2 ≡ M zx ). This fact explains


the denomination of transverse isotropy given to this type of materials, since there
exists isotropy with respect to all the directions contained in the (transverse) plane
Oyz. After some algebra, it can be shown that this symmetry group is equivalent
to a combination of a rotation of the type of Qπy (indicated above) along with a
symmetry group containing rotations about the axis Ox

 
1 0 0
Qθx =  0 cos θ − sin θ 
0 sin θ cos θ

where θ is arbitrary. For this last reason, we also say that the x-axis is the axis of
symmetry (or preferred direction) of the transversely isotropic material. Following
analogous steps as the ones performed for monoclinic and orthotropic materials, the
following relations between the components of [C] are obtained

C22 − C23
C22 = C33 C12 = C13 C55 = C44 = C66 (86)
2

- 29-
2 Mechanical Analysis of a Lamina

x
Figure 15: Planes of material symmetry for a transversely isotropic material. α ∈
[−π/2, π/2]. Note the plane Oyz is also a symmetry plane.

and the stiffness matrix for transversely isotropic materials [C] reads
 
C11 C12 C12 0 0 0
 C12 C22 C23 0 0 0 
 

 C12 C23 C22 0 0 0 

[C] =  0 0 0 C44 0 0  (87)
 
 C22 − C23 
 0 0 0 0 0 
2
0 0 0 0 0 C44

with 5 independent components (material parameters) associated to the preferred


reference frame Oxyz (or any other preferred frame rotated an arbitrary angle about
the x-axis). We remark that the relation C55 = (C22 − C23 ) /2 simply manifests
the isotropy condition within the plane Oyz (compare with Eq. (55)). Analogous
arguments lead to the following compliance matrix
 
S11 S12 S12 0 0 0

 S12 S22 S23 0 0 0 

 S12 S23 S22 0 0 0 
[S] =   (88)

 0 0 0 S44 0 0 

 0 0 0 0 2 (S22 − S23 ) 0 
0 0 0 0 0 S44

for transversely isotropic materials, where now S55 = 2 (S22 − S23 ) (compare with
Eq. (58)).
Finally, if we consider a material in which all the planes that pass through a
point are planes of symmetry, then the material is said to be isotropic in that point
(it is no longer anisotropic). Equations (55) and (58) are recovered for this simple
case, with only two material parameters characterizing the elastic response of the
material.

- 30-
2.4 Apparent material constants. Coupling coefficients.

2.4 Apparent material constants. Coupling coefficients.


The stiffness [C] and compliance [S] matrices obtained in the previous section for
anisotropic materials (monoclinic, orthotropic and transversely isotropic materials
being special anisotropic cases) have been expressed in some specific, preferred or
principal, systems of reference, denoted as X in each studied case. From Eq. (54),
these matrices clearly relate the stress {σ} to the strain {ε} when both entities are
also expressed in the reference frame X. However, as we will see through the next
chapter, there are very usual situations for which the material preferred directions
of a lamina are not aligned with the coordinate system in which the mechanical
problem is analyzed, X ′ say. For these cases, the expressions of [C] and [S] in the
coordinate system used as reference, i.e. X ′ , are needed.

2.4.1 Apparent material constants


It follows from Appendix B that a change of basis of the second-order tensor σ from
X = Oxyz to X ′ = Ox′ y ′ z ′ , where X ′ is clockwise rotated an angle α about the Oz
axis with respect to X, see Figure 17 below, is obtained through
[σ]X ′ = [r]X→X ′ [σ]X [r]TX→X ′ (89)
where [r]X→X ′ is the orthogonal matrix of the corresponding transformation rule
from X to X ′
   
ex′ · ex ex′ · ey ex′ · ez cos α − sin α 0
[r]X→X ′ =  ey′ · ex ey′ · ey ey′ · ez  =  sin α cos α 0  (90)
ez ′ · ex ez ′ · ey ez ′ · ez 0 0 1
It is readily obtained from a simple terms identification exercise, that the result of
Eq. (89) can be rewritten in engineering notation as
{σ}X ′ = [Rσ ]X→X ′ {σ}X (91)
where
 
cos2 α sin2 α 0 −2 cos α sin α 0 0

 sin2 α cos2 α 0 2 cos α sin α 0 0 

 0 0 1 0 0 0 
[Rσ ]X→X ′ = 

 cos α sin α − cos α sin α 0 cos2 α − sin2 α 0 0 

 0 0 0 0 cos α sin α 
0 0 0 0 − sin α cos α
(92)
Due to the use of total angular distortions γij in the strain vector {ε}, the transfor-
mation matrix for {ε}, namely [Rε ]X→X ′ , is slightly different to [Rσ ]X→X ′ , i.e.
 
cos2 α sin2 α 0 − cos α sin α 0 0

 sin2 α cos2 α 0 cos α sin α 0 0  
 0 0 1 0 0 0 
[Rε ]X→X ′ = 

2 2

 2 cos α sin α −2 cos α sin α 0 cos α − sin α 0 0  
 0 0 0 0 cos α sin α 
0 0 0 0 − sin α cos α
(93)

- 31-
2 Mechanical Analysis of a Lamina

so

[ε]X ′ = [r]X→X ′ [ε]X [r]TX→X ′ (94)


{ε}X ′ = [Rε ]X→X ′ {ε}X (95)

Then, operating on Eq. (54)1

{σ}X = [C]X {ε}X (96)


[Rσ ]−1 −1
X→X ′ {σ}X ′ = [C]X [Rε ]X→X ′ {ε}X ′ (97)
{σ}X ′ = [Rσ ]X→X ′ [C]X [Rε ]−1
X→X ′ {ε}X ′ (98)
{σ}X ′ = [C]X ′ {ε}X ′ (99)

we are able to identify the stiffness matrix [C] expressed in the coordinate system
X ′ , i.e.

[C]X ′ = [Rσ ]X→X ′ [C]X [Rε ]−1


X→X ′ = [Rσ ]X→X ′ [C]X [Rε ]X ′ →X (100)

which is uniquely determined for a specific matrix [C]X and an angle α. Note that
[Rε ]X ′ →X = [Rε ]−1 T
X→X ′ 6= [Rε ]X→X ′ , that is, the matrix [Rε ]X ′ →X is not orthogonal
(neither [Rσ ]X ′ →X ).
We are mostly interested in the treatment of orthotropic laminae. In order
to understand this interest, in Figure 16, a unidirectionally reinforced lamina is
represented. For such a lamina, it can be assumed from first observation that the
Ox axis (reinforced direction) represents the preferred direction of a transversely
isotropic material, the plane Oyz representing the isotropic plane. However, firstly,
the different dimensions of the lamina in the Oy and Oz directions may generate
slight differences in the overall mechanical properties between both directions and, in
the second instance, attending to fabrication issues this type of laminae are usually
compacted in the Oz direction during the curing process, making the differences
between directions Oy and Oz more noticeable. Hence, we justify the treatment of
this kind of reinforced laminae as an orthotropic material in the three-dimensional
space. Obviously, a transversely isotropic behavior could arguably be accepted as a
good enough approximation in some applications.
For composite laminae, Eq. (80) gives the expression for the stiffness matrix
in principal material directions [C]X . Subsequently, Eq. (100) yields the following
matrix [C]X ′ expressed in a reference frame X ′ rotated an angle α about the Oz axis,
as shown in Figure 17 (axes Ox′ and Oy ′ remain within the plane of the lamina)
 
C11′ C12′ C13′ C14′ 0 0

 C12′ C22′ C23′ C24′ 0 0 

 C13′ C23′ C33′ C34′ 0 0 
[C]X ′ =  (101)

 C14′ C24′ C34′ C44′ 0 0 

 0 0 0 0 C55′ C56′ 
0 0 0 0 C56′ C66′

- 32-
2.4 Apparent material constants. Coupling coefficients.

x y

Figure 16: Unidirectionally fiber-reinforced lamina and principal material axes X.

with non-vanishing components

C11′ = C11 cos4 α + 2 (C12 + 2C44 ) sin2 α cos2 α + C22 sin4 α (102)
4 2 2 4
C22′ = C11 sin α + 2 (C12 + 2C44 ) sin α cos α + C22 cos α (103)
C33′ = C33 (104)

C44′ = (C11 + C22 − 2C12 − 2C44 ) sin2 α cos2 α + C44 sin4 α + cos4 α (105)
2 2
C55′ = C55 cos α + C66 sin α (106)
2 2
C66′ = C55 sin α + C66 cos α (107)
2 2 4 4

C12′ = (C11 + C22 − 4C44 ) sin α cos α + C12 sin α + cos α (108)
2 2
C13′ = C13 cos α + C23 sin α (109)
C14′ = (C11 − C12 − 2C44 ) sin α cos3 α − (C22 − C12 − 2C44 ) sin3 α cos α (110)
2 2
C23′ = C23 cos α + C13 sin α (111)
C24′ = (C11 − C12 − 2C44 ) sin3 α cos α − (C22 − C12 − 2C44 ) sin α cos3 α (112)
C34′ = (C13 − C23 ) sin α cos α (113)
C56′ = (C66 − C55 ) sin α cos α (114)
—for the matter of simplicity in the notation, we have written C11′ instead of
C(11)′ ≡ C1′ 1′ , and so on. Thus, note that in this non-principal coordinate system,
the stiffness matrix adopts the form of the monoclinic stiffness matrix in principal
material axes, Eq. (66), and that the number of material parameters has apparently
increased from 9 to 13. However, with the angle α being known, the non-vanishing
13 Cij ′ components are uniquely determined from the 9 Cij independent preferred
components through Eqs. (102)–(114), so the number of independent material pa-
rameters remains to be 9. The 13 matrix components in the working frame Ox′ y ′ z ′
are known as apparent material constants associated to X ′ . In the most general case,
for a three-dimensional rotation of the system of reference (not only about the Oz
axis), 21 apparent components should be considered for an orthotropic material. As
before, they could always be related to nine independent quantities. Regarding this

- 33-
2 Mechanical Analysis of a Lamina

y
y
x

Figure 17: Reference frames X and X ′ for a unidirectionally reinforced lamina. The
auxiliary axes X ′ are clockwise rotated an angle α about the z-axis with respect to
the principal material axes X.

z ≡ z z ≡ z

y
x ≡ x y ≡ y x y
x

Figure 18: Left: Specially orthotropic lamina. Right: Generally orthotropic lamina.

issue, an orthotropic lamina with its principal material directions X aligned with the
working basis X ′ is called a specially orthotropic lamina, while an orthotropic lamina
expressed in non-principal coordinates is known as a generally orthotropic lamina
(see Figure 18). It is remarkable that if we do not recognize the orthotropic pre-
ferred directions of a generally orthotropic lamina (containing 21 non-independent
components Cij ′ in the most general case) from direct observation, then we cannot
differentiate it from a truly anisotropic lamina (with 21 independent components
Cij ) from a mechanical standpoint.
The components of the compliance matrix [S] transform in a slightly different
way than the components of [C] when the system of representation is rotated about
the z-axis, as shown in Figure 17. Operating over Eq. (54)2

{ε}X = [S]X {σ}X (115)


[Rε ]−1
X→X ′ {ε}X ′ = [S]X [Rσ ]−1
X→X ′ {σ}X ′ (116)
{ε}X ′ = [Rε ]X→X ′ [S]X [Rσ ]−1
X→X ′ {σ}X ′ (117)
{ε}X ′ = [S]X ′ {σ}X ′ (118)

we are able to identify the compliance matrix [S] expressed in the coordinate system

- 34-
2.4 Apparent material constants. Coupling coefficients.

X ′ , i.e.

[S]X ′ = [Rε ]X→X ′ [S]X [Rσ ]−1


X→X ′ = [Rε ]X→X ′ [S]X [Rσ ]X ′ →X (119)

which is uniquely determined for a specific matrix [S]X and an angle α. The non-
vanishing (apparent) components of [S]X ′ , expressed in terms of the (principal) com-
ponents of [S]X and the angle α are

S11′ = S11 cos4 α + (2S12 + S44 ) sin2 α cos2 α + S22 sin4 α (120)
S22′ = S11 sin4 α + (2S12 + S44 ) sin2 α cos2 α + S22 cos4 α (121)
S33′ = S33 (122)
2 2 4 4

S44′ = 2 (2S11 − 4S12 + 2S22 − S44 ) sin α cos α + S44 sin α + cos α (123)
S55′ = S55 cos2 α + S66 sin2 α (124)
2 2
S66′ = S55 sin α + S66 cos α (125)
2 2 4 4

S12′ = (S11 + S22 − S44 ) sin α cos α + S12 sin α + cos α (126)
2 2
S13′ = S13 cos α + S23 sin α (127)
3 3
S14′ = (2S11 − 2S12 − S44 ) sin α cos α + (2S12 − 2S22 + S44 ) sin α cos α (128)
S23′ = S23 cos2 α + S13 sin2 α (129)
3 3
S24′ = (2S11 − 2S12 − S44 ) sin α cos α + (2S12 − 2S22 + S44 ) sin α cos α (130)
S34′ = 2 (S13 − S23 ) sin α cos α (131)
S56′ = (S66 − S55 ) sin α cos α (132)

which are to be compared with the expressions given in Eqs. (102)–(114).

2.4.2 Coupling coefficients

In the spirit of the well-known strain-stress response for isotropic materials, engi-
neeringly represented by the constants E (extension behavior), ν (transverse-axial
extension coupling) and G (shear behavior) —cf. Eqs. (58) and (59)—, we can de-
scribe the 21 components of a fully anisotropic material in terms of other additional
engineering parameters. To do that, we rewrite the matrix [S] of an anisotropic
material, expressed in the coordinate system X, in the form

- 35-
2 Mechanical Analysis of a Lamina

 1 νyx νzx ηx,xy ηx,yz ηx,zx 


− −
 Ex Ey Ez Gxy Gyz Gzx 
 

 νxy 1 νzy ηy,xy ηy,yz ηy,zx 

 − − 
 Ex Ey Ez Gxy Gyz Gzx 
 
 νxz νyz 1 ηz,xy ηz,yz ηz,zx 
 − − 

 Ex Ey Ez Gxy Gyz Gzx 

[S] = 
 ηxy,x ηxy,y ηxy,z 1 µyz,xy µzx,xy

 (133)
 

 Ex Ey Ez Gxy Gyz Gzx 

 ηyz,x ηyz,y ηyz,z µxy,yz 1 µzx,yz 
 
 
 Ex Ey Ez Gxy Gyz Gzx 
 

 ηzx,x ηzx,y ηzx,z µxy,zx µyz,zx 1 

Ex Ey Ez Gxy Gyz Gzx

In order to understand these new material constants, assume that the stress com-
ponent σ1 of {σ} takes a value σ1 = σxx 6= 0 and that the remaining components of
{σ} are zero. The resulting strains are
 σxx 

 



 Ex 


 

 σ 
xx 

  
 −ν xy   
 εxx 


 E x


  ε xx 
     

 ε yy

 
 σxx  
 −ν xy ε xx










 −ν xz









εzz Ex −νxz εxx
= [S] {σ} = σxx  = (134)
 γxy 
   η
    ηxy,x εxx  
   xy,x   
 γyz 

 



 Ex  
 
 ηyz,x εxx 
 

   σ   
γzx 
 xx 
 η ε
zx,x xx

 ηyz,x 



 Ex  

 

 σxx 
 ηzx,x
 

Ex

which indicates that for a single tensile loading, the anisotropic material undergoes
an axial extension (εxx = σxx /Ex ), transverse extensions (εyy /εxx = −νxy and
εzz /εxx = −νxz ) and shear deformations (γxy /εxx = ηxy,x , γyz /εxx = ηyz,x and
γzx /εxx = γzx ). The remaining entries in [S] may be similarly identified assuming
other stress vectors with only one non-vanishing component. Hence,

Ex , Ey and Ez (135)

stand for the Young’s moduli in the Ox, Oy and Oz directions,

νij , i, j = {x, y, z} , i 6= j (136)

are the Poisson’s ratios, which characterize the transverse extension-axial extension
coupling,
Gxy , Gyz and Gzx (137)

- 36-
2.4 Apparent material constants. Coupling coefficients.

S11
S12 S22 sym
S13 S23 S33
S14 S24 S34 S44
S15 S25 S35 S45 S55
S16 S26 S36 S46 S56 S66

Figure 19: Physical significance of the components of the compliance matrix [S]. In
red: pure extension behavior. In black: transverse extension-axial extension cou-
pling. In blue: pure shear behavior. In magenta: shear deformation-axial extension
coupling. In green: shear-shear coupling.

represent the elastic shear moduli measured in the Oxy, Oyz and Ozx planes,

ηi,jk , i = {x, y, z} , jk = {xy, yz, zx} (138)

are the so-called Lekhnitskii’s coefficients of mutual influence of the first kind, rep-
resenting the axial extension-shear deformation coupling,

ηij,k , ij = {xy, yz, zx} , k = {x, y, z} (139)

are the Lekhnitskii’s coefficients of mutual influence of the second kind, characteriz-
ing the shear deformation-axial extension coupling and

µij,kl , ij, kl = {xy, yz, zx} , ij 6= kl (140)

are the so-called Chentsov’s coefficients, that are to shearing strains what Poisson’s
ratios are to transverse-axial strains. The components of [S] are accordingly labelled
in Figure 19. Following the symmetry of [S], note that some reciprocal relations
involving the foregoing coupling coefficients must be satisfied for a linear elastic
anisotropic material, i.e.
νij νji
= , i, j = {x, y, z} , i 6= j (141)
Ei Ej

ηij,k ηk,ij
= , ij = {xy, yz, zx} k = {x, y, z} (142)
Ek Gij
and
µij,kl µkl,ij
= , ij, kl = {xy, yz, zx} , ij 6= kl (143)
Gij Gkl
which may be used to determine the value of some engineering material constants
in terms of the independent, measured from experiments, material parameters.

- 37-
2 Mechanical Analysis of a Lamina

2.4.3 Particular case: Specially orthotropic lamina


Specifically, and for further use, the compliance matrix [S] of an orthotropic lamina
expressed in principal material coordinates in terms of the 9 independent, engineer-
ing material constants is
 1 νyx νzx 
− − 0 0 0
 Ex Ey Ez 
 νxy 1 νzy 

 − − 0 0 0 

 Ex Ey Ez 
 νxz νyz 1 
 − − 0 0 0 

[S] =  Ex Ey Ez 
 (144)
 1 
 0 0 0 0 0 

 Gxy 

 1 
 0 0 0 0 0 

 Gyz 

1
0 0 0 0 0
Gzx

where the reciprocal relations of Eq. (141) are to be used. Furthermore, Eq. (82)
gives the expression of the stiffness matrix of a specially orthotropic lamina in terms
of these useful and easy-to-measure parameters
 
1 − νyz νzy νyx + νyz νzx νzx + νyx νzy
0 0 0

 Ey Ez S Ey Ez S Ey Ez S 

 ν +ν ν 1 − ν ν ν + ν ν 
 xy xz zy xz zx zy xy zx 
 0 0 0 
 Ex Ez S Ex Ez S Ex Ez S 
 
[C] =  νxz + νxy νyz νyz + νxz νyx
 1 − νxy νyx  (145)
0 0 0 

 E E
x y S E E
x y S E E
x y S 

 

 0 0 0 Gxy 0 0 
 0 0 0 0 Gyz 0 
0 0 0 0 0 Gzx

with
 1 νyx νzx 
− −
 Ex Ey Ez 
 ν 1 ν 
 xy zy 
S = det  − −  (146)
 Ex Ey Ez 
 νxz νyz 1 
− −
Ex Ey Ez
1 − νxy νyx − νyz νzy − νzx νxz − 2νyx νzy νxz
= (147)
Ex Ey Ez

and where the reciprocal relations resulting from the following symmetries

C12 = C21 , C23 = C32 , C31 = C13 (148)

evidently hold.

- 38-
2.4 Apparent material constants. Coupling coefficients.

If we go back to the aforementioned concept of apparent material constants, we


remark that for an orthotropic material, apparent Lekhnitskii and Chentsov coupling
coefficients may be measured for a completely generic orientation of the working co-
ordinate system (generally orthotropic material). However, for the specific case in
which the reference frame is aligned with the principal material directions (spe-
cially orthotropic material), these particular families of coupling coefficients vanish;
evidently the Poisson’s coupling effects remain (see Eq. (144)).

2.4.4 Restriction on engineering constants for an orthotropic material


For the linear infinitesimal theory, recall the strain energy per unit volume is given
by
1
w (ε) = ε : C : ε (149)
2
and the complementary strain energy density is given by
1
wc (σ) = σ : S : σ (150)
2
From a computational standpoint, note that for linear elastic materials
1 1 1 1
w= ε : (C : ε) = ε : σ = σ : ε = σ : (S : σ) = wc (151)
2 2 2 2
that is, the strain energy w takes the same numerical value that the complementary
strain energy wc for a given pair of associated strain and stress tensors (related by
σ = C : ε or ε = S : σ). The physical restriction on the stored energy w (equiva-
lently, the complementary strain energy wc ) to be positive for any deformation state
represented by the strain field ε (equivalently, σ) imposes some constraints to be
satisfied by the components of C (equivalently, S). From a mathematical standpoint,
those restrictions are satisfied if both matrices [C] and [S] are positive-definite. Since
[C] = [S]−1 , the positive definiteness of [S] implies the positive definiteness of [C]
(or vice versa) and we can just analyze how this condition applies to [S], which
expression in terms of the engineering constants, Eq. (144), is more simple.
First, consider the special (six) cases in which a single component of the stress
vector {σ} is strictly positive and all the other are zero. Then, Eq. (150) along with
the restriction wc > 0 leads to
1 1
wc = σl Sll σl = Sll σl2 > 0 , l = 1, 2, ..., 6 (152)
2 2
and we obtain the partial conditions

Sll > 0 , l = 1, 2, ..., 6 (153)

which, expressed in terms of the engineering constants yield


Ex > 0 , Ey > 0 , Ez > 0
(154)
Gxy > 0 , Gyz > 0 , Gzx > 0

That way, mechanical work is produced over the orthotropic material (wc = w > 0)
when any normal or shear stress component is acting alone, which is a plausible

- 39-
2 Mechanical Analysis of a Lamina

physical interpretation. We note that the conditions given in Eq. (153) are equiva-
lent to impose that the 1 × 1 principal minors of the matrix [S] are positive-definite,
that is, we have applied the first condition of the well-known Sylvester’s criterion
for symmetric matrices (where the order of the basis elements is permuted six times
for convenience).
Following the Sylvester’s criterion (again, let the basis elements permute), the
determinant of the following 2 × 2 principal minors must be strictly positive
 1 νyx   1 νzy   1 νzx

− − −
 Ex Ey   Ey Ez   Ex Ez 
 νxy 1   νyz 1   νxz 1  (155)
− − −
Ex Ey Ey Ez Ex Ez

which provides the additional relations

1 − νxy νyx > 0 1 − νyz νzy > 0 1 − νxz νzx > 0 (156)

Use the reciprocal relations of Eq. (141) to rewrite the conditions of Eq. (156) as

2 < Ey 2 < Ez 2 < Ex


νyx , νzy , νxz
Ex Ey Ez
Ex Ey Ez (157)
2
νxy < , 2
νyz < , 2
νzx <
Ey Ez Ex

Finally, the determinant (S) of the following 3 × 3 principal minor must also be
greater than zero  1 νyx νzx 
− −
 Ex Ey Ez 
 ν 1 νzy 
 − xy − 
  (158)
 Ex Ey Ez 
 νxz νyz 1 
− −
Ex Ey Ez
whereupon
1 − νxy νyx − νyz νzy − νzx νxz − 2νyx νzy νxz > 0 (159)
The simultaneous fulfillment of Eqs. (154), (156) and (159) ensures the positive
definiteness of the matrix [S]. We say then that the orthotropic material is pointwise
stable and the resulting strain energy w becomes convex.
For isotropic materials, these constraints coalesce to

E>0, G>0 (160)

ν2 < 1 ⇒ − 1 < ν < 1 (161)


and
1 − 3ν 2 − 2ν 3 = (1 + ν)2 (1 − 2ν) > 0 ⇒ ν < 1/2 (162)
which, rewritten in terms of the Young’s modulus E and the Poisson’s ratio ν (taken
as a pair of independent isotropic material parameters), are

E > 0 and − 1 < ν < 1/2 (163)

- 40-
2.4 Apparent material constants. Coupling coefficients.

From physical grounds, the condition E > 0 and the partial limit ν > −1 (equivalent
to G = E/[2(1 + ν)] > 0) imply that any normal or shear stress component acting
alone over the isotropic solid produces positive work, whereas the remaining limit
ν < 1/2 is equivalent to impose that the bulk modulus

trσ/3 E
κ= = (164)
trε 3 (1 − 2ν)

which relates the value of a hydrostatic pressure ∆p = trσ/3 (mean stress) being
applied to the unit volume to the associated variation of volume ∆V = trε (dilata-
tion), is strictly positive. Lastly, we remark that a negative value of the Poisson’s
ratio means that the material expands in transverse direction under uniaxial tension.
Then, for conventional isotropic materials the Poisson’s ratio usually remains within
the reduced range 0 < ν < 1/2.

2.4.5 Reduced stress-strain relations for plane stress

By construction, fiber-reinforced laminae inherently provide high mechanical prop-


erties in the direction of the fibers. We will see in the next chapter that a general
laminate is made bonding together two or more layered composite plies, each one
of them oriented and laid up in a particular direction. Thus, a laminate may be
essentially treated as a two-dimensional structure which bears loads in the plane of
the fibers. Since laminates are built from laminae, we are interested in determining
the strength properties of a single unidirectionally reinforced lamina not only in the
direction of the fibers, but in the plane containing them, that is, the plane Oxy
shown in Figure 16. Finally, the overall in-plane strength of the laminate will be
based on the in-plane strengths of its laminae and their interactions.
The usual approach to treat a lamina as a two-dimensional structure consists in
assuming a plane stress state throughout the lamina. That is, the lamina is only
stressed in the plane containing the fibers. In those cases, if we denote by t the
lamina thickness and consider the plane Oxy to be located at the middle surface of
the lamina, the boundary conditions at both faces of the lamina are

σzx = σzy = σzz = 0 z = ± t/2 (165)

These boundary conditions usually lead to consider the assumption of vanishing


out-of-plane components throughout the thickness of the lamina, namely

σzx = σzy = σzz = 0 , −t/2 ≤ z ≤ t/2 (166)

with the remaining in-plane stress components σxx , σyy and σxy being non-zero in
general. Certainly, Eq. (166) is not strictly exact in a general situation. However, it
is a good approximation within the context of thin plate theory (t = Lz ≪ Lx ∼ Ly ),
as we assume herein (see Section 7.3 of Ref. [3]).
For an orthotropic lamina under plane stress, the strain-stress relation in pre-

- 41-
2 Mechanical Analysis of a Lamina

ferred material directions can be reduced to


    
 ε1  S11 S12 0  σ1 
ε =  S12 S22 0  σ ⇒ (167)
 2   2 
ε4 0 0 S44 σ4
 1 νyx 
  E − 0
 Ey  
 εxx   νx 1   σ xx 
 xy
εyy = − 0  σyy (168)
   Ex Ey  
γxy  1  σxy
0 0
Gxy

where the remaining three equations (σzx = σzy = σzz = 0) yield the out-of-plane
strains
νxz νyz
εzz = − σxx − σyy , εzx = εzy = 0 (169)
Ex Ey
As it is observed in Eq. (168), only four independent constants (Ex , Ey , Gxy and
νyx or νxy ) are needed to be known when the in-plane behavior of the lamina is
investigated (typical values are given in Table 5). Additionally, two more Poisson’s
ratios (νxz and νyz ) are necessary if one wants to calculate the out-of-plane strain
εzz by means of Eq. (169).

Material
Ex Ey νxy Gxy
system
Glass/Epoxy 54 18 0.25 9
Kevlar/Epoxy 76 5.5 0.34 2.1
Graphite/Epoxy 207 5 0.25 2.6
Boron/Epoxy 207 21 0.3 7

Table 5: Typical in-plane elastic constants (E and G in GP a) of unidirectionally


fiber-reinforced laminae. Adapted from Ref. [4].

The stress-strain relation for a plane stress state is not directly obtained removing
the equation numbers 3, 5 and 6 in Eq. (54)1 along with Eq. (145), as we did in
Eq. (54)2 along with Eq. (144) to directly obtain Eq. (168). The reason for that
is that the stress component σzz is zero, but the strain component εzz is not zero.
Then, we simply invert Eq. (168) to obtain
 Ex νyx Ex 
  1 − νxy νyx 0
 1 − νxy νyx  
 σxx     εxx 
σyy

= νxy Ey Ey 
 εyy ⇒ (170)
   0  
σxy  1 − νxy νyx 1 − νxy νyx  γxy
0 0 Gxy
      
 σ1  Q11 Q12 0  ε1   ε1 
σ =  Q12 Q22 0  ε = [Q] ε (171)
 2   2   2 
σ4 0 0 Q44 ε4 ε4

- 42-
2.4 Apparent material constants. Coupling coefficients.

This last stress-strain relation should be complemented with Eqs. (166) and (169)
for a three-dimensional analysis. The 3 × 3 matrix [Q] relating in-plane stresses
to in-plane strains for a plane stress state is known as the reduced stiffness matrix
of the lamina under study. As before, only four material parameters govern the
in-plane stress-strain behavior. Alternatively, the third equation in Eq. (54)1 along
with Eq. (80)
σ3 = 0 = C13 ε1 + C23 ε2 + C33 ε3 (172)
gives
C13 C23
ε3 = − ε1 − ε2 (173)
C33 C33
so the components of the reduced stiffness matrix [Q] are given in terms of the
components of the stiffness matrix [C] through (α, β = {1, 2}, α 6= β; no sum on
repeated indices)

σα = Cαα εα + Cαβ εβ + Cα3 ε3 (174)


 
Cα3 Cβ3
= Cαα εα + Cαβ εβ + Cα3 − εα − εβ (175)
C33 C33
   
C2 Cα3 Cβ3
= Cαα − α3 εα + Cαβ − εβ (176)
C33 C33
= Qαα εα + Qαβ εβ (177)

That is
Cα3 Cβ3
Qαβ = Cαβ − , α, β = {1, 2} (178)
C33
Finally, note that Q44 = C44 = Gxy .

Exercises using MATLAB


Exercise 1 Use MATLAB (or similar) to create a function file (laminaprop.m)
which returns the in-plane elastic constants for a given lamina.
Solution: The code for MATLAB is

function prop = laminaprop(lamina_name)


%
% function prop = laminaprop(lamina_name)
%
% This function returns the material properties of a lamina (SI units)
%
% Input:
% lamina_name = name of lamina (case-sensitive), see below
%
% Output:
% prop = array with material properties
%

- 43-
2 Mechanical Analysis of a Lamina

%
% Names and indices of position (columns) of Material in Mprop
% You can include here more materials
%
Glass_Epoxyn = ’Glass_Epoxy’; Glass_Epoxy = 1;
Kevlar_Epoxyn = ’Kevlar_Epoxy’; Kevlar_Epoxy = 2;
Graphite_Epoxyn = ’Graphite_Epoxy’; Graphite_Epoxy = 3;
Boron_Epoxyn = ’Boron_Epoxy’; Boron_Epoxy = 4;
%
% Material properties, each column is a material
%
% Indices of position (rows) of material properties in Mprop
% are defined in an independent script named matpropindices
% (indirect positioning allows for easy extension of the code)
%
% Gl-E Kv-E Gr-E B-E
Mprop = [ 54e9, 76e9, 207e9, 207e9;
18e9, 5.5e9, 5e9, 21e9;
0.25, 0.34, 0.25, 0.3;
9e9, 2.1e9, 2.6e9, 7e9];
%
% Assignment of material index
%
imat = 0;
if (strfind(lamina_name,Glass_Epoxyn)), imat = Glass_Epoxy; end,
if (strfind(lamina_name,Kevlar_Epoxyn)), imat = Kevlar_Epoxy; end,
if (strfind(lamina_name,Graphite_Epoxyn)), imat = Graphite_Epoxy; end,
if (strfind(lamina_name,Boron_Epoxyn)), imat = Boron_Epoxy; end,
if (imat == 0),
disp([’** ERROR, ’, lamina_name, ’ not in library’]);
return,
end,
%
% Assignment of material data
%
prop = Mprop(:,imat);
%
return

The MATLAB script (matpropindices.m) that defines the indices of position


(rows) of variable Mprop in the preceding function is

%
% ** matpropindices.m script
%

- 44-
2.4 Apparent material constants. Coupling coefficients.

% Indices of position (rows) of material properties in Mprop


% (indirect positioning allows for easy extension of the code)
% This is done in this independent script (instead of a function)
% so new material properties can be easily and readily added
% just adding a row below
%

%
% Propname stores the property name corresponding to each index
%
% Ex = Largest Young modulus (fiber direction = X)
% Ey = Young modulus perpendicular (Y) to fiber direction
% nuxy = Poisson ratio
% Gxy = Shear stiffness in XY direction
%
Exi = 1; Propname(Exi,:) = ’Ex ’;
Eyi = 2; Propname(Eyi,:) = ’Ey ’;
nuxyi = 3; Propname(nuxyi,:) = ’nuxy’;
Gxyi = 4; Propname(Gxyi,:) = ’Gxy ’;

Exercise 2 Write a MATLAB function (laminaSQ.m) to build the reduced (in-


plane) stress-strain relationships of a lamina in principal material directions for
given preferred properties of a lamina.
Solution: The code for MATLAB is

function [Smat,Qmat] = laminaSQ(prop)


%
% function [Smat,Qmat] = laminaSQ(prop)
%
% function to build the stress-strain relationships of a lamina
% in principal in-plane material directions
% standard Voigt notation is used (engineering shear strains)
%
% Input:
% prop = vector of material properties, index contained in
% matpropindices.m
%
% Output:
% Smat, Qmat = reduced (in-plane) compliance and stiffness matrices
% in principal material directions
%

%
% retrieves material properties indices
%

- 45-
2 Mechanical Analysis of a Lamina

matpropindices;
%
% retrieves relevant material data from array of material properties
%
Ex = prop(Exi);
Ey = prop(Eyi);
nuxy = prop(nuxyi);
Gxy = prop(Gxyi);
nuyx = nuxy/Ex*Ey; % nuyx from symmetry considerations
%
% Builds matrices
%
Smat = zeros(3); % initializes matrix to zero
Smat(1,1) = 1./Ex;
Smat(2,1) = -nuxy/Ex; % warning: note indices convention for nu
Smat(1,2) = Smat(2,1); % S is symmetric
Smat(2,2) = 1./Ey;
Smat(3,3) = 1./Gxy; % note that this is S44
%
Qmat = zeros(3); % initializes matrix to zero
aux = 1. - nuxy*nuyx;
Qmat(1,1) = Ex/aux;
Qmat(2,1) = nuxy*Ey/aux; % warning: note indices convention for nu
Qmat(1,2) = Qmat(2,1); % Q is symmetric
Qmat(2,2) = Ey/aux;
Qmat(3,3) = Gxy; % note that this is C44
%
return

Exercise 3 Create a MATLAB function to obtain the reduced (in-plane) stress-


strain relationships of a lamina expressed in a system of representation rotated a
given angle with respect to the principal lamina axes.
Solution: First, the following function file (planerotationmtx.m) computes the trans-
formation matrices associated to stresses and strains in Voigt notation

function [Rs,Re] = planerotationmtx(angle)


%
% function [Rs,Re] = planerotationmtx(angle)
%
% This funtion creates the plane rotation matrices for stress
% and strain tensors in plane Voigt notation (11,22,12 components)
%
% Input:
% angle = clockwise rotation angle in degrees:
%

- 46-
2.4 Apparent material constants. Coupling coefficients.

% / (old x-axis)
% /
% / ) angle
% ------- (new x-axis)
%
% Output:
% Rs = plane rotation matrix for stresses
% Re = idem strains
%

c = cosd(angle);
s = sind(angle);
%
% Rotation mtx for stresses
%
Rs = [ c^2, s^2, -2*c*s; ...
s^2, c^2, 2*c*s; ...
s*c, -s*c, c^2-s^2];
%
% Rotation mtx for strains in Voigt notation (engrg shear strain)
%
Re = [ c^2, s^2, -c*s; ...
s^2, c^2, c*s; ...
2*s*c, -2*s*c, c^2-s^2];
%
return

Second, the MATLAB function (laminarotate.m) for the stiffness and compliance
matrices of a lamina expressed in a given system of representation is

function [Smat,Qmat] = laminarotate(Sold,Qold,angle)


%
% function [Smat,Qmat] = laminarotate(Sold,Qold,angle)
%
% Function to obtain rotated constitutive matrices for a given angle
%
% Input:
% Sold = old compliance (flexibility) matrix
% Qold = old stiffness matrix
% angle = clockwise rotation angle (deg):
%
% / (old x-axis, for example principal material one)
% /
% / ) angle
% ------- (new x-axis, for example laminate one)

- 47-
2 Mechanical Analysis of a Lamina

%
% Output:
% Smat = new compliance matrix
% Qmat = new stiffness matrix
%

%
% rotation matrices Rs for stresses and Re for strains in Voigt notation
%
[Rs,Re] = planerotationmtx(angle);
%
% represent matrices and vectors in laminate axes
%
Qmat = Rs*Qold/Re; % Rs*Q*inv(Re)
Smat = Re*Sold/Rs; % Re*S*inv(Rs)
%
return

Exercise 4 Create a MATLAB function (plotapparent.m) to plot the distribution


of apparent elastic constants Ex′ , Gxy′ , νxy′ and ηx,xy′ in the interval α ∈ [0, 90] (see
Figure 17) for given preferred properties of a lamina.
Solution: The code for MATLAB is

function plotapparent(prop)
%
% function plotapparent(prop)
%
% Function to plot apparent lamina material property values
%
% Input:
% prop = vector with material properties
% arranged as given by script matpropindices
%

%
% Build matrices in principal material directions
%
[S,Q] = laminaSQ(prop);
%
% loop on different angles
%
dang = 1.; % increment of angle in degrees for the plot
n = 90. / dang; % number of steps to go from 0 to 90
%
for i=1:n+1,

- 48-
2.4 Apparent material constants. Coupling coefficients.

angle = (i-1) * 90 /n;


[Smat,Qmat] = laminarotate(S,Q,angle); % rotates
ang(i) = angle; % saves angle
%
% Apparent stiffnesses for uniaxial tests
%
Ex(i) = 1/Smat(1,1); % apparent Young’s modulus
Gxy(i) = 1/Smat(3,3); % apparent shear modulus
nuxy(i) = -Smat(2,1)/Smat(1,1); % apparent Poisson’s ratio
lxxy(i) = Smat(1,3)/Smat(3,3); % apparent Lekhnitskii’s coef.
%
end,
%
% Perform plots of apparent elastic constants
%
figure
subplot(1,2,1); hold on;
hEx = plot(ang,Ex, ’-r’, ’linewidth’,2);
hGxy = plot(ang,Gxy,’--k’, ’linewidth’,2);
xlabel(’Angle [deg]’); ylabel(’E_x and G_{xy}’);
legend([hEx,hGxy],’E_x’,’G_{xy}’);
set(gca,’xlim’,[0,90]); grid on;
subplot(1,2,2); hold on;
hnuxy = plot(ang,nuxy, ’-r’, ’linewidth’,2);
hlxxy = plot(ang,lxxy, ’--b’, ’linewidth’,2);
xlabel(’Angle [deg]’); ylabel(’\nu_{xy} and \eta_{x,xy}’);
legend([hnuxy,hlxxy],’\nu_{xy}’,’\eta_{x,xy}’);
set(gca,’xlim’,[0,90]); grid on;
suptitle(’Apparent uniaxial material constants’);
%
return

Auxiliary function file suptitle.m is

function ha = suptitle(ttext)
%
% function ha = suptitle(ttext)
%
% plots a general title over all subplots
%
% Input:
% ttext = title to be plotted
%
% Output:
% ha = handle to the axes of the title

- 49-
2 Mechanical Analysis of a Lamina

ha = axes(’Position’,[0 0 1 1],’Xlim’,[0 1],’Ylim’,[0 1],...


’Box’,’off’,’Visible’,’off’,’Units’,’normalized’,’clipping’,’off’);
%
text(0.5,1,ttext,’HorizontalAlignment’,’center’,...
’VerticalAlignment’,’top’,’interpreter’,’none’);
%
return

Exercise 5 Plot the distribution of apparent elastic constants Ex′ , Gxy′ , νxy′ and
ηx,xy′ in the interval α ∈ [0, 90] (see Figure 17) for a Glass-Epoxy lamina.
Solution: Just type in the MATLAB Command Window

lamina_name = ’Glass_Epoxy’;
prop = laminaprop(lamina_name);
plotapparent(prop);

The MATLAB output is shown in the following Figure

10
Apparent uniaxial material constants
x 10
5.5 0.3

5
0.2
4.5 Ex νxy

4 Gxy 0.1 ηx,xy

3.5
νxy and ηx,xy
Ex and Gxy

0
3
−0.1
2.5

2 −0.2

1.5
−0.3
1

0.5 −0.4
0 20 40 60 80 0 20 40 60 80
Angle [deg] Angle [deg]

Figure of Exercise 5

- 50-
2.5 Temperature, moisture and piezoelectric effects

Exercise 6 Plot the distribution of apparent elastic constants Ey′ and ηy,xy′ in the
interval α ∈ [0, 90] (see Figure 17) for a Glass-Epoxy lamina.
Solution: Modify the file plotapparent.m so that it performs the requested plots. We
leave this task to the reader.

2.5 Temperature, moisture and piezoelectric effects


In the previous sections we have dealt with the mechanical response of a lamina
taking into account only purely “mechanical” actions. We have not considered the
temperature effects and the change in humidity.
In composite materials the influence of both temperature and humidity are of ma-
jor importance. Composites are cured at relatively high temperature and lowering
humidity conditions. Then, composite structures work frequently at low temper-
atures. For example the outside air temperature in a plane at 35000 feet (about
10.000m) is about -55o C (although friction may rise the temperature to -25o C). In
space, temperatures vary a lot depending on the distance to the sun and if it is
under a shade; they may be as low as -270o C. In nonsymmetric laminates, warping
may appear as a consequence of temperature changes. Humidity is also important
because composite materials are strongly affected by their moisture content.
A further problem regarding humidity effects is that moisture is slowly released
and, hence, an important gradient may be present in a laminate. Piezoelectric
extensometers and actuators are common composite material devices which take
advantage of the piezoelectric effect to create “smart” actuators and “smart” struc-
tures (morphing) which slightly change their shape in order to obtain an enhanced
behavior under varying conditions. This is precisely one of the most promising fu-
ture applications of composites. The mathematical treatment of all these effects is
similar. Hence we will study them altogether in this section.

2.5.1 Temperature effects


In contrast to isotropic or homogeneous anisotropic materials and structures, tem-
perature changes affect significantly the strength of composite laminates. This is
due to the different dilatation properties that fibers and matrix have. Different ply
orientations produce internal stresses when a change of temperature occurs respect
to the curing (bonding) temperature. These temperature effects add to the usual
(also present in isotropic materials) stress and strain changes when temperature
gradients are present. Thus, in composite laminates, there are internal stresses and
strains even in the case of a homogeneous temperature in the specimen.
The basic equations of temperature effects result from thermodynamical princi-
ples (see Appendix C for a formal treatment) that leads to Eq. (1058), which we
reproduce here for the reader comfort in incremental form

ε = S : σ + α (T − T0 ) (179)

where S is the tensor of elastic compliances and α is the tensor of dilatation con-
stants. Basically this equation means that thermal strains add to “mechanical

- 51-
2 Mechanical Analysis of a Lamina

strains” produced by loads, i.e. derived from stresses2 . For example, for an isotropic
bar loaded in x−direction so σ = {σxx , 0, 0, 0, 0, 0} T

Total = “mechanical” + “thermal” 


εxx = (1/E) σxx + α (T − T0 )
(180)
εyy = −ν (1/E) σxx + α (T − T0 ) 


εzz = −ν (1/E) σxx + α (T − T0 )

i.e. each dimension is enlarged a quantity ∆l = α l (T − T0 ), where l is the original


length and T0 the original temperature. Anisotropy simply implies that we have
different material properties and hence different dilatation properties, in the different
directions. For an orthotropic lamina, Eq. (179) may be written in principal material
directions in Voigt matrix notation as
 1 νyx νzx 
− − 0 0 0
 Ex Ey Ez 
 νxy 1 νzy 

εxx
 
 − − 0 0 0 
 σxx






  Ex Ey Ez  


 εyy 
  νxz νyz 1  σyy 


 
  − − 0 0 0  

εzz  Ex Ey Ez  σzz

= 1
 + ...
 γxy
 


 0 0 0 0 0

 σxy 

   
 γ   Gxy  σyz 
 yz

 
  
  1 
 

γzx 
 0 0 0 0 0  σzx

 Gyz 

1
0 0 0 0 0
Gzx
 

 αxx 


 αyy 


 

 
αzz
... + (T − T0 ) (181)

 0 

 
 0





 
0
Of course, this equation may be expressed in any arbitrary system of representation
X′

{ε}X = [S]X {σ}X + {α}X (T − T0 ) (182)


[Rε ]−1 −1
X→X ′ {ε}X ′ = [S]X [Rσ ]X→X ′ {σ}X ′ + {α}X (T − T0 )
{ε}X ′ = [Rε ]X→X ′ [S]X [Rσ ]−1
X→X ′ {σ}X ′ + [Rε ]X→X ′ {α}X (T − T0 )
{ε}X ′ = [S]X ′ {σ}X ′ + {α}X ′ (T − T0 ) (183)

so dilatation coefficients change as strains, which should be apparent from Eq. (179),
i.e.
{α}X ′ = [Rε ]X→X ′ {α}X (184)
2
The terms “mechanical” is somehow misleading but we keep the usual terminology.

- 52-
2.5 Temperature, moisture and piezoelectric effects

As a consequence, the Voigt representation of the second-order tensor α in the


general system of representation X ′ includes the factor of 2 in the shear terms
   

 α1′ 
 
 αxx′ 

  
 α2′ 
 
 αyy′ 

   
αxx′ αxy′ αyz ′ 
 ′ 
 
 

V oigt α3 αzz ′
[α]X ′ = αxy′ αyy′
 αzx′  ; {α}X ′ = = (185)
 α4′   2αxy′ 
αyz ′ αzx′ αzz ′ 
 
 
 

 α′   2αyz ′
 
 5

 
 
   

α6′ 2αzx′

where αij ′ stand for the apparent dilatation constants associated to the system of
representation X ′ , which are uniquely determined from the three (independent)
preferred dilatation constants αxx , αyy and αzz and the given orientation of X ′ with
respect to X.
Equation (179) may be inverted factoring-out the stresses, or alternatively ob-
tained from a different thermodynamic potential to yield Equation (1056) of Ap-
pendix C, which we reproduce here in incremental form for the reader comfort

σ = C : ε − β (T − T0 ) (186)

where
β=C:α (187)

Alternatively
σ = C : [ε − α (T − T0 )] (188)

where εM = ε − α (T − T0 ) is usually named mechanical strains3 and are the strains


that produce stresses. Eq.(186) may be written in principal material directions using
Voigt notation as
    

 σxx 
 C11 C12 C13 0 0 0 
 εxx 


 σyy 
  C12 C22 C23 0 0 0  εyy 


 
   

   
σzz  C13 C23 C33 0 0 0 εzz
=  − ...
 σxy
 


 0 0 0 C44 0 0 
 γxy 

   
 σ   0 0 0 0 C55 0  γyz 
 yz

 
 
 

  
σzx 0 0 0 0 0 C66 γzx

 

 βxx 


 βyy 


 

 
βzz
... − (T − T0 ) (189)
 0
 

 


 0 


 
0

3
Perhaps a more accurate name would have been “stressing strains”, but we keep the usual
terminology.

- 53-
2 Mechanical Analysis of a Lamina

As for the previous expressions, this equation may be written in any arbitrary system
of representation (for example that of a laminate)

{σ}X = [C]X {ε}X − {β}X (T − T0 ) (190)


[Rσ ]−1
X→X ′ {σ}X ′ = [C]X [Rε ]−1
X→X ′ {ε}X ′ − {β}X (T − T0 )
{σ}X ′ = [Rσ ]X→X ′ [C]X [Rε ]−1
X→X ′ {ε}X ′ − [Rσ ]X→X ′ {β}X (T − T0 )
{σ}X ′ = [C]X ′ {ε}X ′ − {β}X ′ (T − T0 ) (191)

We note that in this case the rotation matrix to be applied to the β coefficients are
those of the stresses, i.e.
{β}X ′ = [Rσ ]X→X ′ {β}X (192)

Also note that in contrast with the Voigt representation of the dilatation coefficients
Eq. (185), the Voigt representation of β does not have the factor of 2 in the shear
terms when changed to an arbitrary system of representation, i.e.
   

 β1′ 
 
 βxx′ 

  
 β2′ 
 
 βyy′ 

   
βxx′ βxy′ βyz ′ 
 ′ 
 
 

V oigt β3 βzz ′
[β]X ′ = βxy′ βyy′
 βzx′  ; {β}X ′ = = (193)
 β4′   βxy ′ 
βyz ′ βzx′ βzz ′ 
 
 
 

 β5′
   β ′
 
  yz
  
 

 
β6′ βzx′

It is customary to separate the thermal contribution from the mechanical con-


tribution writing
σ = C : ε − σT (194)

where σ T = β∆T = C : α (T − T0 ) are the “thermal” stresses.


In the case of an orthotropic lamina under plane stress in Voigt notation, we can
use Eq. (188) to write in the principal lamina directions
    
 σ1  Q11 Q12 0  ε1 − α1 (T − T0 ) 
σ =  Q12 Q22 0  ε − α2 (T − T0 ) (195)
 2   2 
σ4 0 0 Q44 ε4

and in an arbitrary system of representation


    
 σ1′  Q11′ Q12′ Q14′  ε1′ − α1′ (T − T0 ) 
σ2′ =  Q12′ Q22′ Q24′  ε ′ − α2′ (T − T0 ) (196)
   2 
σ4′ Q14′ Q24′ Q44′ ε4′ − α4′ (T − T0 )

i.e.
{σ} = [Q] {ε − α∆T } = [Q] {ε} − [Q] {α} ∆T (197)

with ∆T = T − T0 .

- 54-
2.5 Temperature, moisture and piezoelectric effects

2.5.2 Moisture effects


As mentioned, the humidity content of a lamina also affects its weight and its di-
mensions. Humidity also transfers heat. The importance of humidity lies in the fact
that composites are cured in an autoclave at high temperatures. Then, the working
condition of composites may differ substantially in terms of ambient humidity.
Here we will simply consider the dilatation contribution of humidity. The equa-
tions are similar in format to those of thermomechanical effects, as shown in Ap-
pendix C. We recall here for the reader comfort Eq.(1075) in incremental form,
including the thermal effects

ε = S : σ + α (T − T0 ) + αH (H − H0 ) (198)
where αH are the coefficients of hygroscopic expansion, (H − H0 ) is the change in
humidity and ε are the hygrothermomechanical strains. This expression can be
inverted to obtain a similar expression to that of Eq. (186)
σ = C : ε − β (T − T0 ) − β H (H − H0 ) (199)
where as for the thermal case
β H = C : αH (200)
and
σ = C : ε − σ T − σ H = C : ε − σ HT (201)
where
σ T = C : α (T − T0 ) , σ H = C : αH (H − H0 ) and σ HT = σ H + σ T (202)
are respectively the thermal, hygroscopic and hygrothermal stresses. We note that
moisture dilatation effects may be included into thermal effects using a fictitious
temperature and fictitious dilatation coefficients. For T > T0 and H ≥ H0 , Eq.
(198) may be written as
(H − H0 )
ε = S : σ + α (T − T0 ) + αH (T − T0 )
(T − T0 )
 
(H − H0 )
= S : σ + α + αH (T − T0 )
(T − T0 )
= S : σ + αHT (T − T0 ) (203)
where
(H − H0 )
αHT = α + αH (204)
(T − T0 )
are the fictitious hygrothermal dilatation coefficients. This is a frequent “trick” to
use standard thermal software to include also hygroscopic effects. Of course, if there
is no change in temperature (T = T0 ) simply take
(H − H0 ) → (T − T0 ) and αH → α (205)
and perform a fictitious “thermal” analysis. Therefore, from now on we will consider
moisture effects embed into thermal effects and no distinction will be made unless
necessary.

- 55-
2 Mechanical Analysis of a Lamina

2.5.3 Piezoelectric effects


As mentioned, piezoelectric sensors and actuators are very common in engineering.
Fiber composites, as man-made fibrous materials present a unique opportunity to
provoke user controlled deformations or to measure them in an integrated manner.
Piezoelectric multilayered composites are used for example for barometers, sonars,
flat loudspeakers, etc.
Piezoelectric composites may be employed both as sensors and as actuators.
For both cases the most comfortable equation set to work with is different. For
a full understanding of the equations, refer to Appendix C, Section C.3. We just
recall from the appendix the three usual sets of equations (for simplicity of this
presentation we omit in this section thermal effects)

• Constitutive equations in natural format:



σ =C:ε−e·E
(206)
D = eT : ε + ǭ E

where E = − ∇V is the electric field strength vector (units of V / m or N / C),


V is the electric potential (units of V), D is the
 dielectric charge displacement
vector (units of C / m2 or N / (V m) = N / m2 / (V / m)), e(ij)k is a third order
tensor containing the piezoelectric moduli or piezoelectric stress constants –
units of N / (V m)–, and ǭ is the second order tensor of permittivities measured
at constant strains, units of F / m or equivalently C2 /(m2 N), see Appendix
C.3.

• Constitutive equations in usual, actuator, form, see Eq. (1111) and (1112):

ε = S : σ + dT · E
(207)
D = d : σ + ǫ̃ · E

where di(jk) is the third-order tensor of piezoelectric (strain) constants or piezo-


electric charge coefficients4 (units of m / V or C / N) and ǫ̃ is the second order
tensor of permittivities measured at constant stresses. The following identities
hold as it is straightforward to show from the previous two sets of equations

ǫ̃ = eT : dT + ǭ and e = C : dT (208)

• Constitutive equations in sensor form, see Eqs.(1115) and (1119) from Ap-
pendix C: 
ε = SD : σ + g T · D
(209)
E = −g : σ + ǫ̃−1 · D
where gi(jk) is the third-order tensor of piezoelectric (stress) constants (units
of m2 / C) or piezoelectric voltage coefficients (because it also has units of
V m / N) and SD is the tensor of elastic compliances measured at constant
4
Typical values for PZT polarized in z−axis are d311 = d322 = −171 × 10−12 C / N, d333 =
374 × 10−12 C / N, d123 = d213 = 584 × 10−12 C / N.

- 56-
2.5 Temperature, moisture and piezoelectric effects

electric displacements (S is measured at constant electric field). The following


equations hold as the reader may check

g = ǫ̃−1 · d and SD = S − dT · ǫ̃−1 · d (210)

The third order tensors d, e, g relate mechanical tensors (stresses and strains)
with electric vectors (electric field and electric displacement). Usually Voigt notation
is used for these tensors. To make easier the tensor-to-Voigt notation conversion we
have enclosed in parenthesis indices that relate with mechanical tensors and, hence,
are collapsible pairs. The remaining index is not collapsible. For example
 
d11 d12 d13 2d14 2d15 2d16
di(kl) → [dij ] =  d21 d22 d23 2d24 2d25 2d26  (211)
d31 d32 d33 2d34 2d35 2d36
where the first index takes the vector values of i = 1, 2, 3 and whereas the second
index takes Voigt notation values j = 1, ..., 6 according to pair (kl).The first index
(rows) is also known as the vector side, whereas the second pair of indices (columns)
is known as the tensor side. The factors of 2 arises from the consideration that the
tensorial operation d : σ in Eq. (207)2 and the matrix operation [d] {σ} must be
in agreement. The transpose is defined as (note that the factors of 2 are consistent
with the shear strain components of {ε} in Eq. (207)1 ) i.e.
 
d11 d21 d31
 d12 d22 d32 
 
 T    d13 d23 d33 
di(kl) = d(kl)i =    (212)
 2d14 2d24 2d34 

 2d15 2d25 2d35 
2d16 2d26 2d36
For an orthotropic lamina, assuming that the material is poled in the (−z) direction
(dipoles to −z), usual representations in principal directions are (we leave to the
reader the task of proving this expression using symmetry transformations)
   
0 0 0 0 0 2d16 ǫ̃11 0 0
di(kl) → [dij ] =  0 0 0 0 2d25 0  ; ǫ̃ =  0 ǫ̃22 0  (213)
d31 d32 d33 0 0 0 0 0 ǫ̃33
Frequently piezoelectric materials (e.g. PZT devices) are considered transversely
isotropic with the z-axis being the axis of symmetry, in such case
 
0 0 d31
 0 0 d31   
  ǫ̃11 0 0
 0 0 d33 
[dij ]T = 
 0 ; ǫ̃ =  0 ǫ̃11 0  (214)
0 0 

 0
 0 0 ǫ̃33
2d16 0 
2d16 0 0
We leave to the reader the task of writing Eqs. (206), (207) and (209) in Voigt
notation for a lamina.

- 57-
2 Mechanical Analysis of a Lamina

Evidently, if we need to work in an arbitrary system of representation X ′ , the


usual rotation rules apply to Eqs. (206)–(210). For example

{E}X ′ = [r]X→X ′ {E}X (215)

[ǫ̃]X ′ = [r]X→X ′ [ǫ̃]X [r]X ′ →X (216)

[d]X ′ = [r]X→X ′ [d]X [Rσ ]X ′ →X (217)

but (note that a different rotation matrix has to be used for the tensor sides of d
and dT )
 T  
d X ′ = [Rε ]X→X ′ dT X [r]X ′ →X (218)

Exercise 7 Write Equations (206), (207) and (209) in Voigt notation for a lamina.
Solution: We write the first of the equations in principal material directions and
leave the rest of the equations to the reader.
Equation (2061 ) in Voigt notation is
    

 σ 1 
 C11 C12 C13 0 0 0 
 ε1 


 σ 
  C12 C22 C23 0 0 0  ε2 


 2 
  
   
 

σ3  C13 C23 C33 0 0 0 ε3
=  ...
 σ4 
   0 0 0 C44 0 0 
 ε4 

   


 σ 5



 0 0 0 0 C55 0 

 ε5 


   
σ6 0 0 0 0 0 C66 ε6
 
0 0 e31

 0 0 e32  
  E1


 0 0 e33 
−  E2 (219)

 0 0 0    E3 
 0 e25 0 
e16 0 0

Exercise 8 Obtain the change of system of representation for tensors e, eT , g


and g T in Voigt notation and verify that ǫ̃ = eT : dT + ǭ, e = C : dT and
SD = S − dT · ǫ̃−1 · d are correctly obtained.
Solution: We leave this exercise to the reader.

2.6 Failure criteria for composite laminae


In the previous sections we have described the linear elastic response of both isotropic
and anisotropic materials. This study has lead to the introduction of the concept of
stiffness (and compliance), which is the determinant design factor for applications
in which small, elastic (or recoverable), deformations are considered. However, the
mechanical behavior beyond this limit has also to be taken into account for the
design of composite structures for general applications.

- 58-
2.6 Failure criteria for composite laminae

σxx

Xt
y
Xt Xt
x

εxx
Xε t
Figure 20: Uniaxial loading in the fiber direction of a unidirectionally reinforced
lamina. Definition of the axial tensile allowable stress Xt and strain Xtε .

2.6.1 Strength of an orthotropic lamina in principal material directions


In this section we focus on the two-dimensional (in-plane) strength analysis of a
unidirectionally reinforced lamina. In this context, besides the in-plane (reduced)
stiffness and compliance constants introduced in Section 2.4.5, five more strength-
like mechanical properties are to be defined.
In Figure 20, the stress-strain curve obtained from a uniaxial tensile test per-
formed on a lamina along the direction of the fibers is shown. The test is carried out
until the fracture of the lamina is obtained. The Young’s modulus Ex characterizes
the initial elastic response and, after a (rather unusual) inelastic response, the failure
of the lamina occurs at σxx = Xt . This last value, which we will show below that
corresponds to a fiber-dominated property, is called the axial strength in tension of
the lamina. Sometimes it is also referred as the axial tensile allowable stress. Aside,
the axial compressive strength Xc results from the application of compressive loads
to the lamina until its failure is reached. Most composite materials present different
strength properties in tension and compression, so both Xt and Xc must be known
to provide a complete description of the axial strength of a lamina. In contrast, note
that only one material constant is assumed to govern the tensile-compressive elastic
response, i.e. Ex , which is a direct consequence of the constitutive linearization.
Similarly, transverse tensile and compressive strengths Yt and Yc , respectively,
can be defined for an orthotropic lamina from the corresponding uniaxial tests per-
formed along the principal material axis Oy (Figure 21.a). As it can be easily
inferred from direct observation (and we will show below), the allowable stresses Yt
and Yc are matrix-dominated properties. Hence, for example, it has not to be a
surprising result that for composite materials Xt usually takes a value that is one
or two orders greater than Yt (see Table 6). This fact reveals a high anisotropic
behavior in terms of strength.
The last fundamental allowable stress, defined in principal material coordinates,
is the shear strength S, as it is shown in Figure 21.b. Due to symmetry issues, we

- 59-
2 Mechanical Analysis of a Lamina

a) b)
Yt
S
y y

x x

S
Yt
Figure 21: a) Uniaxial loading in the transverse direction of a unidirectionally rein-
forced lamina. Definition of the transverse tensile allowable stress Yt . b) Pure shear
loading in principal coordinates of a unidirectionally reinforced lamina. Definition
of the shear strength S.

note that no distinction is needed to be considered between “positive” and “negative”


shear strengths when they are measured in the preferred material axes. Several tests
may be used to measure the strength S of an orthotropic lamina. Some examples
are the so-called torsion-tube test, sandwich cross-beam test and rail shear test.

Material
Xt Xc Yt Yc S
system
Glass/Epoxy 1.04 −1.04 0.028 −0.14 0.041
Kevlar/Epoxy 1.38 −0.28 0.028 −0.14 0.044
Graphite/Epoxy 1.04 −0.69 0.041 −0.12 0.069
Boron/Epoxy 1.38 −2.76 0.083 −0.28 0.124

Table 6: Typical in-plane strength characteristics (in GP a) of unidirectionally fiber-


reinforced laminae. Adapted from Ref. [4]

Exercise 9 Modify the MATLAB function file and the script of the Exercise 1
(laminaprop.m and matpropindices.m, respectively) to include the principal allow-
able stresses of Table 6.
Solution: We leave this exercise to the reader.

As a summary, we conclude that, besides the four elastic material constants Ex ,


Ey , Gxy and νyx of a unidirectionally reinforced lamina, we need to know the five
strengths
Xt , Xc , Yt , Yc and S (220)
which are defined (and measured) in principal material coordinates, to be able to
completely characterize the in-plane behavior of the lamina. In this respect, it should

- 60-
2.6 Failure criteria for composite laminae

be taken into account that the experimental determination of laminae allowable


stresses is complicated by the varied and sometimes complex failure modes inherent
to this type of composite materials. As mentioned above, both stiffness and strength
characteristics of laminae (and of course of the resulting laminate built from laminae)
will be of central importance for the design of composite structures. The five values
Xt , Xc , Yt , Yc and S characterizing the fracture of a fiber-reinforced lamina compare
to the one mechanical property needed to characterize the yielding of conventional
isotropic metallic materials, for which the only axial yield strength is X = Xt =
−Xc = Yt = −Yc and the shear yield strength S directly results from the value of
X and the specific constitutive model being assumed, as we will see in the following
section. However, other nonideal isotropic materials could need two yield properties
to characterize its failure behavior, see for example Ref. [5].

2.6.2 Usual strength criteria for biaxial in-plane stresses


In the previous section we have defined what are the fundamental allowable stresses
of an orthotropic lamina and how they can be measured from some specific exper-
imental tests. However, a lamina can be subjected to forces in its own plane in
other multiple different ways. A generalization of the previous results for arbitrary
in-plane biaxial states results necessary.

Isotropic metallic laminae


First, we provide an overview of the well-known strength criteria for isotropic ma-
terials, widely used for the design of metallic structures. In this simple study, we
will consider that the material fails when the yield strength is reached. That is,
we restrict the present study to the context of metal plasticity. From physical
grounds (crystal plasticity), it is accepted in engineering practice that the yielding
phenomenon in most metals depends only on the deviatoric stresses, so we work
herein with the deviatoric stress tensor
tr (σ)
σD = σ − I =P:σ (221)
3
where P is the fourth-order projector (or deviatoric) tensor, expressed in matrix
notation as  
2/3 −1/3 −1/3 0 0 0
 −1/3 2/3 −1/3 0 0 0 
 
 −1/3 −1/3 2/3 0 0 0 
[P] = 
 0
 (222)
 0 0 1 0 0  
 0 0 0 0 1 0 
0 0 0 0 0 1
Then, in the principal stress basis, the tensors σ and σD are expressed in terms of
the principal stresses σ1 , σ2 and σ3 as
 
σ1 0 0
[σ] =  0 σ2 0  (223)
0 0 σ3

- 61-
2 Mechanical Analysis of a Lamina

and  
2σ1 − σ2 − σ3 0 0
 D
 1
σ = 0 2σ2 − σ3 − σ1 0  (224)
3
0 0 2σ3 − σ1 − σ2
If we consider a plane stress (biaxial) state throughout a metallic lamina, then σ3 = 0
and Eq. (224) reduces to
 
 D  1 2σ1 − σ2 0 0
σ =  0 2σ2 − σ1 0  (225)
3
0 0 −σ1 − σ2
The first yield condition we introduce is the maximum shear stress failure criterion
or the Tresca criterion, which states that yielding occurs at a point of the lamina if
D
σt = S = τy (226)
max

where σtD stands for any shear component (in absolute value) of σ D and S = τy
represents the yield shear strength of the material. Since the only difference Dbetween

D
σ and σ is the hydrostatic contribution, i.e. (tr (σ) /3)I, note that σt max =
|σt |max . The shear strength-axial strength relation S = τy = σy /2 = X /2, associated
to the Tresca failure criterion, follows when the yield point is reached in a uniaxial
tensile test, see Figure 22, and indicates that only one of the strengths X = σy and
S = τy can be independently defined. The elastic limit σy is easily measured and
will definitely be the value used to define this failure criterion. For specific values
σ1 and σ2 , the maximum shear stress σtD max obtained from Eq. (225) is
( )
D σ D − σ D σ D − σ D σ D − σ D
σt 1 2 2 3 3 1
max
= max , , (227)
2 2 2
 
|σ1 − σ2 | |σ2 | |σ1 |
= max , , (228)
2 2 2
= |σt |max (229)

which combined with Eq. (226), along with the relation τy = σy /2, gives the expres-
sion of the Tresca criterion applied to biaxial states of stress

max {|σ1 | , |σ2 | , |σ1 − σ2 |} = σy (230)

When |σ1 |, |σ2 | or |σ1 − σ2 | reaches the value σy at some point of the solid, then
failure of the material occurs at that point. Accordingly, yielding is prevented if
max {|σ1 | , |σ2 | , |σ1 − σ2 |} < σy . The maximum shear stress envelope is represented
in Figure 23 in the space of principal stresses. That is, that figure is valid whatever
the orientation of the principal stresses is; only the principal stress values σ1 and σ2
are relevant for the failure study. This feature is a direct consequence of the isotropic
behavior and facilitates the yield strength characterization to a large extent.
Regarding the Tresca criterion, it has been observed that the failure of some
brittle metals is completely independent of the value |σ1 − σ2 |, so we rewrite Eq.
(230) as
max {|σ1 | , |σ2 |} = σy (231)

- 62-
2.6 Failure criteria for composite laminae

σt
τY σ1=σY=2τY

σ1 σ1
σn
σ2=σ3=0

Figure 22: Uniaxial loading of an isotropic lamina until the Tresca criterion
(|σt |max = τy ) is satisfied. As a result σy = 2τy .

which is known as the maximum normal stress failure criterion or the Rankine
criterion. It is remarkable that for the normal stress envelope (Figure 23), the
failure along a principal stress direction is the same regardless the value of the other
principal stress. The reader may prove that for this model the shear strength-axial
strength relation is S = τy = σy = X .
For a large variety of metals, the Tresca criterion has shown to be too restrictive,
that is, the actual yielding of the material occurs beyond the maximum shear stress
envelope. The maximum distortional energy failure criterion, also known as the von
Mises yield criterion, circumvents this issue5 . This failure criterion states that the
yielding begins at a certain point of the solid when the distortion energy per unit
volume reaches a critical value at that point. The distortion energy represents the
energy employed to deform/distort the unit cube without an associated change of
volume, that is, it is a genuine deviatoric measure. For linear isotropic elasticity,
the total stored energy is decomposed into its deviatoric (D) and volumetric (V )
contributions through
1 1 1
w = σ : ε = σ D : εD + σ V : εV = wD + wV (232)
2 2 2
where the products σD : εV = σV : εD = 0, not indicated, vanish as does the
double-dot product of any deviatoric and volumetric tensors6 . For isotropic materials
distortion and dilatation are said to be fully uncoupled because deviatoric stresses
are directly related to deviatoric strains and volumetric stresses are directly related
to volumetric strains through σ D = 2µεD (µ = G is the second Lame’s constant)
and σ V = 3κεV , respectively.
Therefore, the maximum distortional energy failure criterion predicts that yield-
ing occurs if
1 1 D
w D = σ D : εD = σ : σ D = K∗ (233)
2 4µ
where K∗ is the critical distortion energy. The von Mises criterion is also called the
J2 −criterion due to the fact that the second invariant J2 of the deviatoric stress
5
This criterion was originally introduced by Maxwell as the distortion energy criterion and later,
under different statements, by von Mises, Huber, Hencky and Nadai.
6
For example, σ V : εD = pI : εD = p(trεD ) = 0, since trεD = 0.

- 63-
2 Mechanical Analysis of a Lamina

τY

σ1 = σY σ1 = τY
σY σY
σ2 =0 σ2 = ̶ τY

τY

σ2
R
VM
σY
T

σ1
(τY )T
(τY )VM
(τY )R
Figure 23: Failure criteria for metals. T ≡ Tresca criterion. R ≡ Rankine criterion.
V M ≡ Von Mises criterion.

- 64-
2.6 Failure criteria for composite laminae

tensor σ D is
1 D 2  1 D
J2 = σ : σD − trσ D = σ : σD (234)
2 2
so, analogously, we can say that yielding begins if the invariant J2 reaches the critical
value J2∗ = 2µK∗
1
J2 = σD : σ D = J2∗ (235)
2
Another useful interpretation emerges directly if we take into account that the prod-
2
D D
uct σ : σ represents the squared norm of the deviatoric stress tensor, i.e. σD ,
so an equivalent limit to those of Eqs. (233) or (235) is simply
D √
σ = σD : σ D = σ D ∗ (236)

that is,
yielding
∗ p occurs if√the norm of the deviatoric tensor σD reaches the critical
value σ
D = 2J2∗ = 4µK∗ . In the three-dimensional setting, Eq. (235) (which
provides a lower physical interpretation of the criterion but facilitates its treatment)
can be written in the alternative forms (the reader can check it)
2 2 2
σ1D + σ2D + σ3D = 2J2∗ (237)
2 2 2
(σ1 − σ2 ) + (σ2 − σ3 ) + (σ3 − σ1 ) = 6J2∗ (238)
2 2 2 2 2 2
(σxx − σyy ) + (σyy − σzz ) + (σzz − σxx ) + 6σxy + 6σyz + 6σzx = 6J2∗ (239)

where the two former are expressed in principal directions of stress and the latter
in an arbitrary system of reference. If we consider now a plane stress state, Eq.
(238) gives the expression of the von Mises yield criterion applied to biaxial states
in terms of the total principal stresses σ1 and σ2 , i.e.

σ12 − σ1 σ2 + σ22 = σy2 (240)

where the value σy2 = 3J2∗ has been obtained through the particularization of Eq.
(238) at the yield point in the uniaxial tensile test (σ1 = σy , σ2 = σ3 = 0). The
failure envelope Eq. (240)
√ is represented
p by an ellipse in Figure 23, with semi-major
and semi-minor axes 2σy and 2/3σy , respectively. This last curve encloses the
maximum shear envelope, so the von Mises failure criterion is less restrictive (for
design purposes) than the Tresca failure criterion. As a matter of fact, conventional
metals generally fails between these two limits, that is after the maximum shear
failure criterion is reached and before the maximum distortional energy criterion
is reached.
√ Finally,
√ note that the von Mises rule predicts the relation S = τy =
σy / 3 = X / 3.

Unidirectionally reinforced laminae


Now, we focus on the strength criteria to be used for orthotropic laminae. In this
case, the formulation become far more complicated since the mechanical properties
of the material are direction-dependent. As indicated above, the allowable stresses
of an orthotropic lamina are usually known in, and only in, principal material axes.
Hence, it follows that the strength failure criteria should be expressed in that specific

- 65-
2 Mechanical Analysis of a Lamina

basis. This is an essential difference with the foregoing isotropic formulation, where
the allowable stresses were the same in every direction, so we were able to work in
the space of principal stresses, regardless their orientation in the space. Aside, it has
to be taken into account that composite materials usually fail in a brittle manner,
so the fundamental strengths Xt , Xc , Yt , Yc and S are generally associated to the
fracture of the lamina, whatever the phenomenon involved, and yielding could not
occur. In clear contrast, for metals we associated the axial and shear strengths, X
and S, to the axial and shear yield strengths, σY and τY , respectively.
By virtue of the current weight this topic is reaching in the industry, this section
may be arguably regarded one of the most important ones in this text. However,
although widely used by practitioners, the presented formulations should not be
considered as the definitive failure criteria for composite laminae since it is, certainly,
an active field of research in constant development.

Maximum Stress Failure Criterion The easiest criterion to consider is the


maximum stress failure criterion, which represents a generalization of the previous
isotropic one. It predicts that the composite lamina fails if one of the following
mutually independent conditions is reached
σxx = Xt , σxx = Xc , σyy = Yt , σyy = Yc , |σxy | = S (241)
where the stresses are expressed in principal material axes and a plane stress state
is considered. In contrast to its isotropic counterpart, namely Eq. (231), the failure
envelope Eq. (241) must be represented in the three-dimensional space of in-plane,
referred to X, stresses, as shown in Figure 24. Hence, according to the maximum
stress failure criterion, permissible in-plane stress states are within this rectangular
cuboid, which shows no interaction between the five modes of failure (any plane
defining this surface can be translated without affecting the other planes).

Maximum Strain Failure Criterion A similar analysis is provided by the max-


imum strain failure criterion, where the strains rather than the stresses have to
remain within specific ranges. A lamina subjected to a generic plane stress state is
said to fail if one of the following mutually independent conditions is reached
εxx = Xtε , εxx = Xcε , εyy = Ytε , εyy = Ycε , |γxy | = S ε (242)
where the fundamental allowable strains Xtε , Xcε , Ytε , Ycε and S ε are obtained from
the allowable stresses Xt , Xc , Yt , Yc and S, all of them being defined in the preferred
material directions. First, the separate consideration of the two first equations of
Eq. (168) for the corresponding uniaxial tests in tension and compression along the
principal in-plane material directions provides
Xt Xc Yt Yc
Xtε = , Xcε = , Ytε = , Ycε = (243)
Ex Ex Ey Ey
On the other hand, from a pure shear test performed in principal in-plane material
axes, the third equation in Eq. (168) yields
S
Sε = (244)
Gxy

- 66-
2.6 Failure criteria for composite laminae

σxy
S
S

σyy
Xc

Xt
σxx
Yc Yt
Figure 24: Maximum Stress failure envelope represented in the space of stresses in
principal material axes X for a Graphite/Epoxy lamina (reproduced using data from
Table 6).

Figure 25 shows the failure surface represented in the space of the in-plane strains
εxx , εyy and εxy . Although each fundamental stress allowable has been directly
related to its corresponding strain allowable, without any type of coupling affecting
them, note that the Poisson’s couplings distort this strain-based rectangular cuboid
when it is represented in the stress space. Accordingly, the maximum stress failure
envelope and the maximum strain failure envelope are not identical when they are
represented in the stress space. That is, for a given plane stress state, failure may
be reached according to the former and may not be reached according to the latter,
or vice versa.

Tsai-Hill Failure Criterion The generalization of the von Mises yield criterion to
the orthotropic case leads to the so-called Hill’s criterion (Ref. [6]). This orthotropic
failure criterion is readily introduced if one rewrites the von Mises isotropic criterion
of Eq. (235) using the relation given in Eq. (221) along with Eq. (222), i.e.

2σy2
σD : σ D = (σ : P) : (P : σ) = σ : P : σ = (245)
3
where both the symmetry property P : σ = σ : P and the identity P : P = P
have been used —note that with the matrix [P] expressed as in Eq. (222), P :
σ = [P] {σ} but σ : P : σ 6= {σ}T [P] {σ}; in this last case a factor of 2 is to
be included in the shear components of [P]. That way, the von Mises criterion is
mathematically expressed as a quadratic form of the stress tensor σ, where P is the
tensorial representation of the associated bilinear symmetric form. Essentially, we
say that Eq. (245) is an isotropic failure condition because the deviatoric tensor

- 67-
2 Mechanical Analysis of a Lamina

εxy

εyy
εxx

Xcε
Ycε Ytε
Xt ε

Figure 25: Maximum Strain failure envelope represented in the space of strains in
principal material axes X for a Graphite/Epoxy lamina (reproduced using data from
Table 6).

- 68-
2.6 Failure criteria for composite laminae

P is an isotropic fourth-order tensor, which representation is independent of the


orientation of the basis in which Eq. (245) is expressed. However, this fact is no
longer true for an orthotropic material because the strengths in different directions
take different (and independent) values, as we already know. Thus, we can generalize
Eq. (245) to account for different strengths in different directions and rewrite it for
an orthotropic material as
 2σ 2
σ D : L̄ : σ D = σ : P : L̄ : P : σ = σ : L : σ = ∗ (246)
3
where σ∗ is a reference (critical) stress and L̄ contains the direction-dependent weight
factors. The tensor L̄ is expressed in principal material coordinates in terms of six
independent weight factors as
 
9L1 0 0 0 0 0
 0 9L2 0 0 0 0 
 
 0 0 9L3 0 0 0 
[L̄] = 
  (247)
 0 0 0 L4 0 0 
 0 0 0 0 L5 0 
0 0 0 0 0 L6

where the factor of 9 is included for convenience. From a mathematical viewpoint,


note that the operation σ D : L̄ : σ D represents the squared norm of σ D weighted
by L̄. For an isotropic material we identify σ∗ = σy and 9L1 = 9L2 = 9L3 = L4 =
L5 = L6 = 1 and the isotropic criterion of Eq. (245), unweighted squared norm of
σ D , is retrieved. For the general case, the matrix representation of L = P : L̄ : P
(algebraically consistent with σ : L : σ = {σ}T [L] {σ}) is
 
L7 + L8 −L7 −L8 0 0 0

 −L7 L7 + L9 −L9 0 0 0 

 −L8 −L9 L8 + L9 0 0 0 
[L] =   (248)

 0 0 0 2L4 0 0 

 0 0 0 0 2L5 0 
0 0 0 0 0 2L6
with

L7 = 2L2 + 2L1 − L3 , L8 = 2L1 + 2L3 − L2 , L9 = 2L2 + 2L3 − L1 (249)

By construction of L = P : L̄ : P, the sum of the entries of each one of the first


three rows (or columns) of the matrix [L] in Eq. (248) is zero. Then we say that
the tensor L is a fourth-order orthotropic deviatoric tensor since the product σ :
L : σ vanishes for pure hydrostatic stresses σ = σ V = pI. However, contrary to
isotropic materials, we remark that we can generate distortions of the infinitesimal
volume of an orthotropic material, together with simultaneous dilatations, with the
single application of a hydrostatic pressure —you can check this assertion using the
relations {ε} = [S] {σ}, Eq. (144) and σ = pI. That is, distortion and dilatation
are coupled for a generic orthotropic material. Hence, although Eq. (246) has been
originally derived from deviatoric stresses, one should be aware of the fact that the

- 69-
2 Mechanical Analysis of a Lamina

Hill’s criterion does not represent a true distortion energy failure criterion since the
operation σ : L : σ is not strictly related to distorted deformations. The orthotropic
failure criterion Eq. (246) is expressed in principal material axes as

L7 (σxx − σyy )2 + L8 (σzz − σxx )2 + L9 (σyy − σzz )2


2 (250)
+ 2L4 σxy2 + 2L σ 2 + 2L σ 2 = 2σ∗
5 yz 6 zx
3
For further consideration, we note that only six of the seven constants appearing in
Eq. (250) are independent. After normalizing the coefficients Li using the critical
value 2σ∗2 /3, we finally rewrite it as

F (σyy − σzz )2 + G (σzz − σxx )2 + H (σxx − σyy )2


2 + 2M σ 2 + 2N σ 2 = 1 (251)
+ 2Lσyz zx xy

which represents the original yield failure envelope proposed by Hill [6] in the con-
text of orthotropic metal plasticity (compare to the isotropic formulation of Eq.
(239)). The six material constants F , G, H, L, M and N are widely referred to in
the literature because are those originally used by Hill. These material constants
can be determined in terms of the six yield strengths X , Y, Z, Sxy , Syz and Szx by
the corresponding uniaxial (F, G, H) and pure shear (L, M, N ) loading cases. Tsai
[7] regarded these last parameters as strength-like failure parameters (they somehow
represent limits of linear elastic behavior) and used this formulation to predict the
fracture of composite laminas, hence the name Tsai-Hill failure criterion which is
often given to this criterion. For practical purposes, Tsai also considered the unidi-
rectionally reinforced lamina as a transversely isotropic material, the axis Ox being
the preferred anisotropic direction. Thus, the subscripts y and z are interchangeable
in Eq. (251) and the relation G = H, among others, is obtained for this type of
laminae. Using the relation G = H and considering a plane stress state throughout
the lamina, Eq. (251) specializes to the following expression
2 2 2
2Hσxx + (F + H) σyy − 2Hσxx σyy + 2N σxy =1 (252)

which limits the permissible in-plane stresses to be applied in a transversely isotropic


reinforced lamina. Particularizing this last expression separately, first, to a uniaxial
tensile test performed along the fiber direction Ox, second, to a uniaxial tensile test
about the perpendicular direction Oy and, then, to a pure shear test in the principal
plane Oxy, we arrive to the most common expression of the Tsai-Hill failure criterion
for in-plane stresses
2 2
σyy 2
σxy
σxx σxx σyy
+ − + =1 (253)
X2 Y2 X2 S2
Note that the von Mises criterion Eq. (240) is recovered if both isotropic behavior
(X = Y = σy ) and principal stress components are used in Eq. (253). For materials
with tensile and compressive axial strengths being equal, that is Xt = −Xc = X
and Yt = −Yc = Y with the usual relation X > Y, the representation of the failure
surface Eq. (253) in the three-dimensional space of in-plane stresses is an ellipsoid,
with its center at the origin. However, composite materials usually present different

- 70-
2.6 Failure criteria for composite laminae

Figure 26: Tsai-Hill failure envelope represented in the space of stresses in principal
material axes X for a Graphite/Epoxy lamina (reproduced using data from Table
6). Four different ellipsoids are needed.

values of strength for tension and compression, so a piecewise ellipsoid (continuous


in value at the vertical planes separating the quadrants, but not in slope) is then
obtained, as it is shown in Figure 26. That is, four different envelopes are to be
considered in order to assess this failure criterion
2 2
σyy 2
σxy
σxx σxx σyy
+ − + =1 if σxx > 0 and σyy > 0
Xt2 Yt2 Xt2 S2

2 2
σyy 2
σxy
σxx σxx σyy
+ − + =1 if σxx < 0 and σyy > 0
Xc2 Yt2 Xc2 S2
(254)
2 2
σyy 2
σxy
σxx σxx σyy
+ − + =1 if σxx < 0 and σyy < 0
Xc2 Yc2 Xc2 S2

2 2
σyy 2
σxy
σxx σxx σyy
+ − + =1 if σxx > 0 and σyy < 0
Xt2 Yc2 Xt2 S2
From the observation of the surface in Figure 26, and in contrast to the maximum
stress and strain criteria, it is apparent that the failure strengths interact themselves
(are coupled) in the Tsai-Hill criterion.

Hoffman Failure Criterion The original Hill’s formulation has found a fertile
ground in metal plasticity. However, when different strengths are considered for

- 71-
2 Mechanical Analysis of a Lamina

tension and compression, its applicability (Tsai-Hill criterion) is clearly lessened


due to the necessary use of different surfaces in separate quadrants of the in-plane
stresses space. It is clear that in order to account for different tensile and compressive
strengths with a single continuously differentiable failure ellipsoid, its center cannot
be located at the origin of the in-plane stresses. We can obtain such a surface
modifying Eq. (246) by means of

  2σ 2
σ D − αD : L̄ : σD − αD = ∗ (255)
3

where αD is a material-dependent deviatoric tensor that represents the new “center”


of the variety Eq. (255) in the sixth-dimensional space of deviatoric stresses. In
elastoplasticity, αD is known as the back-stress tensor and is used as a mean to
model the Bauschinger effect in reverse loading. In contrast, αD is used herein as
a static material parameter derived from material experimental data. Using the
symmetries of L̄, Eq. (255) can be written as

2σ∗2
σ D : L̄ : σ D − 2αD : L̄ : σ D = − αD : L̄ : αD = K∗ (256)
3
Then, we arrive to an equation containing both a quadratic term and a linear term
of the total stresses, i.e.
σ : L : σ + B D : σ = K∗ (257)
where K∗ is a new reference (critical) value, L is the fourth-order tensor given in Eq.
(248) and B D = −2αD : L is a second-order deviatoric tensor. Analogously as we
did for L̄, the principal directions of αD are assumed to be aligned with the preferred
material directions of the lamina because of material symmetry requirements. Hence,
since the principal basis of L is the same, the principal directions of B D results to
be the preferred material directions of the lamina. As a result, the representation
of B D in principal material axes is diagonal and depends on two material constants
only  
B1 0 0
[B D ] =  0 B2 0  (258)
0 0 −B1 − B2
Note that, like the von Mises Eq. (245) and Hill Eq. (246) criteria, the left hand
side of Eq. (257) vanishes for pure volumetric stresses, that is, these three equations
depend only on deviatoric stresses.
Certainly, composite materials behave in a different manner than metals. Using
experimental data, it can be seen that the previous deviatoric criteria provide rela-
tively better approximations for metals than for composites. Thus, the assumption
that only deviatoric stresses contribute to yielding (experimentally verified for met-
als) may not be so accurate in the context of a failure criterion for fiber-reinforced
materials. With this fact in mind, a possible modification of Eq. (257) is simply to
consider a generic non-deviatoric second-order tensor B, i.e.

σ : L : σ + B : σ = K∗ (259)

- 72-
2.6 Failure criteria for composite laminae

which is a failure criteria containing a deviatoric quadratic term and a deviatoric-


hydrostatic linear term. In this case, the tensor B contains three independent
components when it is represented in the preferred axes of orthotropy
 
B1 0 0
[B] =  0 B2 0  (260)
0 0 B3

Following similar steps as for the derivation of Eq. (251) starting from Eq. (246), we
arrive to the expression originally proposed by Hoffman [8], known as the Hoffman
failure criterion
F (σyy − σzz )2 + G (σzz − σxx )2 + H (σxx − σyy )2
2 + 2M σ 2 + 2N σ 2 + P σ (261)
+ 2Lσyz zx xy xx + Qσyy + Rσzz = 1

The nine material constants appearing in Eq. (261) are now to be determined in
terms of the nine possible orthotropic strengths Xt , Xc , Yt , Yc , Zt , Zc , Sxy , Syz
and Szx by means of three uniaxial tensile tests and three uniaxial compressive tests
(F, G, H, P, Q, R) and three pure shear tests (L, M, N ). If we assume a pure isotropic
behavior of the lamina within the transverse plane Oyz, then G = H (among other
relations), and we obtain for a plane state of stresses (σzx = σyz = σzz = 0) the
in-plane failure condition
2 2 2
2Hσxx + (F + H) σyy − 2Hσxx σyy + 2N σxy + P σxx + Qσyy = 1 (262)

with five material constants to be related to the five in-plane strengths Xt , Xc , Yt ,


Yc and Sxy = S. After a simple identification of constants exercise, we arrive to the
common expression of the Hoffman failure criterion for in-plane stresses
2 2
σyy 2
σxy
σxx σxx σyy Xt + Xc Yt + Yc
− − + + σxx + σyy + 2 = 1 (263)
Xt Xc Y t Y c Xt Xc Xt Xc Yt Yc S
where algebraic values are considered for the compressive strengths, that is, Xc < 0
and Yc < 0. Note that the Hill’s criterion Eq. (253) is recovered if Xt = −Xc = X
and Yt = −Yc = Y. Moreover, when the lamina is assumed to be isotropic but with
different tensile and compressive strengths, then Xt = Yt = σyt and Xc = Yc = σyc in
every direction. In this case, Eq. (263) can be expressed in principal directions of
stress to give a generalization of the von Mises criterion for in-plane stresses, namely
Eq. (240), to this type of isotropic materials with different strength in tension and
compression
 
σ12 + σ22 − σ1 σ2 − σyt + σyc σ1 − σyt + σyc σ2 = −σyc σyt (264)

which is represented in the space of principal stresses by a single smooth ellipse.


The relation S 2 = τy2 = −σyc σyt /3 = −Xc Xt /3 is predicted by Eq. (264).
The continuously differentiable (anisotropic) failure equation given by Eq. (263)
is represented in the space of in-plane stresses by a single ellipsoid. As mentioned,
this is a clear advantage with respect to the Tsai-Hill failure criterion (defined by
means of a piecewise failure surface) in design use. In Figure 27 the contour levels

- 73-
2 Mechanical Analysis of a Lamina

σyy
σxy = S σxy = 0
σxx

Figure 27: Contour levels (ellipses) of the Hoffman failure envelope (ellipsoid) rep-
resented in the space of stresses in principal material axes X for a Graphite/Epoxy
lamina (reproduced using data from Table 6).

of a Hoffman ellipsoid are shown. Note that the center of the ellipses are no longer
at the origin of the stress axes due to the presence of the linear terms in Eq. (263).
The coupling term σxx σyy /Xt Xc causes the axes of the ellipses to be slightly rotated
with respect to the stress (preferred) axes.

Tsai-Wu Failure Criterion The next strength formulation we consider is the


Tsai-Wu failure criterion. Some authors refer to this criterion as the tensor fail-
ure criterion. Certainly, as we show below, this condition is expressed in tensorial
form, but the Tsai-Hill and the Hoffman failure criteria have been also formulated
in tensorial form, so this denomination is somehow inappropriate. Tsai and Wu [9]
generalized the Hoffman tensorial condition of Eq. (259) by means of the consider-
ation of both deviatoric and volumetric stress contributions in the quadratic term.
That is, they proposed the failure criterion

σ : M : σ + B : σ = K∗ (265)

where M is a fourth-order orthotropic tensor with nine independent principal com-


ponents (compare to the six independent components of the deviatoric tensor L)
 
M11 M12 M13 0 0 0

 M12 M22 M23 0 0 0 

 M13 M23 M33 0 0 0 
[M] =   (266)

 0 0 0 2M44 0 0 

 0 0 0 0 2M55 0 
0 0 0 0 0 2M66

and B is the tensor given in Eq. (258) in principal material coordinates. In the
sixth-dimensional space of stresses, it is clear that Eq. (265) depends on a total of 12
material constants. Then three more tests (additional to the nine mentioned above
for the Hoffman criteria) are needed in order to completely determine the Tsai-Wu
model. If all these necessary experimental data were available, the Tsai-Wu failure
criterion could better characterize the overall behavior of an orthotropic material
and would provide closer approximations in a wider range of situations. However,
note that this is a huge task, which will not be always possible. A more feasible
situation is obtained when Eq. (265) is particularized to plane stress and properly

- 74-
2.7 Determination of the overall behavior of a lamina through the
behavior of its constituents. Strength of materials approach.

σyy σxy = 0

σxx

(F12 )H F12 = 2(F12 )H F12 = 1/2(F12 )H F12 = 0

Figure 28: Contour levels σxy = 0 of the corresponding Tsai-Wu failure envelopes
for a Graphite/Epoxy lamina computed with different values of the parameter F12
(reproduced using data from Table 6). Representation in the space of stresses in
principal material axes X. The ellipse in black color corresponds to the Hoffman
failure criterion (F12 )H ≡ (F12 )Hof f man = 1/(2Xt Xc ).

normalized as in Eq. (261)


2 2 2
F11 σxx + F22 σyy + 2F12 σxx σyy + 2F44 σxy + F1 σxx + F2 σyy = 1 (267)

This expression contains six material constants to be determined through mechan-


ical testing. In this case, the independent material characteristics Fij = Mij /K∗
and Fi = Bi /K∗ have been used in Eq. (267). The main difference between Eq.
(262) and Eq. (267) is that for the former expression the coefficients of the terms
2 and σ σ
σxx xx yy are directly related due to the assumption of transversely isotropic
behavior, while for the latter expression they are utterly independent even for pure
transversely isotropic laminae. Specializing Eq. (267) to four uniaxial tests (tension
and compression along the preferred axes Ox and Oy) and a single pure shear test
(in the plane Oxy), we obtain
2 2
σyy 2
σxy
σxx Xt + Xc Yt + Yc
− − + 2F12 σxx σyy + σxx + σyy + 2 = 1 (268)
Xt Xc Y t Y c Xt Xc Yt Yc S
and F12 still remains undetermined (cf. Hoffman criterion of Eq. (263)). This
material parameter can be calculated from (sometimes expensive) biaxial or off-axis
uniaxial tests, in which σxx and σyy are simultaneously distinct to zero. The influence
of the value F12 in Eq. (268) is shown in Figure 28, where different failure curves,
with σxy = 0, are represented. We remark that the value taken by F12 provides little
differences regarding the centre, size and orientation of the corresponding failure
ellipsoids. This little influence of F12 has led to some authors [10] to simply assume
F12 = 0 in the in-plane Tsai-Wu condition of Eq. (268).

2.7 Determination of the overall behavior of a lamina through the


behavior of its constituents. Strength of materials approach.
In the previous section, composite laminae have been characterized as anisotropic
materials from a macroscopic standpoint. Both elastic and strength directional re-
sponses have been described without taking into account the specific microstructure

- 75-
2 Mechanical Analysis of a Lamina

of the material. That way, the foregoing formulations are not only valid for typical
fiber reinforced materials, but also for any other material which behaves anisotrop-
ically in macroscopic terms. Among a variety of possible examples, metal sheets
show anisotropic elastic properties due to the technological process involved in the
material production. Hence, the formulations derived above can be used to analyze
their mechanical response to generic loads.
This section is devoted to examine the specific behavior of unidirectionally rein-
forced laminae when they are seen as materials formed with two perfectly bonded
constituents, namely matrix and fibers. First of all, it will be assumed that the me-
chanical properties of the matrix itself, on one hand, and of the fibers themselves,
on the other hand, are known. Then, the question that naturally arises is: Can the
macroscopic response of the composite material formed by the fibers and the matrix
be described in terms of the mechanical properties of its constituents? We will see
that, with some assumptions at hand, it will be possible to infer that behavior to
some extent. Furthermore, regardless of the goodness of the quantitative approxi-
mations being reached, we will be able to conclude some interesting results which
will be relevant in the design process from a qualitative point of view.
The most simplistic approximation to the overall behavior of a lamina through
the behavior of its constituents is obtained from the strength of materials theory.
Using this basic approach, we are interested herein on the analytical prediction of
the elastic behavior of the composite lamina in terms of the elastic properties of the
matrix and the fibers. Both matrix and fibers are assumed to be homogeneous and
to have local isotropic properties. Moreover, the bond between the fibers and the
matrix is assumed to be initially perfect and to be preserved when the lamina is
deformed.

Young’s modulus in the fiber direction Ex


In Figure 29, an sketch of a piece of a unidirectionally reinforced lamina containing
the matrix material and an isolated fiber perfectly bonded together is shown. In that
figure, the deformed configuration when this representative volume is subjected to
the normal stress σxx = σ is also represented. According to the interpretation given
in Eq. (10) the macroscopic axial infinitesimal strain εxx = ε, which is equal to the
axial strains of fibers and matrix, is simply
∆L
εf = εm = ε = (269)
L
In this basic study, we assume that the transverse and through-the-thickness
stresses vanish throughout the whole lamina, i.e. σyy = σzz = 0. Although σzz = 0
is arguably a good approximation (it is the plane stress condition itself), note that
the enforcement of σyy = 0 is more questionable. Then, the isotropic Hooke’s law
Eq. (49) separately applied to the matrix and the fibers yield
σf = Ef εf = Ef ε σm = Em εm = Em ε (270)
where Ef and Em are the Young’s moduli of the fiber and matrix materials, respec-
tively. If we consider the volume as a unique entity, then we can also write
σ = Ex ε (271)

- 76-
2.7 Determination of the overall behavior of a lamina through the
behavior of its constituents. Strength of materials approach.

ΔW
y
W σ x σ

L ΔL
Figure 29: Initial and infinitesimally deformed representative volume of a unidi-
rectionally reinforced lamina subjected to uniaxial tension in the direction of the
fibers.

where Ex represents the macroscopic Young’s modulus of the composite lamina


in the fiber direction. Note that we deliberately specify the subscript x in the
corresponding Young’s modulus of the anisotropic lamina but we do not need to
specify any subscript in the Young’s moduli of its isotropic constituents. Our aim
here is to obtain an analytical expression for the mean property Ex in terms of Ef
and Em . Since the stresses given in Eq. (270) are respectively applied to different
points of the lamina, they cannot be directly related. However, let us denote by Af
and Am the cross-sectional area of the deformed lamina being occupied by the fibers
and by the matrix, respectively, and by A = Af + Am the total area of the deformed
transverse section. Now, we can readily relate the partial loads carried by the fibers,
Pf = σf Af , and by the matrix, Pm = σm Am , to the total axial load applied to the
lamina, P = σA, through

P = Pf + Pm (272)
σA = σf Af + σm Am (273)

so
Af Am
σ = σf + σm = σf V̂f + σm V̂m (274)
A A
where V̂f = Vf /V = Af /A and V̂m = Vm /V = Am /A are the volume fractions of
fibers and matrix, respectively. Note that V̂f + V̂m = 1. Substituting Eqs. (270)
and (271) into Eq. (274) we finally arrive to the sought relation

Ex = Ef V̂f + Em V̂m = Em + (Ef − Em ) V̂f (275)

which gives a linear dependence on the volume fraction V̂f for a given combination
of matrix-fiber materials usually known as the rule of mixtures for Ex . Note we only
consider infinitesimal deformations, so no distinction is needed between undeformed
and deformed configurations when computing areas and volumes. In Eq. (275)
both elastic properties of the constituents and geometrical parameters relating them
are needed to obtain the macroscopic Young’s modulus Ex . Both dependences,

- 77-
2 Mechanical Analysis of a Lamina

mechanical and geometrical, will be present throughout the following microstructure-


based derivations of macrostructure properties.
Although the intuitive Eq. (275) really seems very simple, it has been proven to
provide very good approximations for the averaged value Ex of several matrix-fiber
material systems for the whole range of fiber-volume fraction V̂f . An example of
such a correlation with actual experimental data is shown in Figure 30. Regarding
this specific graphic, it is remarkable that for the surroundings of the usual volume
fraction used in fiber-reinforced materials, namely V̂f ≈ 0.6, the resulting Young’s
modulus expressed in terms of Ef is approximately Ex ≈ 0.6Ef ∼ Ef , whereas if it
is expressed in terms of Em is Ex ≈ 60Em ≫ Em , which is a direct consequence of
the linear relation of Eq. (275) and the fact that Ef ≫ Em , as usual. Hence, the
elastic modulus in the fiber direction of a conventional reinforced lamina is of the
same order than the elastic modulus of the reinforcing material. Accordingly, we
say that Ex is a fiber-dominated property. Finally, it is straightforward to obtain
that the longitudinal iso-strain condition εxx = ε = εf = εm and the assumptions
σyy = σzz = 0 yield for the transverse and through-the-thickness strains of the fiber
and the matrix
εyyf = εzzf = −νf εxxf = −νf εf = −νf ε (276)
εyym = εzzm = −νm εxxm = −νm εm = −νm ε (277)
where νf and νm are the Poisson’s ratios of the isotropic constituents. Aside, the
conditions of a perfect bond between fibers and matrix and the pure transverse
isotropic behavior of the lamina of Figure 29 imply that

εyyf = εyym εzzf = εzzm (278)

so the strength of materials formulation presumes the following underlying relation


for the Poisson’s ratios of the isotropic constituent materials

νf = νm (279)

Of course, the requirement of Eq. (279) is not satisfied for a generic composite
lamina. However, as we will see below the Poisson’s ratios νf and νm take very
similar values in practical cases and the mechanics of materials prediction given in
Eq. (275) can be regarded an acceptable approximation of Ex .

Young’s modulus in the transverse direction Ey


The other macroscopic elastic modulus that we can determine from this basic ap-
proach is Ey , that is, the Young’s modulus in the in-plane transverse direction to
the fibers. Figure 31 shows the deformed configuration which is consistent with the
aforementioned premises. First, by axial equilibrium in direction Oy

σf = σm = σ = σyy (280)

and by means of the associated Hooke’s laws (upon taking the assumptions σxx =
σzz = 0 throughout the whole lamina)

Ef εf = Em εm = Ey ε (281)

- 78-
2.7 Determination of the overall behavior of a lamina through the
behavior of its constituents. Strength of materials approach.

400 Ex (G Pa) Ef = 410


(Boron)

300

200

100

Em = 4.1
0 0.2 0.4 0.6 0.8 1
(Epoxy)
Vf

Figure 30: Rule of mixtures for Ex (Eq. (275)) contrasted to experimental data for
a Boron/Epoxy material system. Adapted from Ref. [4].

σ
ΔW
y
x W

σ
L ΔL
Figure 31: Initial and infinitesimally deformed representative volume of a unidirec-
tionally reinforced lamina subjected to uniaxial tension in the transverse direction
to the fibers.

- 79-
2 Mechanical Analysis of a Lamina

In order to obtain Ey in terms of Ef and Em , we previously need to relate the


strains in the transverse direction undergone by the fibers and the matrix εf and
εm to the mean (total) strain ε. The assumption of perfect bond between fibers and
matrix yields the relation
∆W = ∆Wf + ∆Wm (282)
so
∆W ∆Wf + ∆Wm ∆Wf Wf ∆Wm Wm Wf Wm
ε= = = + = εf + εm (283)
W W Wf W Wm W W W

If we consider now the approximations Wf /W ≈ V̂f and Wm /W ≈ V̂m , which are


not strictly exact because the configuration shown in Figure 31 is not uniform along
the Oz direction within the lamina, we obtain the approximated expression for ε in
terms of εf , εm and the (initial) volume fractions

ε ≈ εf V̂f + εm V̂m (284)

Then, using the two relations contained in Eq. (281) and the expression for ε given
in Eq. (284), it is straightforward to obtain the result
 
1 V̂f V̂m 1 1 1
= + = − − V̂f (285)
Ey Ef Em Em Em Ef
which is a rule of mixtures for the compliance 1/Ey in terms of the compliances
1/Ef and 1/Em represented as a linear relation on the fiber-volume fraction V̂f , as
shown in Figure 32.a. The explicit expression for Ey is
Ef Em 1
Ey = =   (286)
Em V̂f + Ef V̂m 1 1 1
− − V̂f
Em Em Ef

which is represented in Figure 32.b as a function of V̂f by a hyperbola with vertical


asymptote at 1/(1 − Em /Ef ) & 1. From the observation of that graphic, and in
clear contrast to Ex , we note that Ey ∼ Em for conventional composite laminae
(Ef ≫ Em and vf ≈ 0.6), that is, the elastic modulus in the transverse direction to
the fibers of a conventional reinforced lamina is of the same order than the elastic
modulus of the matrix. We say then that Ey is a matrix-dominated property.
The comparison of predicted values given by Eq. (286) with actual experimental
data is also shown in Figure 32.b. The agreement is not so good as for Ex , so other
more elaborated microstructure-based formulations are to be derived for Ey . A clear
difference between the present formulation for Ey and the previous one for Ex is that
Eq. (284) used herein is an approximation while its homologous expression used
above, i.e. Eq. (274), is exact. After proposing a geometrical equivalence taking
into account the effects in the z-direction, the so-called modified rule-of-mixtures
model (see Ref. [12]) considers a representative volume with modified fiber and
matrix widths W̄f and W̄m such that
q q
W̄f W̄m W − W̄f
= V̂f > V̂f ; = = 1 − V̂f < V̂m (287)
W W W

- 80-
2.7 Determination of the overall behavior of a lamina through the
behavior of its constituents. Strength of materials approach.

1/Em 1/Ey

a)

1/Ef
0 0.2 0.4 0.6 0.8 1

Ey (GPa) Ef = 73
60

b)
40

20

Em = 3.45
0 0.2 0.4 0.6 0.8 1

Ey (GPa) Ef = 73
60

c) 40

20

Em = 3.45
0 0.2 0.4 0.6 0.8 1
Vf

Figure 32: a) Rule of mixtures for 1/Ey , Eq. (285). b) Representation of Ey as a


function of the fiber-volume fraction V̂f , Eq. (286). c) Modified representation of Ey
as a function of the fiber-volume fraction V̂f making use of Eq. (287). Experimental
data for a Glass/Epoxy material system from Ref. [11].

- 81-
2 Mechanical Analysis of a Lamina

It can be observed in Figure 32.c that the prediction capability is improved if this
simple modification of the rule of mixtures is considered in Eqs. (283)–(286).
Moreover, taking the assumed stress field, σyy = σf = σm = σ and σxx = σzz =
0, to be strictly true has another underlying implication. Considering this stress
field, the longitudinal (in the direction of the fibers) strains associated to the matrix
and the fibers are
νf νf
εxxf = − σyyf = − σ (288)
Ef Ef
νm νm
εxxm =− σyym = − σ (289)
Em Em

Then, if a perfect match of displacements is required across the fiber-matrix bound-


aries of the lamina, as depicted in Figure 31, the following compatibility equation
must hold
∆Lf ∆L ∆Lm
εxxf = = = = εxxm (290)
Lf L Lm
so
νf νm
= (291)
Ef Em
which is an implicit relation for the transverse compliances S12 = −ν/E of the
constituents being included in the strength of materials solution for Ey . However,
we already know that Ef ≫ Em and νf ∼ νm . Thus, Eq. (291) is far to be
satisfied for a conventional fiber-matrix material system, which may explain the
significant deviations observed between the mechanics of materials predictions and
the experimental data. Note that both relations Eq. (279) and Eq. (291) cannot be
simultaneously fulfilled unless νf = νm and Ef = Em , that is, unless the constituents
have the same elastic properties (resulting into a macroscopically isotropic material),
which directly opposes to the underlying idea for an effective composite material.
Simple models of Eqs. (275) and (285) using spring arrangements are shown
in Figure 33. These equivalent models give very intuitive interpretations of the
foregoing formulations. In the equivalent model for Ex , the springs are arranged in
parallel, so they deform the same amount but carry different loads. If the springs’
stiffnesses are such that Ef V̂f ≫ Em V̂m , then the fiber-spring bears much more load
than the matrix-spring. The response is fiber-dominated. In the equivalent model
for Ey , the springs are arranged in series, so they carry the same load but undergo
different deformations. If the springs’ stiffnesses are such that Ef /V̂f ≫ Em /V̂m ,
then the matrix-spring deforms much more than the fiber-spring. The response is
matrix-dominated.

In-plane Poisson’s ratio νxy


Recall that the in-plane elastic response of an orthotropic lamina is characterized
by the Young’s moduli Ex and Ey , the Poisson’s ratio νxy and the shear modulus
Gxy . Expressions for νxy and Gxy are still to be derived. As it was explained above
using the engineering expression of the compliance matrix given in Eq. (133), the
Poisson’s ratio νxy measures the coupling between the transverse strain εyy and the

- 82-
2.7 Determination of the overall behavior of a lamina through the
behavior of its constituents. Strength of materials approach.

a) Kf = Ef Vf
Ff Ff
Fx = Ff + Fm

Fx Fx Kx ΔL = Kf ΔL + Km ΔL

Fm Fm Kx = Kf + Km
Km = Em Vm

b)
Kf = Ef / Vf Km = Em / Vm ΔWy = ΔWf + ΔWm
Fy Fy F/Ky = F/Kf + F/Km
ΔWf ΔWm
1/Ky = 1/Kf + 1/Km

Figure 33: Spring models for the longitudinal (a) and transverse (b) micromechanical
elastic behaviors of a fiber reinforced lamina.

axial strain εxx when a single axial stress σxx is acting on the material. Then,
following the sketch of Figure 29, νxy reads
εyy ∆W/W
νxy = − =− (292)
εxx ∆L/L
From the direct observation of the initial and the deformed configurations, we deduce
the relations
L = Lf = Lm ∆L = ∆Lf = ∆Lm (293)
W = Wf + Wm ∆W = ∆Wf + ∆Wm (294)
Then, by simple substitution into Eq. (292) we obtain
∆Wf /W ∆Wm /W ∆Wf /Wf Wf ∆Wm /Wm Wm
νxy = − − =− − (295)
∆L/L ∆L/L ∆Lf /Lf W ∆Lm /Lm W

As for the derivation of Ey we consider the approximations Wf /W ≈ V̂f and


Wm /W ≈ V̂m . Then, we finally arrive to the following rule of mixtures for νxy
νxy = νf V̂f + νm V̂m (296)
where νf and νm are the Poisson’s ratios of the fiber and matrix materials, respec-
tively. For conventional fiber-reinforced materials the Poisson’s ratios νf and νm
take very similar values (between 0.25 and 0.35), so certainly νxy ∼ νf ∼ νm and
any of both constituents play a relevant role against the other one in the macroscopic
description of νxy . The agreement between predicted and measured values for νxy
is similar to that for Ex in Figure 30.

- 83-
2 Mechanical Analysis of a Lamina

In-plane shear modulus Gxy

The calculations performed to the determination of the shear modulus Gxy are, in all
terms, equivalent to those detailed above for the obtaining of the transverse elastic
modulus Ey . The deformation state to be analyzed is shown in Figure 34. The only
non-vanishing stress component acting on the representative volume is σxy = τ .
This stress component is presumed to act on both the fiber and the matrix, i.e.

τf = τm = τ (297)

The separate application of the Hooke’s law Eq. (49) to the fiber, the matrix and
the macroscopic element yields

Gf γf = Gm γm = Gxy γ (298)

Aside, the averaged angular distortion undergone by the piece of lamina, γ, can be
expressed in terms of the angular distortions of the fibers, γf , and the matrix, γm ,
through
u uf W f um Wm Wf Wm
γ= = + = γf + γm (299)
W Wf W Wm W W W
which is an interpretation only valid for infinitesimal strains and where we have used
the fact that u = uf + u1 + u1 = uf + um . Then, we may express γ by means of

γ ≈ γf V̂f + γm V̂m (300)

Equations (298) and (300) are identical in form to Equations (281) and (284). We
only change E by G, ε by γ and Ey by Gxy . Thus, the solution for Gxy is
 
1 V̂f V̂m 1 1 1
= + = − − V̂f (301)
Gxy Gf Gm Gm Gm Gf

or
1 Gf Gm
Gxy =   = (302)
1 1 1 Gm V̂f + Gf V̂m
− − V̂f
Gm Gm Gf
Thus, a rule of mixtures is obtained for the shear compliance 1/Gxy and a hyperbola
with vertical asymptote at 1/(1 − Gm /Gf ) & 1 represents the dependence of the
predicted shear modulus Gxy as a function of V̂f . Similarly to the above case for Ey ,
some deviation between predictions and experimental
q data has been found. Again,
the use of the modified fiber-volume fraction V̂f (modified rule of mixtures) instead
of V̂f in Eq. (302) provides a better prediction for measured values of Gxy . Aside,
since the relation Gf ≫ Gm usually holds, we can infer from this basic approach
that the resulting elastic shear modulus of a fiber-reinforced lamina Gxy is a matrix-
dominated property (as for Ey ). That is, the matrix shear deformation mostly
dominates the shear deformation of the composite material.

- 84-
2.8 Determination of the overall behavior of a lamina through the
behavior of its constituents. Theory of elasticity approach.

Figure 34: Initial and infinitesimally deformed representative volume of a unidirec-


tionally reinforced lamina subjected to simple shear in the plane Oxy.

2.8 Determination of the overall behavior of a lamina through the


behavior of its constituents. Theory of elasticity approach.
The strength of materials approach has demonstrated to provide just a mere approx-
imation to the definition of the in-plane elastic properties of a composite lamina in
terms of those of its fiber and matrix constituents. We have seen that some un-
usual matrix-fiber relations implicitly emerge as a consequence of the simplifying
assumptions on the stress and strain fields being considered. In some cases, the
approximations may be acceptable, in other cases a better refinement of the theory
seems to be necessary. In this section use the theory of elasticity in order to show
that the preceding strength of materials derivations provide an upper and a lower
bound for the true elastic macroscopic properties of the composite lamina. Then, we
introduce a curve-fitting procedure, the Halpin-Tsai equations, consisting of a set
of theory-of-elasticity-based equations which are commonly used to reproduce the
actual (experimental) elastic behavior of an orthotropic lamina in a very accurate
way.

2.8.1 Lower and upper bounds on preferred Young’s moduli


The definition of an upper bound and a lower bound on the value of the Young’s
modulus of a composite lamina (measured in a preferred direction) in terms of
the Young’s moduli and volume fractions of its constituents is a valuable result
for design purposes of composite structures. We can obtain such bounds invoking
well-known Energy Theorems defined within the theory of elasticity. First of all,
consider a composite material conformed by two different isotropic constituents, i.e.
a reinforcing material and a binding material. Assume that the reinforcement is
in the form of particles approximately equal in size (equiaxed), which are evenly
distributed within the matrix material (we keep on using the letters f and m in
the following derivation in order to refer to the reinforcement and to the matrix).
That way, the resulting composite is heterogeneous in the microscale, but behaves

- 85-
2 Mechanical Analysis of a Lamina

isotropically in the macroscale. The constituents are defined by their corresponding


Young’s moduli, Ef and Em , and Poisson’s ratios, νf and νm , which are different
relative to each other in the most general case.

Lower bound
A lamina of this material is subjected to a uniaxial tension test. The macroscopic
stress and strain fields in any point of the solid, in terms of the prescribed stress σ,
are
     

 σ 1   σ   1 
 0 
  
  0 
 



 −ν




 
 
 
 
 

    σ  
0 0 −ν
{σ} = = {ε} = [S] {σ} = (303)

 0  
 0 
 E  0 
     


 0 
  0 
  
 0 
    
 
 

0 0 0
where E and ν are the macroscopic Young’s modulus and Poisson’s ratio of the
isotropic composite material, respectively. The resulting complementary strain en-
ergy density is
1 1 1 σ2
wc (σ) = σ : S : σ = σ : ε (σ) = (304)
2 2 2E
—recall that the strain energy and the complementary strain energy densities co-
incide numerically for linear elasticity—, which integrated over the volume of the
lamina gives the total complementary strain energy
Z Z
c c c 1 σ2
W = w dV = w dV = V (305)
V V 2E
From physical grounds, this energy, obtained at the macroscopic scale, must coincide
with the value of the complementary strain energy calculated at the microscopic scale
1 σ2
Wc = V (306)
Z2 E Z
= wfc dVf + c
wm dVm (307)
Vf Vm
Z Z
1 1
= (σ f : Sf : σ f ) dVf + (σ m : Sm : σ m ) dVm (308)
2 Vf 2 Vm

where Sf and Sm are the isotropic compliance tensors of the constituents. In the
preceding section, we have analyzed two very special cases of an heterogeneous
composite lamina. In the first of them, we assumed εf = ε = εm throughout the
lamina and obtained σf 6= σ 6= σm , see Eqs. (269) and (270). In the second case, we
assumed σf = σ = σm throughout the lamina and obtained εf 6= ε 6= εm , see Eqs.
(280) and (281). If we consider herein a more general case in which both materials
interact themselves through a different manner than the two previous special cases,
then neither the stresses nor the strains of the constituents are necessarily equal to
the macroscopic stresses and strains (as given in Eq. (303)), respectively, that is
σ f 6= σ 6= σ m and εf 6= ε 6= εm (309)

- 86-
2.8 Determination of the overall behavior of a lamina through the
behavior of its constituents. Theory of elasticity approach.

and Eq. (308) remains to be determined for a generic composite lamina under a
tension test.
We specialize now the Principle of Minimum Complementary Energy to the
problem under study. Then, we know that among all the statically admissible stress
fields (i.e. in equilibrium with the actual –zero– body forces and satisfying the
traction boundary conditions of the tension test), the actual stress field makes the
complementary strain energy W c an absolute minimum. A statically admissible
stress field proposal at the microscale is simply
σ̃ f = σ and σ̃m = σ (310)
which is uniform. The associated complementary strain energy densities of the
constituents are
1 1 1 σ2
w̃fc = σ̃ f : Sf : σ̃f = σ : Sf : σ = (311)
2 2 2 Ef
c 1 1 1 σ2
w̃m = σ̃ m : Sm : σ̃ m = σ : Sm : σ = (312)
2 2 2 Em
Then
1 σ2
Wc = V (313)
Z2 E Z
= wfc dVf + c
wm dVm (314)
Vf Vm
Z Z
≤ w̃fc dVf + c
w̃m dVm (315)
Vf Vm
1 σ2 1 σ 2
= Vf + Vm (316)
2 Ef 2 Em
!
1 V̂f V̂m
= + σ2 V (317)
2 Ef Em
= W̃ c (318)
where we readily identify the inequality
1 V̂f V̂m
≤ + (319)
E Ef Em
or, equivalently
Ef Em
E≥ (320)
Em V̂f + Ef V̂m
This way, we have obtained a lower bound on the Young’s modulus E in terms of
Ef and Em (νf and νm do not appear in this analysis) of an heterogeneous isotropic
composite lamina which is coincident with the mechanics-of-materials Young’s mod-
ulus associated to the transverse direction of a unidirectionally reinforced composite
lamina, Eq. (286). However, the inequality Eq. (320) can also be interpreted to
be valid with E being any principal Young’s moduli of a fiber-reinforced composite
material.

- 87-
2 Mechanical Analysis of a Lamina

Upper bound
In order to obtain an upper bound on the Young’s modulus E of the lamina under
study, we have to deal with the Principle of Minimum Potential Energy instead of
the Principle of Minimum Complementary Energy used above. The proper applica-
tion of the present principle requires the macroscopic stress and strain fields to be
expressed in terms of the prescribed strain ε, i.e.
   

 ε1 
 
 ε 

 ε2 
 
 −νε 


 
 
 

   
ε3 −νε
{ε} = = (321)

 0 
 
 0 
 0 
    0  

 
  
    

0 0
 
C11 ε − 2νC12 ε 


    

 (1 − ν) C12 ε − νC11 ε   
 0 


    

 (1 − ν) C12 ε − νC11 ε  =
 0
{σ} = [C] {ε} = 
   0  (322)
 0    


 0
 
   0 

 
0 0

In other words, we prescribe the displacements (strains) on the faces of the lamina
and the stress field is obtained as a result. The macroscopic strain energy density is
1 1 1
w (ε) = ε : C : ε = ε : σ (ε) = Eε2 (323)
2 2 2
which integrated over the volume of the lamina gives the total strain energy
Z Z
1
W= wdV = w dV = Eε2 V (324)
V V 2

From physical grounds, this macroscopic value must coincide with the value of the
strain energy calculated at the microscopic scale
1 2
W= Eε V (325)
Z2 Z
= wf dVf + wm dVm (326)
Vf Vm
Z Z
1 1
= (εf : Cf : εf ) dVf + (εm : Cm : εm ) dVm (327)
2 Vf 2 Vm

where Cf and Cm are the isotropic stiffness tensors of the constituents. Similarly to
the preceding case, we know that, in general

σ f 6= σ 6= σ m and εf 6= ε 6= εm (328)

and Eq. (327) remains to be determined for the case under study.

- 88-
2.8 Determination of the overall behavior of a lamina through the
behavior of its constituents. Theory of elasticity approach.

We apply now the Principle of Minimum Potential Energy to this particular case.
Then, we know that among all the kinematically admissible strain fields (i.e. with
an associated, compatible, displacement field satisfying the prescribed displacement
boundary conditions of the tension test), the actual strain field makes the strain
energy W an absolute minimum. Consider the following proposal of a kinematically
admissible, uniform, strain field at the microscale

ε̃f = ε and ε̃m = ε (329)

The strain energy densities result into

1 1 1 1 − νf − 4νf ν + 2ν 2
w̃f = ε̃f : Cf : ε̃f = ε : Cf : ε = Ef ε2 (330)
2 2 2 1 − νf − 2νf2
1 1 1 1 − νm − 4νm ν + 2ν 2
w̃m = ε̃m : Cm : ε̃m = ε : Cm : ε = 2
Em ε2 (331)
2 2 2 1 − νm − 2νm
Then
1
W = Eε2 V (332)
Z2 Z
= wf dVf + wm dVm (333)
Vf Vm
Z Z
≤ w̃f dVf + w̃m dVm (334)
Vf Vm
1 1 − νf − 4νf ν + 2ν 2 1 1 − νm − 4νm ν + 2ν 2
= 2 Ef ε2 Vf + 2
Em ε2 Vm (335)
2 1 − νf − 2νf 2 1 − νm − 2νm
!
1 1 − νf − 4νf ν + 2ν 2 1 − νm − 4νm ν + 2ν 2
= Ef V̂f + Em V̂m ε2 V (336)
2 1 − νf − 2νf2 1 − νm − 2νm2

= W̃ (337)

which yields the following upper bound on the Young’s modulus E in terms of the
known values Ef , Em , νf , νm , V̂f , V̂m and the still unknown value ν

1 − νf − 4νf ν + 2ν 2 1 − νm − 4νm ν + 2ν 2
E≤ Ef V̂ f + Em V̂m (338)
1 − νf − 2νf2 2
1 − νm − 2νm

The actual value of the macroscopic Poisson’s ratio ν of the mixture is obtained
upon the consideration that, for given elastic and geometrical parameters of the
constituents, the strain energy Eq. (336) reaches effectively an absolute minimum.
After some algebra, the condition dW̃ (ν) /dν = 0 yields
  
1 − νm − 2νm2 ν E V̂ + 1 − ν − 2ν 2 ν E V̂
f f f f f m m m
ν=   (339)
2 ) E V̂ + 1 − ν − 2ν 2 E V̂
(1 − νm − 2νm f f f f m m

which provides an absolute minimum on W̃ due to the fact that d2 W̃ (ν) /dν 2 is
always positive when the constraints on the isotropic Poisson’s ratios νf < 1/2 and

- 89-
2 Mechanical Analysis of a Lamina

νm < 1/2 are introduced. The substitution of Eq. (339) into Eq. (338) gives the
final expression of the upper bound on E in terms of Ef , Em , νf and νm .
Recall that the implicit condition νf = νm , Eq. (279), was obtained during
the strength of materials approach to the longitudinal Young’s modulus Ex of a
fiber reinforced lamina, Eq. (275), as a result of prescribing the longitudinal strains
εxx = ε and the transverse stresses σyy = σzz = 0 for both fibers and matrix.
However, for the upper bound derived herein we have prescribed the longitudinal
strains εxx = ε and also the transverse strains εyy = εzz = −νε for both constituents,
see Eq. (329) along with Eq. (321). The stress field at the microscopic scale
corresponding to the upper bound case is then derived from the prescribed strains.
We obtain
 
 1 − νf − 2νf ν

 2 Ef ε 




 1 − νf − 2νf 


 


 ν f − ν 



 2 Ef ε 



 1 − ν f − 2ν f



{σ̃ f } = [Cf ] {ε̃f } = [Cf ] {ε} = νf − ν (340)

 1 − ν − 2ν 2 Ef ε 


 f f 


 


 0 


 


 0 


 

 
0

and  1 − ν − 2ν ν 
m m

 2
Em ε 



 1 − ν m − 2ν m



 

 νm − ν 


 2
Em ε 

 1 − νm − 2νm

 


{σ̃m } = [Cm ] {ε̃m } = [Cm ] {ε} = νm − ν (341)
 2
Em ε 


 1 − νm − 2νm 


 


 0 


 


 0 


 
0
That is, the transverse stresses associated to the upper bound case are not zero in
the constituents if νf 6= ν 6= νm . However, for the special case in which νf = νm ,
first, Equation (339) yields ν = νf = νm , second, we recover the stress fields of the
strength of materials approach
   

 Ef ε 
 
 Em ε 


 0   
 0  

 
 
 

   
0 0
{σ̃f } = {σ̃ m } = (342)

 0   
 0 
   


 0  



 0 

   
0 0

and, finally, the upper bound of Eq. (338) reduces to

E ≤ Ef V̂f + Em V̂m (343)

- 90-
2.8 Determination of the overall behavior of a lamina through the
behavior of its constituents. Theory of elasticity approach.

Approximated upper bound Ef


Exact upper bound
Lower bound

Em
0 0.2 0.4 0.6 0.8 1
Vf

Figure 35: Bounds for the Young’s moduli of a composite material.

This last upper bound is coincident with the mechanics of materials Young’s modulus
associated to the longitudinal direction of a unidirectionally reinforced lamina, Eq.
(275), and is commonly used provided that νf ≈ νm . Again, the inequalities Eq.
(338) or Eq. (343) can also be interpreted to be valid with E being any principal
Young’s moduli of a fiber-reinforced composite material.
The lower bound Eq. (320) and the approximated upper bound Eq. (343) of
a conventional unidirectionally fiber-reinforced composite material, uniquely deter-
mined for given Ef and Em , are shown in Figure 35. The exact upper bound given
in Eqs. (338) and (339), using the values νf = 0.25 and νm = 0.35, is also depicted.
Note the approximated and exact upper bounds are almost indistinguishable in this
case. The lower and upper bounds define a corresponding admissible zone in which
the values of the principal Young’s moduli of any (fiber, ribbon or particulate) com-
posite lamina containing two constituents with the considered values Ef and Em
(and also νf and νm if the exact upper bound is to be calculated) must remain.

2.8.2 The Halpin-Tsai Equations


For practical (design) purposes, Halpin and Tsai developed an interpolation proce-
dure which provides an approximated solution for the ply elastic properties expressed
in terms of the constituents elastic properties. Their proposal is motivated by com-
plex elasticity-based solutions (specially Refs. [13] and [14]) which use is restricted
to very few cases and packing geometries of the reinforcing material within the ma-
trix material. In contrast to this type of elasticity models, the Halpin-Tsai equations
adopt a relatively simple form and their applicability is much wider. Moreover, they
still provide rather accurate approximations of the micromechanical behavior of or-
thotropic laminae. Hence, these equations are widely used for design of composite

- 91-
2 Mechanical Analysis of a Lamina

structures.
The set of micromechanical expressions proposed by Halpin and Tsai for the
principal elastic properties of an orthotropic lamina comprises the rule-of-mixtures
for Ex (fiber-dominated) and νxy , i.e.
Ex = Ef V̂f + Em V̂m (344)
νxy = νf V̂f + νm V̂m (345)
along with the following particular expression for other moduli
E 1 + ξη V̂f
= (346)
Em 1 − η V̂f
where E represents the composite modulus Ey , Gxy or Gyz ; Ef is the corresponding
fiber modulus Ef or Gf ; Em is the corresponding matrix modulus Em or Gm ; the
term ξ is an empirical parameter which depends on the loading conditions and
the (fiber, ribbon or particulate) reinforcement geometry and distribution; and η is
defined through
Ef
−1
E
η= m (347)
Ef

Em
Note that, besides the micromechanical elastic and volume fraction properties, an
additional factor ξ is needed to be known for a given composite geometry and bound-
ary conditions so as to the Halpin-Tsai formulas can be applied. From this aspect,
it is remarkable that when ξ = 0, then η = 1 − Em /Ef and the matrix-dominated
rule-of-mixtures 1/E = V̂f /Ef + V̂m /Em for the compliance 1/E is recovered. Recall
that this expression corresponds to the admissible lower bond for the stiffness E. On
the other hand, if we take the limit ξ → ∞, then η → 0 and ξη → Ef /Em − 1, so the
fiber-dominated rule-of-mixtures E = Ef V̂f + Em V̂m (or approximated upper bound)
is obtained for the stiffness E. Values of ξ between 0 and ∞ will give expressions for
E between these bounds (see Figure 36). The limit cases ξ = 0 and ξ → ∞ indicate
that, among other factors, the parameter ξ represents a direct measure of how the
reinforcing material contributes to the corresponding macroscopic stiffness of the
composite material. Reliable values (or even parametric expressions) of the empir-
ical factor ξ to be used in Eqs. (346)–(347) can be estimated by means of making
these formulae conform to exact elasticity solutions or available experimental data.
For further derivations of the Halpin-Tsai formulas, elaborated expressions for ξ
commonly used and some other specific constitutive semi-empirical formulae to be
applied for particulate composites, porous solids, oriented continuous or discontin-
uous fibers or ribbons, the interested reader is referred to the Halpin’s book, Ref.
[15], and the references given therein.

2.9 Determination of the strength properties of a lamina from its


constituents. Strength of materials approach.
In the previous Section we have dealt with the stiffness properties of a lamina from
those of its constituents. In this section we give some insight on the calculation

- 92-
2.9 Determination of the strength properties of a lamina from its
constituents. Strength of materials approach.

ξ = 0 (upper bound)
ξ=5 Ef
ξ = 20
ξ = 80
ξ = ∞ (lower bound)

Ey

Em
0 0.2 0.4 0.6 0.8 1
Vf

Figure 36: Halpin-Tsai calculations for Ey using different values of the empirical
parameter ξ.

of the axial strength of a fiber-reinforced lamina in the reinforcing direction using


simple theories of mechanics of materials. Obviously, some inelastic phenomenons
may occur in both fibers and matrix before the composite material fails, so elasticity
solutions cannot be derived for a general situation. In this case, the continuous ad-
vances of the Solid Mechanics field are necessary in order to obtain more elaborated
micromechanical models for the strength behavior of a composite lamina.

Tensile Strength
For conventional fiber-reinforced composites, the fiber material is stiffer, stronger
and more brittle than the matrix constituent. This consideration adds to the fact
that, for example, boron, carbon and ceramic fibers are essentially elastic up to
fracture and that polymeric and ceramic matrices can be regarded, for practical
purposes, to behave elastically up to failure. Then, the behavioral curves for the
constituents shown in Figure 37 are proposed based on those facts.
When the composite lamina is pulled on tension in the direction of the fibers until
its ultimate load σu = X is reached, assume that all the fibers within the lamina
fracture at the same time. Of course this is quite unrealistic, but this simplifying
hypothesis will be useful to present a simple yet useful formulation. Assume also
the implicit relation νf = νm used previously in the strength of materials approach,
which implies absence of transverse stresses in the constituents of Figure 29. During
the simultaneous elastic range of both fibers and matrix, see Figure 38, the average
tensile stress applied to the lamina is (Eqs. (271), (274) and (275))

σ = σf V̂f + σm V̂m = Ex ε , ε < Xfε (348)

- 93-
2 Mechanical Analysis of a Lamina

σ
Xf Xf Xf

Xm
Xm Xm
(σm )X ε
f
ε
Xfε Xmε

Figure 37: Assumed responses up to fracture for the isolated constituents.

Since a perfect bond is presumed (ε = εf = εm ) and Xfε < Xm


ε holds, the axial strain
ε
first reaches the value ε = Xf , whereupon Eq. (348) reads

σXfε = Xf V̂f + (σm )X ε V̂m = Ex Xfε , ε = Xfε (349)


f

and all the fibers simultaneously fail before the matrix fractures. Subsequently, the
fibers do not bear any load (another hypothesis being introduced) and the matrix
alone remains subjected to the total axial force being applied at that instant. The
average tensile stress σXfε preserves its value, but the stress in the matrix abruptly
changes from (σm )X ε to σm ∗ , which value can be obtained through
f

∗ ∗
σXfε V̂f
σXfε = σm V̂m ⇒ σm = = Xf + (σm )X ε (350)
V̂m V̂m f

Two situations are then possible. On one hand, if the resulting stress σm ∗ is such

that σm∗ > X , the matrix cannot resist the axial load being applied and the failure
m
∗ < X , the matrix is
of the lamina instantaneously occurs. On the other hand, if σm m
capable of bearing more load until the stress σm = σ/V̂m = Xm is reached, instant
at which the matrix (lamina) fails. The special case for which σm∗ = X separates
m
these two possible situations. Thus

  Xm − (σm )X ε
∗ f
σm = Xm ⇒ V̂f = (351)
min Xf + Xm − (σm )X ε
f

and we obtain a minimum value (V̂f )min of the fiber-volume fraction for which the
lamina fails when the fibers fracture. In this respect, for a given composite lamina,
∗ > X and the composite failure takes place at
if V̂f > (V̂f )min then σm m

X ε = Xfε (352)

- 94-
2.9 Determination of the strength properties of a lamina from its
constituents. Strength of materials approach.

with the fiber-dominated tensile strength of the lamina being


 
X = Xf V̂f + (σm )X ε V̂m = (σm )X ε + Xf − (σm )X ε V̂f (353)
f f f

On the contrary, if V̂f < (V̂f )min , then σm∗ < X


m and the failure of the composite
will occur at
X ε = Xmε
(354)
with the matrix-dominated tensile strength of the lamina being
 
X = Xm V̂m = Xm 1 − V̂f (355)

In Figure 38, several cases associated to different fiber-volume fractions are repre-
sented by their corresponding paths in a σ-vs-ε graph and a X -vs-V̂f graph. Note
that, when the strain ε = Xfε is reached and the fibers fail, the instantaneous matrix
∗ is followed by an abrupt strain change from X ε
stress variation from (σm )X ε to σm
f f
to ε∗ = ε∗m , which can be calculated by means of

∗ Ef V̂f + Em V̂m
σXfε = σm V̂m ⇒ Ex Xfε = Em ε∗ V̂m ⇒ ε∗ = Xfε (356)
Em V̂m
This instantaneous increment of the axial strain leads to the definitive fracture of
the lamina whenever V̂f > (V̂f )min .
Obviously, we design the composite material to gain strength from having fiber
reinforcement (when compared to the matrix alone). Then, the requirement X > X m
should be satisfied by the composite lamina, that is
  Xm − (σm )X ε  
f
(σm )X ε + Xf − (σm )X ε V̂f > Xm ⇒ V̂f > = V̂f (357)
f f Xf − (σm )X ε crit
f

where (V̂f )crit is the critical fiber-volume fraction for which the fiber-dominated
strength of the composite equals the strength of the matrix constituent (see Figure
38). Note (V̂f )crit > (V̂f )min . Finally, it can be observed in the X -vs-V̂f graph
that the strength of the lamina expressed in terms of the strength properties of
its constituents is piecewise defined, in clear contrast to the single curves which
define the elastic properties of the lamina in terms of the elastic characteristics of
its constituents.
The mechanics of materials formulation just given provides commonly accepted
behavioral tendencies that should be taken into account just from a qualitative
viewpoint. Too many simplifying hypothesis have been assumed so as to this mi-
cromechanical approach to the strength of a fiber-reinforced lamina can be used
for design purposes. First, we already know that generally νf 6= νm , so the pro-
posed displacement and stress fields are only approximated. Second, we have not
considered the possible existence of inelastic phenomenons in both fibers and ma-
trix (specially in the latter) before the composite material fails. In addition, the
assumption that all the fibers within a composite lamina simultaneously break is
questionable, mainly because fibers do not all have the same strength due to surface

- 95-
2
σ Abrupt change of strain X
X X
Composite lamina failure

Xf Vf = 1 Xf

Vf > (Vf )crit

(Vf )crit

erties for different fiber-volume fractions.


(Vf )min
Mechanical Analysis of a Lamina

- 96-
Vf = 0 Xm
Xm

Vf < (Vf )crit (σm )X ε


f

ε ε (Vf )crit Vf > (Vf )crit Vf = 1


Xf Xmε
(Vf )min

Vf < (Vf )min

Vf = 0

Figure 38: Failure of composite laminae in terms of the constituents strength prop-
2.9 Determination of the strength properties of a lamina from its
constituents. Strength of materials approach.

imperfections. Indeed, fractures usually begin at half the ultimate load and then
they rapidly accumulate as the load increases (Ref. [16]). Finally, it has also been
supposed that when a fiber fails, then it is no longer capable of bearing more load.
This is arguably not entirely true. In Figure 39, a representative volume contain-
ing a broken fiber embedded within a matrix is shown. At the exact location of
the fracture, x ≈ 0, the fiber remains unloaded, so its longitudinal stress becomes
σf (x ≈ 0) = 0. However, for x ≥ δ (δ being a relatively short distance) the matrix
and the fiber are experiencing equal axial displacements and strains (imaginary ver-
tical lines are not distorted), which indicates that both of them are subjected to axial
stresses. That is, for x ≥ δ the longitudinal stress in the fiber is σf (x ≥ δ) = σ0 ,
where σ0 represents the stress level of any other fiber in the composite far from the
break, and the fiber is resisting load partially. If the fiber is isolated in the range
0 < x ≤ δ, as it is also shown in Figure 39, it is apparent that shear stresses must
appear in the fiber surface to equilibrate the fiber axial stresses. The fact that fiber
and matrix are still perfectly bonded for 0 < x ≤ δ (i.e. there is no sliding between
them, whereupon imaginary vertical lines are distorted), means that these shear
stresses are actually being transferred from the matrix to the fiber along its com-
mon interface. The shear stresses in the matrix are a consequence of the shear strain
being imposed in the matrix material to ensure the perfect bonding compatibility
condition. Obviously, for x > δ the shear stresses vanish. The tensile stress in the
fiber and shear stress in the fiber/matrix interface distributions are also shown in
Figure 39. Hence, it should be clear that around any fiber fracture there exists a
shear stress transfer from the matrix to the fiber which results in a fiber longitudinal
stress increment from zero at the break to the value σ0 at a given distance δ. This
stress transfer mechanism provokes that broken fibers are capable of bearing par-
tial load far from their associated breaks and that the energy-absorption capacity
of an actual composite material exceeds that of the mechanics of materials model
presented above. Hence, the expression for the strength X of an actual lamina in
terms of the fiber-volume fraction V̂f usually locates above the curve represented
in Figure 38. Finally, we remark that a possible failure mechanism not included in
the strength of materials model emerges when one considers that high shear stresses
may appear in the surroundings of a fiber fracture. The shear stress level could
exceed the matrix shear strength, leading to a shear failure of the matrix material.

Compression Strength

The determination of the compressive strength of a lamina in the fiber direction in


terms of the compressive strength characteristics of its constituents is quite more
involved than its tensile counterpart. A very simple model can be derived as follows.
The compressive strength of an isolated fiber with length L relatively much longer
than its diameter d (i.e. L/d ≫ 1) corresponds, evidently, to buckling failure.
Moreover, the value of the corresponding allowable stress, associated to the first
mode of a bidimensional Euler column model without lateral support, i.e.
 π 2
(Xc )f = Ef I (358)
L

- 97-
2 Mechanical Analysis of a Lamina

y τ
σ0

σ x σ
σ σ0
τ
δ x
δ
Figure 39: Stress transfer mechanism in the surroundings of a broken fiber embedded
within a binder medium.

will be very low. However, when the transverse displacement constraint introduced
by the surrounding matrix material is included in the analysis of the fiber buckling,
the previous result can be modified by means of increasing the number of lateral
equally-spaced supports on the column. Using this rigid and discrete modelling of
the elastic and continuous matrix lateral support is analogous to increase the mode
number m of the Euler column model, which yields for the composite compressive
strength
 mπ 2
Xc = Ef I = m2 (Xc )f (359)
L
If m ≫ 1, then Xc ≫ (Xc )f as one could expect. As a matter of fact, experimental
observations have shown that the buckle wavelengths of the fibers embedded within
a matrix are directly proportional to the fiber diameter. Hence, a high value of the
mode number m is effectively to be considered in Eq. (359) if L/d ≫ 1. However,
the fact that a smaller value of d implies a lower value of the second moment of area
I in Eq. (359) has also to be taken into account.
Finally, note that this is an empirical model based on very simple buckling the-
ories in which the strength characteristics of the matrix are not included anywhere.
Thus, it is far away from being regarded a definitive micromechanics model of the
compressive strength of a composite lamina. More elaborated formulations, also
based on the overall hypothesis that fiber buckling is responsible for compressive
failure, and with an argued consideration of the elastic properties of both fibers
and matrix can be found in Ref. [4]. Anyway, modifications based on empirical
observations seem also to be necessary to make those models conform to the avail-
able experimental data. Certainly, micromechanical formulations of the compressive
strength of a composite material should also reproduce the macroscopic effects of
the structural buckling associated to the lamina (plate) itself. In this respect, much
work remains to be done.

- 98-
3 Constitutive Equations for Laminated Composites
The two main features of filamentary composites that make them very attractive
for structural applications are, on one hand, the high specific mechanical properties
they have and, on the other hand, the ability they present to be tailored to the me-
chanical design requirements. How the stiffness-to-density and strength-to-density
characteristics of composites compare to those of conventional materials was out-
lined in Chapter 1. The controlled anisotropy, or how the property values in different
directions can be varied in a composite laminate in order to be capable of sustaining
external loads efficiently, is investigated in the present chapter.
Controlled anisotropy of composites is accomplished by the ability of laminae to
be oriented and laid up on specific, desired, directions to conform a laminate (recall
Figure 2). Some simplistically, consider the loading cases depicted in Figure 40.
First, for a uniaxial loading situation, it is evident that a unidirectionally reinforced
composite may be sufficient to bear the axial load with the desired longitudinal
stiffness and strength characteristics. Then, because of the equivalence between a
pure shear state of stresses and a biaxial state of normal stresses rotated 45o degrees
relative to each other, we can efficiently resist a pure shear stress state defined by
σxy by means of a set of unidirectional layers being arranged in directions forming
±45o with respect to the x-axis. Finally, if a specific design includes the application
of a generic system of normal and shear loads (defined with respect to the axes Oxy),
the laminate would require composite plies being laid up at the orientations 0o and
90o , to bear the normal stresses, and at the orientations ±45o , to resist the shear
stresses. Obviously, the loading cases of Figure 40 are just simple examples in which
the direction of the fibers within the laminate can be easily inferred. The design of
other structures subjected to more complex loading situations can be much more
involved. Moreover, note the specific thickness and number of layers to be laid up
at each direction, along with the stacking sequence in which they are arranged, are
also to be defined by the designer. In what follows, we will able to see how these
factors definitely define both the in-plane and bending macromechanical behavior
of the corresponding laminate being conformed.
One important fact to be taken into account when working with laminates is that
the laminae are perfectly bonded all together. In order to see the implications of a
perfect bonding condition between adjacent laminae within a laminate, consider the
well-known two-beam problem of Figure 41. From elementary Strength of Materials,
we know that the deflection of a beam with rectangular cross-section (b × h) loaded
at midspan by a force P is
P L3 P L3 P L3
δbeam = = = (360)
48EI bh3 4Ebh3
48E
12
If we consider now two unbonded beams bearing the load P , then the bending
stiffness becomes twice the bending stiffness of the single beam, yielding
P L3 P L3 δbeam
δunbonded = = = (361)
bh3 8Ebh 3 2
2 × 48E
12

- 99-
3 Constitutive Equations for Laminated Composites

o
0
σxx

σxx

σxy ±45
o

σxy

o o o
0 , ±45 , 90

σxx σxy σyy

Figure 40: Mechanical tailoring to simple loading cases through ply orientation
within the laminate.

- 100-
3.1 Review of the theory of isotropic plates

P P
P

Beam Unbonded beams Bonded beams

Figure 41: Deflection of equivalent unbonded and bonded beams.

However, if the two beams are perfectly bonded, then they work together as a unit
beam with cross-sectional area b × 2h. The deflection is
P L3 P L3 δbeam δunbonded
δbonded = 3 = 3
= = (362)
b (2h) 32Ebh 8 4
48E
12
and the bending resistance of the bonded system results to be four times the bending
resistance of the unbonded pair of beams. This last result directly explains both the
importance of a perfect bonding between adjacent laminae and the reason for which
inspection is extremely necessary in current manufacturing practice to ensure good
(nearly perfect) bonding in every manufactured laminate.

3.1 Review of the theory of isotropic plates


A plate is a plane solid with a thickness h much smaller than any characteristic
length L defining the main plane of the plate. In Structural Mechanics, the Theory
of Plates studies the response of plates under generic loads, see Figure 42. These
loads are commonly divided into loads being perpendicular to the middle surface
of the plate (where the axes Oxy are embedded) and loads being parallel to this
main plane of the plate. In isotropic plates, loads applied in direction Oz will be
equilibrated with out-of-plane shear forces and bending and twist moments, while
in-plane loads will generate in-plane shear and axial forces. Moreover, in-plane
and out-of-plane (bending) behavior are uncoupled in isotropic plates under small
strains and deflections. Hence both effects can be dealt with separately. However,
in composite materials it is frequent the case when both effects are coupled, as we
will see below. When a plate deforms within its main plane only, then we refer to it
as a slab.
Similarly to what happens in the Theory of Beams, where the Navier-Bernoulli
hypotheses are usually adopted in order to describe the pure bending of beams,
within the Theory of Plates it is usual to accept the so-called Kirchhoff hypothe-
ses, which simply generalizes the assumptions taken for unidimensional beams to
bidimensional plates. The first hypothesis simply characterizes the through-the-
thickness displacement uz of a given point of the plate, with coordinates (x, y, z),
in terms of the displacement u0z of the corresponding middle-surface point, with
coordinates (x, y, 0), namely
uz (x, y, z) = u0z (x, y) (363)
From now on, the superscript (0 ) is used to represent values at the middle plane.
The second assumption states that straight lines normal to the undeformed middle

- 101-
3 Constitutive Equations for Laminated Composites

fz (x, y) z

ext
F2
h

x ext y
F1 h ≪ Lx ∼ Ly

Figure 42: Plate subjected to external loads.

surface of the plate remain straight and normal to the deformed middle surface after
bending. The implications of this last hypothesis can be obtained from Figure 43.
Focusing on the displacements that take place within the plane Oxz, we observe
that the rotation angle about the Oy axis of the straight line under consideration is

ux (x, y, z) − u0x (x, y)


θy (x, y) ≃ tan θy (x, y) = (364)
z
—note u0x (x, y) and u0y (x, y) are zero in that figure for gaining simplicity—, and by
virtue of the second Kirchhoff hypothesis it is also equal to

u0z (x + dx, y) − u0z (x, y) ∂u0 (x, y)


θy (x, y) ≃ tan θy (x, y) = − =− z (365)
dx ∂x
Combining the last two equations, we obtain the expression of the displacement
ux (x, y, z)
∂u0 (x, y)
ux (x, y, z) = u0x (x, y) − z z (366)
∂x
Analogous operations within the Oyz plane lead to

∂u0z (x, y)
uy (x, y, z) = u0y (x, y) − z (367)
∂y

Therefore, the consideration of both Kirchhoff hypothesis gives the displacement


field u (x, y, z) defined by Eqs. (363), (366) and (367).
For such a displacement field, the resulting in-plane strain field (we omit the
explicit dependences) is obtained by means of Eqs. (28)1 , (28)2 and (29)1

∂u0x ∂ 2 u0z ∂u0y ∂ 2 u0z


εxx = −z εyy = −z (368)
∂x ∂x2 ∂y ∂y 2
!
1 ∂u0x ∂u0y ∂ 2 u0z
εxy = + −z (369)
2 ∂y ∂x ∂x∂y

- 102-
3.1 Review of the theory of isotropic plates

z z
ux(x,y,z)>0 uy(x,y,z)<0

uz0(x,y+dy)>0
θy(x,y)>0 x θx(x,y)>0
π/2 uz0(x+dx,y)<0 π/2 y

Figure 43: Displacement field for a Kirchhoff plate. In-plane displacements of the
middle surface are not depicted.

In the preceding expressions we identify the strains associated to the middle surface
!
0 ∂u 0 0 ∂u 0
∂ux y 1 ∂u x y
ε0xx = ε0yy = ε0xy = + (370)
∂x ∂y 2 ∂y ∂x

and also its associated bending and twist curvatures

∂ 2 u0z ∂ 2 u0z ∂ 2 u0z


κ0xx = − κ0yy = − κ0xy = − (371)
∂x2 ∂y 2 ∂x∂y

Then, the in-plane strains can be rewritten in matrix notation as


   0   0 
εxx  εxx  κ 
 xx 

  
  

    
εyy = ε0yy +z κ0yy (372)

 
    
    0    0 
 
γxy γxy κxy

or, using the index numeration


   0   0 
ε ε κ 
 1
    
 
  1 
   1 
 
ε2 = ε02 +z κ02 (373)

 
    
    0    0 
 
ε4 ε4 κ4

which results into a through-the-thickness linear variation for the strains of points
P (x, y, z) in terms of the strains and curvatures of the corresponding middle-surface
point P 0 (x, y). Aside, note we have defined (for convenience) the total twist curva-
0 as twice the twist curvature κ0 (κ 0 = 2κ0 = κ0 ) in the same way as the
ture κxy xy xy xy 4
total angular distortion γxy is defined as twice the angular strain εxy (γxy = 2εxy =
ε4 ).
We assume hereafter that the plate is under a plane stress condition. Hence,
once the expressions of the in-plane strains are known, the direct application of Eq.

- 103-
3 Constitutive Equations for Laminated Composites

(170) specialized for an isotropic material (Young’s modulus E and Poisson’s ratio
ν) gives the following expression for the in-plane stress field
 0   0 

σ

 εxx + νε0yy   κxx + νκ0yy 

 xx 
 
 
 
 

  E   E  
σyy = ε0yy + νε0xx +z κ0yy + νκ0xx (374)
  1 − ν 2   1 − ν 2  
 1 − ν γ0   1 − ν κ0

 
 
  
 

σxy xy

xy

2 2
which is, again, linear through the thickness and where the middle-surface (z = 0)
stresses are  0 
 0 
σ  εxx + νε0yy 

 xx 
 
 

  E  0 0

0
σyy = εyy + νε xx (375)
  1 − ν2  
 0 
  
 1 − ν 0


σxy  γxy 
2
It is important to emphasize that the first Kirchhoff hypothesis, Eq. (363), gives
an strain field under plane strain (εzz = ∂uz /∂z = 0). On the other hand, we have
later used Eq. (170), which was obtained for laminae under plane stress (σzz = 0).
We want to remark that if the through-the-thickness strain is prescribed to be zero,
then σzz cannot vanish, in general (and vice versa). This apparent contradiction
(frequently found in the literature) may be explained by the assumption that the
plane strain condition introduced above is only approximated (i.e. εzz ≈ 0). Then,
we use the (also approximated) displacement field derived from the Kirchhoff hy-
potheses along with the plane stress condition, which governs more realistically the
behavior of thin plates subjected to in-plane loads and bending moments.
We are interested in obtaining the resultant axial and shear forces (Nx , Ny and
Nxy ) and bending and twist moments (Mx , My and Mxy ) per unit width acting on
the plate, as shown in Figure 44. They can be obtained integrating the preceding
stresses through the thickness coordinate z as
Z h/2 Z h/2 Z h/2
Nx = σxx dz Ny = σyy dz Nxy = σxy dz (376)
−h/2 −h/2 −h/2

Z h/2 Z h/2 Z h/2


Mx = zσxx dz My = zσyy dz Mxy = zσxy dz (377)
−h/2 −h/2 −h/2

Using Eq. (374), the forces and moments can be expressed in terms of the in-plane
strains and curvatures of the middle surface as
  1−ν 0
Nx = A ε0xx + νε0yy Ny = A ε0yy + νε0xx Nxy = A γxy (378)
2
 1−ν 0

Mx = D κ0xx + νκ0yy My = D κ0yy + νκ0xx Mxy = D
κxy (379)
2
where we have used the so-called extensional A and bending D stiffnesses of the
plate
Eh Eh3
A= D = (380)
1 − ν2 12(1 − ν 2 )

- 104-
3.1 Review of the theory of isotropic plates

z z

dx dy dx dy
σxy
x y x y
Nx Ny Mxy
σxx σyy Mxy
My Nxy Nxy
Mx

Figure 44: Resultant forces and moments per unit width at a point of the plate.
Positive senses are drawn.

Note that the forces given in Eq. (378) depend only on the in-plane middle-surface
strains and that the moments of Eq. (379) depend only on the middle-surface cur-
vatures. Hence, as previously commented, the extension-shear behavior and the
bend-twist behavior are completely uncoupled for an isotropic plate. Moreover, we
observe that the extensional behavior is also uncoupled from the shear behavior
and that bending is uncoupled from twist as well. As we will see below, some of
these kinds of coupling may naturally appear for a general anisotropic plate. The
anisotropy of the plate can be generated by either the use of a single anisotropic ma-
terial conforming the plate or the use of specific ply layups conforming a laminated
plate.
Finally, if we rewrite the component σxx , for example, of Eq. (374) using Eqs.
(378)1 and (379)1 we arrive to the well-known Strength of Materials expression for
beams
Nx Mx
σxx = + 3 z (381)
h h /12
where, clearly, h plays the role of the area per unit width of the transverse section
which normal axis is Ox (Ax = h × 1) and h3 /12 is the second moment of area (or
area moment of inertia) per unit width of that section with respect to the axis Oy
(Iy = (h3 × 1)/12).

Exercise 10 Use MATLAB to create a function file (plotplatedef.m)  0which


plots
the deformed
 0 configuration of a plate for given middle-surface strains ε and cur-
vatures κ .
Solution: The code for MATLAB is

function hsurf = plotplatedef(ek,nfig)


%
% function hsurf = plotplatedef(ek,nfig)
%
% function to plot a deformed plate using middle-surface strains
% and curvatures (assumed constant; i.e. local deformation)
% we assume no rotation and no displacement at the origin (x=y=0)
%
% Input:

- 105-
3 Constitutive Equations for Laminated Composites

% ek = vector of strains (ek(1:3)) and curvatures (ek(4:6))


% nfig = figure where to plot the plate
%
% Output:
% hsurf= handle to the surface
%

figure(nfig); hold on;


L = 1; % length of the plate
N = 2; % number of divisions of the undef. plate
x = -L/2:L/N:L/2; % points X-coord
[X,Y] = meshgrid(x,x); % points of the grid (lines)
z = 0*X; % original plate position
hsurf = surfl(X,Y,z); % plots undeformed plate with some transparency
set(hsurf,’FaceColor’,[0.9 0.9 0.9],’FaceAlpha’,0.6);
%
% compute x,y deformation scale factor
%
magfac = 0.25;
xscale = 1/max(abs(ek(1:3)))*magfac;
%
% compute z-coordinates
%
N = 10; % number of divisions of the def. plate (even)
x = -L/2:L/N:L/2; % points X-coord
[X,Y] = meshgrid(x,x); % points of the grid (lines)
a = -ek(4); % x-curvature: note sign convention
b = -ek(5); % y-curvature: note sign convention
c = -ek(6)/2; % xy-curvature
z = a/2 * X.^2 + b/2 * Y.^2 + c * X.*Y; % vertical displacements
%
% compute z-scaling factor and definite factor
%
zscale = L/max(max(abs(z)))*magfac;
scale = min(xscale,zscale);
ek(1:3) = ek(1:3) * scale;
z = z * scale;
%
% compute x,y coordinates
%
x = X + ek(1) * X + ek(3)/2 * Y; % horizontal displacements
y = Y + ek(2) * Y + ek(3)/2 * X; % id
%
% actual plot
%
hsurf = surfl(x,y,z); % plots deformed plate, no transparency

- 106-
3.1 Review of the theory of isotropic plates

set(hsurf,’FaceColor’,[0.6 0.6 0.6],’FaceAlpha’,1.);


view(3); % standard 3D view
axis equal;
title([’Plate deformation; magnified by ’,num2str(scale)]);
%
return

 0
For example, the resulting deformed plate for ε = {0, 0, 0}T and {k0 } =
{2, −1, 0}T × 10−3

ek = [0,0,0,2,-1,0]*1e-3;

is shown in the following Figure (note the high value assigned to the scaling
factor)

Figure of Exercise 10

- 107-
3 Constitutive Equations for Laminated Composites

z N
zN
zN-1 k
zk
h zk-1
x, y
z1
z0 1
Figure 45: Geometry of a generic N-layered laminate.

3.2 Laminated Plate Theory

The procedure to obtain the extensional and bending stiffnesses of a laminate, which
relate internal forces and moments to the laminate middle-surface strains and curva-
tures, is essentially analogous to the one that has been followed for obtaining these
structural properties for an isotropic plate. We depart from the generic laminate
represented in Figure 45, which total thickness is h. The plane Oxy is embedded in
the laminate middle surface, there are N ≥ 1 (possibly different) laminae and the
k-th lamina is located between the coordinates zk−1 and zk . Obviously, the case
N = 1 corresponds to the case of a single lamina, with z0 = −h/2 and z1 = h/2.
For N > 1, each lamina has its specific thickness, mechanical properties and orien-
tation of the preferred axes with respect to the laminate axes X = Oxyz. We are in
the position to describe the mechanical response of each one of these laminae when
they act on its own using the theories explained in the previous Chapter. We are
interested herein in obtaining the global structural behavior of the laminae when
they are joined all together and work as a unit structure, i.e. as a laminate. In
what follows, we derive the present formulation for flat laminates. However, it is
important to note that the laminates can, evidently, be curved or shell-like.
As we have done for the common analysis of isotropic plates, we approximate the
displacement field throughout the laminate by means of the Kirchhoff hypotheses
stated above. Because the (simplified) kinematic description is identical to the
case of isotropic plates, the in-plane strain field results into the linear dependence
given in Eq. (372), which, note, is continuous through the thickness. However,
since the mechanical properties and orientation are different for each layer, we must
apply Eq. (170) separately to each lamina. Following the arguments given in the
preceding chapter, each unidirectionally reinforced lamina is assumed to behave
orthotropically, although only the in-plane elastic properties will be needed. In
Figure 46, the particular placement of the lamina k, which preferred axes are Xk =
Oxk yk zk , with respect to the working basis X is shown. According to Eq. (171), the
in-plane stresses for the lamina k in terms of its reduced stiffnesses and the in-plane

- 108-
3.2 Laminated Plate Theory

strains, all of them being expressed in its associated principal basis Xk , are
 (k)  (k)  (k)
 σ1k  Q11k Q12k 0  ε1k 
σ2k =  Q12k Q22k 0  ε2 (382)
   k 
σ4k 0 0 Q44k ε4k
or
(k) (k) (k)
{σ}Xk = [Q]Xk {ε}Xk (383)
where the superscript (k) indicates that we are referring to the k-th lamina of the
laminate and the subscript Xk (or simply k) indicates the system of representation
being used. Again, note Q11k stands for Q(11)k ≡ Q1k 1k , and so on. In order to
account for the different stress contribution of each laminae to the resultant laminate
forces and moments, we have to refer all measures to the same reference basis, i.e.
the system of reference of the laminate X = Oxyz. We change the representation
basis of Eq. (382) from Xk to X using an analogous transformation law to Eq. (100)
to arrive to
(k) (k) (k)
{σ}X = [Q]X {ε}X (384)
where the (apparent) reduced stiffness matrix relative to the k-th lamina becomes
 (k)
Q11 Q12 Q14
(k)
[Q]X =  Q12 Q22 Q24  (385)
Q14 Q24 Q44

with

Q11 = Q11k cos4 α + 2 (Q12k + 2Q44k ) sin2 α cos2 α + Q22k sin4 α (386)
Q22 = Q11k sin4 α + 2 (Q12k + 2Q44k ) sin2 α cos2 α + Q22k cos4 α (387)
2 2
 4 4
Q44 = (Q11k + Q22k − 2Q12k − 2Q44k ) sin α cos α + Q44k sin α + cos α (388)

Q12 = (Q11k + Q22k − 4Q44k ) sin2 α cos2 α + Q12k sin4 α + cos4 α (389)
Q14 = (Q11k − Q12k − 2Q44k ) sin α cos3 α − (Q22k − Q12k − 2Q44k ) sin3 α cos α
(390)
Q24 = (Q11k − Q12k − 2Q44k ) sin3 α cos α − (Q22k − Q12k − 2Q44k ) sin α cos3 α
(391)

where the angle α = α(k) is defined in Figure 46. Introducing the strain linear
variation law given in Eq. (372)
(k)  
{ε}X = ε0 X + z κ0 X z ∈ [zk−1 , zk ] (392)

into Eq. (384) yields the following expression for the in-plane stresses of the k-th
lamina
(k) (k)  0  
{σ}X = [Q]X ε X + z κ0 X z ∈ [zk−1 , zk ] (393)
(k) (k+1)
Since the matrices associated to contiguous plies, [Q]X and [Q]X , will be different
in a general case, then the in-plane stresses will be continuous within each lamina

- 109-
3 Constitutive Equations for Laminated Composites

yk y
xk
(k)
α

Figure 46: Orientation of the principal material axes of the k-th lamina with respect
to laminate axes X. The angle α(k) depicted is regarded positive.

(k)
(indeed, linear with z if [Q]X is uniform at the lamina level) but discontinuous at
boundaries between laminae (compare with the isotropic case presented above). An
example of a typical linear strain through-the-thickness profile and its associated
piecewise linear stress distribution at a given point P 0 (x, y) of a laminate can be
seen below, in the figure obtained as the Matlab output of Exercise 14. In the
following lines we proceed to explain the needed tools to be able to compute these
distributions when the resultant forces and moments acting on a point of a laminate
are known.
From now on, all the variables will be represented onto the laminate basis X
and the subscript indicating the system of representation will be omitted when no
confusion is possible. The integration of the stresses through the laminate thickness
will provide the desired relation between the internal forces and moments and the
laminate middle-surface strains and curvatures
 
 Nx  Z h/2
{N } = N = {σ} dz (394)
 y  −h/2
Nxy
 
 Mx  Z h/2
{M } = My = z {σ} dz (395)
  −h/2
Mxy
Taking into consideration that the stress fields is piecewise linear through the thick-
ness, then the axial and shear forces of Eq. (394) are
N Z zk !
X
{N } = {σ}(k) dz (396)
k=1 zk−1

Using Eq. (393) and noticing that ε0 and κ0 are middle-surface measures and do
not depend on z, we obtain the following expression for the axial and shear forces
 
{N } = [A] ε0 + [B] κ0 (397)

- 110-
3.2 Laminated Plate Theory

where   !
A11 A12 A14 XN Z zk
(k)
[A] =  A12 A22 A24  = [Q] dz (398)
A14 A24 A44 k=1 z k−1

and   !
B11 B12 B14 XN Z zk
[B] =  B12 B22 B24  = z [Q](k) dz (399)
B14 B24 B44 k=1 zk−1

are the so-called extensional stiffness matrix and the bending-extension coupling
stiffness matrix (associated to the reference frame X), respectively. Note that they
are symmetric by virtue of the symmetry of every contribution [Q](k) . If, as usual,
the reduced stiffnesses are constant for each layer, the preceding matrices simply
reduce to
N
X
[A] = [Q](k) (zk − zk−1 ) (400)
k=1
N
1 X 
[B] = [Q](k) zk2 − zk−1
2
(401)
2
k=1

In case the reduced stiffnesses were non-uniform for the given laminate constituents
(an example would consist in laminae with temperature-dependent elastic properties
subjected to a through-the-thickness temperature gradient), then Eqs. (398) and
(399) are to be used. The integration of Eq. (395), using Eq. (393), gives the
following expression for the bending and twist moments
 
{M } = [B] ε0 + [D] κ0 (402)

where the symmetric matrix [D], known as the bending stiffness matrix (associated
to the reference frame X), is obtained through
  !
D11 D12 D14 XN Z zk
(k)
[D] =  D12 D22 D24  = z 2 [Q] dz (403)
D14 D24 D44 k=1 z k−1

For a laminate with uniform reduced stiffnesses [Q](k) for every layer, [D] reduces to
N
1X 
[D] = [Q](k) zk3 − zk−1
3
(404)
3
k=1

It is remarkable that the bending-extension coupling stiffness matrix [B] appears


in both the force-strain-curvature relation of Eq. (397) and the moment-strain-
curvature relation of Eq. (402); in the former being multiplied by curvatures to
give forces (per unit width) and in the latter being multiplied by strains to give
moments (per unit width). Rewriting Eqs. (400), (401) and (404) using the relations
zk = z (k) + t(k) /2 and zk−1 = z (k) − t(k) /2, where
zk + zk−1
z (k) = (405)
2

- 111-
3 Constitutive Equations for Laminated Composites

and
t(k) = zk − zk−1

represent the z-coordinate of the middle surface of the k-th lamina and its thickness,
we obtain
N
X
[A] = [Q](k) t(k) (406)
k=1
XN
[B] = [Q](k) z (k) t(k) (407)
k=1
"  #
N
X  2 t (k) 3
[D] = [Q](k) z (k) t(k) + (408)
12
k=1

These last expressions indicate that only the thickness t(k) of each composite ply con-
tributes to the laminate extensional stiffness matrix [A] (that is, the layout of the
laminae along the Oz axis, represented by z (k) , is not relevant for these stiffnesses),
while both the thickness and the location of each lamina are relevant for the calcula-
tion of the bending-extension coupling stiffness matrix [B] and the bending stiffness
2 3
matrix [D]. Basically, in these last equations, t(k) , z (k) t(k) and z (k) t(k) + t(k) /12
represent the area, the first moment of area respect to the laminate middle surface
and the second moment of area (calculated using the Huygens–Steiner theorem) re-
spect to the laminate middle surface, respectively, of a unit-width transverse section
of the k-th lamina.
Hence, the resultant laminate forces {N } and moments {M } are obtained in
terms of the matrices [A], [B] and [D] and the laminate middle-surface strains {ε0 }
and curvatures {κ0 } by means of the compact expression
 
  ..   0 
{N }  [A] . [B]  ε
 ···  =  ··· · · ·  ···  (409)
   0
{M } .. κ
[B] . [D]

which constitutes the final equations to be used for a generic composite laminate,
as the one shown in Figure 45. It is remarkable that the coefficient matrix in Eq.
(409) contains 18 different stiffnesses (6 per each sub-matrix [A], [B] and [D]) to be
determined from the integration process detailed just above, that is from Eqs. (400),
(401) and (404) (or other equivalent). Therefore, each orthotropic lamina contributes
(k) (k) (k)
to the laminate stiffnesses through its 4 reduced elastic stiffnesses Q11 , Q22 , Q12
(k)
and Q44 (which are obtainable in terms of its four in-plane preferred elastic constants
(k) (k) (k) (k) (k)
Ex , Ey , Gxy and νyx or νxy , as shown in Eq. (170)), its in-plane orientation
α(k) , its thickness t(k) and its placement z (k) respect to the laminate middle surface.
For this reason, we refer to the coefficients Aij , Bij and Dij as structural stiffnesses,
since both elastic properties of the material(s) and geometrical (layup) properties of
the fiber-reinforced composite structure are involved in their definitions.

- 112-
3.2 Laminated Plate Theory

Extension Shear Bend Twist

Extension
A11 A12 A14 B11 B12 B14
A12 A22 A24 B12 B22 B24
Shear A14 A24 A44 B14 B24 B44
Bend
B11 B12 B14 D11 D12 D14
B12 B22 B24 D12 D22 D24
Twist B14 B24 B44 D14 D24 D44

Figure 47: Possible couplings in the coefficient matrix for the force-moment-strain-
curvature relations. Note the extensional and bending behaviors can additionally
be decomposed into their longitudinal and transverse counterparts. Symmetric co-
efficients are depicted in grey color.

3.2.1 Coupling coefficients


We have called the matrices [A], [B] and [D] as the extensional stiffness matrix, the
bending-extension coupling stiffness matrix and the bending stiffness matrix of the
laminate, as it is usually done in the literature. However, shear and twist behaviors
are also included in the definition of those stiffnesses. In order to properly see all
the possible mechanical responses that can be investigated using the formulation
presented just above (i.e. the Laminated Plate Theory), we can expand Eq. (409)
as     0 

 N x 
 A 11 A 12 A 14 B 11 B 12 B 14 
 ε1 



 Ny 



 A12 A22 A24 B12 B22 B24    

 ε02 


    0  
Nxy A 14 A 24 A 44 B 14 B 24 B 
44  ε 4
=   κ0  (410)
 Mx 
    B11 B12 B14 D11 D12 D14   1 
   

 My    B12 B22 B24 D12 D22 D24    κ0 

 
  20 
 
Mxy B14 B24 B44 D14 D24 D44 κ4
where, recall, due to the symmetry of the sub-matrices [A], [B] and [D], only 18
structural stiffnesses define the coefficient matrix in the most general case. Note also
that all the variables are referred to the laminate system of reference X. Similarly as
we did when we analyzed the behaviors and couplings being present in Eq. (133), we
readily observe in Eq. (410) that the 6 stiffnesses of the matrix diagonal are directly
related to the pure extensional, shear, bend and twist responses of the laminate.
Each one of the remaining 12 stiffnesses is associated to a specific kind of coupling
between these four pure mechanical responses. Figure 47 shows the corresponding
label for each structural stiffness present in Eq. (410).
Some of these mechanical couplings are well known. For example, for an isotropic
plate, Eqs. (378) and (379) include the rather intuitive Poisson’s effects for both
extension-extension (A12 = Aν) and bending-bending (D12 = Dν) coupled behav-
iors. However, for some specific composite layered laminates, some non-intuitive
responses may be obtained when other coupling coefficients present in Eq. (410) do
not vanish. For example, in Figure 48 the effect of a non-vanishing extension-twist

- 113-
3 Constitutive Equations for Laminated Composites

Nx

Nx εxx κxy κxy εyy

Figure 48: Illustration of the (possible) extension-twist coupling behavior for a lam-
inate subjected to uniaxial loading.

stiffness coupling B14 can be observed for a particular laminate under tension load-
ing. In absence of other couplings (except for the associated to A12 ), the axial force
causes the usual axial extension and transverse contraction along with a, maybe
unexpected, twisting of the laminate. Finally, note the effect of the bend-twist stiff-
ness coupling D14 (or D24 ) is crucial for accomplishing the aeroelastic tailoring of a
swept-forward wing (Figure 6) in order to prevent undesired aeroelastic phenomena
and delay the stall of this type of wings, as it was explained in Chapter 1. Clearly,
the Laminated Plate Theory presented herein is able to reproduce all these somehow
unexpected, but physically possible, responses.

3.2.2 Standard Laminate Code


We introduce a standard laminate code in order to establish a stacking-sequence
terminology needed to be able to describe a specific laminate uniquely. Assume
first that all the layers conforming the laminate are made of the same fiber-matrix
system material and have the same thickness (typical case in industry). Following
these guidelines, a number is assigned to each lamina within the laminate. This
number represents the angle (in degrees) formed between the reinforced direction of
the lamina and the x-axis of the laminate, taking values comprised between −90o
and 90o . If adjacent laminae have different orientations relative to one another, then
their representative numbers are separated in the code by a slash. On the contrary,
adjacent laminae with the same representative angle are denoted by a numerical
subscript, which indicates the number of adjacent layers for which the orientation is
preserved. We list the laminae in sequence starting with the lamina on the bottom
face of the laminate and ending with the lamina on the other face7 , that is, following
the order in Figure 45. This code is enclosed with brackets (alternatively, parentheses
may also be used). Finally, a subscript “T” is used to emphasize that the total
laminate is shown. This last subscript is somehow redundant and is commonly
omitted. An example of this stacking sequence terminology is shown in Figure 49.a.
Note that a positive angle indicates that the fibers of the corresponding lamina are
counterclockwise oriented with respect to the axis Ox (following the right-hand rule,
positive rotations are in the sense of the axis Oz). Note also the compact notation
[±α] is equivalent to [α/ − α] and [∓α] is equivalent to [−α/α].±
This general notation can be further simplified in some cases. If sequences of
7
Other authors invert this order in the sequence because they define the z-axis positive downward.

- 114-
3.2 Laminated Plate Theory

laminae are repeated, they are called sets and are enclosed in brackets. A numerical
subscript is appended to the brackets in order to indicate the number of repeating
sets that conform the laminate. Aside, if the laminae layup is symmetric relative
to the laminate middle surface, only half the laminae are included in the notation
code and a subscript “S” is used. For a symmetric laminate with an odd number of
laminae, the central lamina is equally split by the laminate middle surface. Then a
bar decoration is put over the angle representing the central lamina. Examples of
these special cases are also represented in Figures 49.b.c.d.
Laminates containing laminae with different thickness are called irregular lami-
nates (laminates with equal-thickness layers are regular) and are simply represented
adding an alpha-numerical subscript (t, 2t, etc.) to each lamina indicating its cor-
responding thickness. Finally, a hybrid laminate is formed by laminae with (a min-
imum of two) different fiber-matrix material systems. In this case, a subscript is
needed to denote the fiber constituent of each layer (the matrix materials are usu-
ally equal to achieve a good curing process). Examples of these two last cases are
shown in Figures 49.e.f .

Exercises using MATLAB


Exercise 11 Use MATLAB to create a function file (plotlaminate.m) which plots
the unbonded view of a laminate for a given stacking sequence in full format.
Solution: The code for MATLAB is

function hlaminas = plotlaminate(laminate,bw,nfig)


%
% function hlaminas = plotlaminate(laminate,bw,nfig)
%
% This function plots a laminate
%
% Input:
% laminate = laminate stacking in full format from bottom to top
% bw = if one, then no color, just black and white (optional)
% nfig = figure number where to plot (optional, default = 40)
%
% Output:
% hlaminas = vector with handles to groups of laminas
%

B = 3; % length of the laminate


H = 2; % width of the laminate
sep = H/5; % separation between fibres in a lamina
nlaminas = length(laminate); % number of laminas
Zstack = B/4*sqrt(nlaminas); % maximum semi-height of the stack plot
dz = Zstack * 2 / (nlaminas-1); % separation between laminae
z = -Zstack-dz; % bottom z-coordinate of the plot
%

- 115-
3 Constitutive Equations for Laminated Composites

a) [-30,0,30,90,±30] b) [(0)2 ,(90)2 ]

x y x y

c) [0,30,30,0] ≡ [0,30]S d) [0,30,0] ≡ [0,30]S

x y x y

e) [02t ,30t ]S f) [0Gr ,30Kv ]S

x y x y

Figure 49: Examples of laminate stacking sequences.

- 116-
3.2 Laminated Plate Theory

if (nargin < 2), bw = 0; end, % black-and-white plot


if (nargin < 3), nfig = 40; end, % default number of figure
%
% loop on laminas
%
for i = 1: nlaminas,
%
% color code (change to fit your taste; see plotlamina)
%
color = ’w’; % default color
if (bw ~= 1),
if (abs(laminate(i)) == 0), color = ’c’; end,
if (abs(laminate(i)) == 90), color = ’c’; end,
if (abs(laminate(i)) == 30), color = ’m’; end,
if (abs(laminate(i)) == 45), color = ’y’; end,
if (abs(laminate(i)) == 60), color = ’m’; end,
end,
%
% plot lamina
%
z = z + dz; % increase z-coordinate for next lamina
hlaminas(i) = plotlamina(B,H,z,laminate(i),sep,nfig,color);
%
end,
%
return

The function file (plotlamina.m) used in plotlaminate.m to plot each lamina is

function hgroup = plotlamina(lx,ly,z,angle,sep,nfig,mcolor)


%
% function hgroup = plotlamina(lx,ly,z,angle,sep,nfig,mcolor)
%
% This function plots a lamina in MATLAB
%
% Input:
% lx = length in x of the lamina
% ly = width in y of the lamina
% z = position of the middle surface of the lamina
% angle = angle of the fibers with x direction
% sep = transversal separation between fibers
% nfig = number of figure where to plot
% mcolor = matrix color, one of ’c’,’m’,’y’,’w’
%
% Output:

- 117-
3 Constitutive Equations for Laminated Composites

% hgroup = handle to the group of plotted objects


%

% map of matrix-fiber colors


fcolor = ’k’; % default fiber color
if (mcolor == ’c’), fcolor = ’b’; end,
if (mcolor == ’m’), fcolor = ’r’; end,
if (mcolor == ’y’), fcolor = ’g’; end,
if (mcolor == ’w’), fcolor = ’k’; end,
%
% figure and group
%
figure(nfig); hold on; view(3); axis off;
hgroup = hggroup; % this is the handle of the group
%
% draws matrix
%
H = ly; B = lx;
X = [0,B,B,0]; Y = [0,0,H,H]; Z = [z,z,z,z];
hpatch = patch(X,Y,Z,mcolor);
set(hpatch,’parent’,hgroup);
%
% prepares data of fibres
%
ang = angle;
if (angle == 90), ang = 89.9999; end, % avoids division by zero
a = tand(ang); % slope
l = max(B,H); n = fix(2*l/sep+1); % guess aprox number of divisions
tol = l/10000.; % a tolerance so fibers can be seen
c = cosd(ang);
%
% draws fibers
%
nlines = 0;
for i=0:n,
pt = 0;
b = (i-n/2)*sep/c; % y = a * x + b
%
% computes first point through intersections at x=0 and y=0
%
if (b >= 0 && b <= H),
pt = pt+1; x(pt) = 0; y(pt) = b;
end,
aux = -b/a;
if (aux > 0 && aux < B),
pt = pt+1; x(pt) = aux; y(pt) = 0;

- 118-
3.2 Laminated Plate Theory

end,
%
% computes second point through intersections at x=B and y=H
%
aux = a*B + b;
if (aux > 0 && aux < H),
pt = pt+1; x(pt) = B; y(pt) = aux;
end,
aux = (H-b)/a;
if (aux >= 0 && aux <= B),
pt = pt+1; x(pt) = aux; y(pt) = H;
end,
%
% plots fiber
%
if (pt > 1),
nlines = nlines + 1;
hline(nlines) = line(x,y,[z+tol,z+tol]);
set(hline(nlines),’color’,fcolor,’linewidth’,2);
set(hline,’parent’,hgroup);
nlines = nlines + 1;
hline(nlines) = line(x,y,[z-tol,z-tol]);
set(hline(nlines),’color’,fcolor,’linewidth’,2);
set(hline,’parent’,hgroup);
end,
end,
%
text(B*1.05,-H*0.05,z,num2str(angle),’color’,fcolor)
axis equal;
%
return

The MATLAB plot for the input laminate

laminate = [0,15,30,45,60,75,90];

is shown in the following Figure

- 119-
3 Constitutive Equations for Laminated Composites

90

75

60

45

30

15

Figure of Exercise 11

Exercise 12 Create a MATLAB function (laminatestiff.m) to compute the struc-


tural stiffness matrices [A], [B] and [D] of a given laminate.
Solution: The code for MATLAB is

function [h,A,B,D,ABBD] = laminatestiff(lamina_name,laminate,t)


%
% function [h,A,B,D,ABBD] = laminatestiff(lamina_name,laminate,t)
%
% Function to create the stiffness matrices A, B and D of a laminate
%
% Input:
% lamina_name = name of the laminas, e.g. ’Boron_Epoxy’
% see laminaprop.m for available laminas
% laminate = laminate stacking sequence in long format bottom-to-top
% e.g. [0,0,30,-30,0,-30,30,0,0]
% t = thickness of each lamina
% t may be a vector if each lamina has different thickness

- 120-
3.2 Laminated Plate Theory

% in such case length(t) = length(laminate), otherwise,


% scalar
%
% Output:
% h = total thickness of the laminate
% A,B,D,ABBD = structural stiffness matrices such that
%
% | A B |
% ABBD = | |
% | B D |
%
% | N | | A B | | e |
% | | = | | | |
% | M | | B D | | k |
%
% where N = in-plane force resultants of the laminate
% M = moment resultants of the laminate
% e = strains in the middle surface of the laminate
% k = curvatures of the laminate
%

%
% all laminas are suposed to be of the same material
% retrieve lamina properties in principal material directions
%
prop = laminaprop(lamina_name); % retrieve properties by name
%
% compute lamina stiffness and compliance matrices
%
[Smat,Qmat] = laminaSQ(prop); % in ppal material directions
%
% obtain laminate z-coordinates
%
[nlaminas,t,~,zmid,~,h] = laminatezs(laminate,t);
%
% loop on for lamina
%
A = zeros(3); B = zeros(3); D = zeros(3);
%
for i=1:nlaminas,
%
% obtain rotated matrices
%
angle = laminate(i); % retrieves angle from stacking sequence
[~,Q] = laminarotate(Smat,Qmat,angle); % rotation
%

- 121-
3 Constitutive Equations for Laminated Composites

% A-matrix (relationship between mid-lamina strains and in-plane loads)


%
Ak = Q * t(i);
A = A + Ak;
%
% B-matrix (relationship between mid-lamina strains and moments)
%
Bk = Ak * zmid(i);
B = B + Bk;
%
% D-matrix (relationship between mid-lamina curvatures and moments)
%
Dk = Q * (t(i) * zmid(i)^2 + (t(i)^3) / 12);
D = D + Dk;
end,
%
% creates global matrix
%
ABBD = [A,B;B’,D];
%
return

The function file (laminatezs.m) used in laminatestiff.m to compute some geo-


metrical parameters of the laminate is

function [nlaminas,t,zbot,zmid,ztop,h] = laminatezs(laminate,t)


%
% function [nlaminas,t,zbot,zmid,ztop,h] = laminatezs(laminate,t)
%
% auxiliar function to compute the z-coordinates of the laminas in a
% laminate
%
% Input:
% laminate = laminate stacking sequence in long format bottom-to-top
% e.g. [0,0,30,-30,0,-30,30,0,0]
% t = thickness of each lamina
% t may be a vector if each lamina has different thickness
% in such case length(t) = length(laminate), otherwise,
% scalar
%
% Output:
% nlaminas = number of laminas
% t = vector of laminae thicknesses
% zbot = coordinates of the bottom of each lamina
% zmid = coordinates of the middle plane of each lamina

- 122-
3.2 Laminated Plate Theory

% ztop = coordinates of the top of each lamina


% h = total thickness of the laminate
%

%
% prepares number and thickness of laminas
%
nlaminas = length(laminate);
if (length(t) == 1 && nlaminas > 1),
t(1:nlaminas) = t;
end,
if (length(t) ~= nlaminas), disp(’**ERROR, check number of laminas’); end,
%
% compute total thickness h, zbot and ztop positions of laminae
% z = 0 is the center (mid-surface) of the laminate
%
h = 0;
for i=1:nlaminas,
zbot(i) = h; h = h + t(i); ztop(i) = h;
end,
zbot = zbot - h/2; ztop = ztop - h/2; zmid = (ztop + zbot)/2;
%
return

The numerical values for [A], [B] and [D] resulting from the function call

[~,A,B,D,~] = laminatestiff(’Glass_Epoxy’,[0,45,90,-45,0],0.001)

in the MATLAB Command Window are

A =
1.0e+008 *
1.8804 0.3715 0
0.3715 1.5128 0
0 0 0.5917
B =
1.0e+004 *
0.0000 0.0000 -1.8383
0.0000 0.0000 -1.8383
-1.8383 -1.8383 0
D =
516.2234 63.2234 0
63.2234 219.0319 0
0 0 109.1011

- 123-
3 Constitutive Equations for Laminated Composites

Exercise 13 Create a MATLAB function (laminatedef.m) to compute the middle-


plane strains and curvatures of a laminate, {ε 0 } and {κ 0 }, subjected to given resul-
tant forces and moments, {N } and {M }.
Solution: The code for MATLAB is

function ek = laminatedef(lamina_name,laminate,t,N,M)
%
% function ek = laminatedef(lamina_name,laminate,t,N,M)
%
% function to compute the deformation of a laminate under
% some resultant loads (in-plane forces and moments)
%
% Input:
% lamina_name = name of the laminae (see laminaprop.m)
% laminate = laminate stacking sequence in full format
% t = thicknesses of the laminae, if scalar, all
% laminae have same thickness, otherwise contains
% a vector with the thickness of each lamina in laminate
% N = vector with in-plane force resultants
% M = vector with moment resultants
%
% Output:
% ek = strain-curvature vector
%

%
% compute structural stiffness global matrix
%
[~,~,~,~,ABBD] = laminatestiff(lamina_name,laminate,t);
%
% solve system of equations
%
NM = [N;M];
ek = ABBD\NM; % inv(ABBD)*NM
%
return

The strains and curvatures resulting from the function call

ek = laminatedef(’Glass_Epoxy’,[0,45,90,45,0],0.01,1e6*ones(3,1),1e3*ones(3,1))

are

- 124-
3.2 Laminated Plate Theory

ek =
0.0003
0.0004
0.0015
0.0012
0.0035
0.0083

We can now plot the deformed laminate by means of the function call

plotplatedef(ek,1);

Figure of Exercise 13

Exercise 14 Create a MATLAB function (laminatestress.m) to compute and plot


the through-the-thickness strain and stress distributions of a deformed laminate.
Solution: The code for MATLAB is

- 125-
3 Constitutive Equations for Laminated Composites

function [e,s] = laminatestress(lamina_name,laminate,t,ek,nfig)


%
% function [e,s] = laminatestress(lamina_name,laminate,t,ek,nfig)
%
% function to compute and plot the strains and stresses of a laminate
%
% Input:
% lamina_name = name of the laminas, e.g. ’Boron_Epoxy’
% see laminaprop.m for available laminas
% laminate = laminate stacking sequence in long format bottom-to-top
% e.g. [0,0,30,-30,0,-30,30,0,0]
% t = thickness of each lamina
% t may be a vector if each lamina has different thickness
% in such case length(t) = length(laminate), otherwise,
% scalar
% ek = strain-curvatures vector
% nfig = if > 0, number of figure to plot strains & stresses
%
% Output:
% e = strains in the laminae
% e(1:3,l) = bottom; e(4:6,l) = top; l = lamina
% s = stresses in the laminae; idem strains
%

%
% all laminas are suposed to be of the same material
% retrieve lamina properties in principal material directions
%
prop = laminaprop(lamina_name); % retrieve properties by name
%
% compute lamina stiffness and compliance matrices
%
[Smat,Qmat] = laminaSQ(prop); % in ppal material directions
%
% compute z-coord. of each lamina
%
[nlaminas,~,zbot,zmid,ztop] = laminatezs(laminate,t);
%
% loop on for lamina
%
for i=1:nlaminas,
%
% obtain rotated matrices
%
angle = laminate(i); % retrieves angle from stacking sequence
[~,Q] = laminarotate(Smat,Qmat,angle); % rotation

- 126-
3.2 Laminated Plate Theory

%
% compute strains, e(1:3,i) = bottom; e(4:6,i) = top
%
e(1:3,i) = ek(1:3) + zbot(i) * ek(4:6);
e(4:6,i) = ek(1:3) + ztop(i) * ek(4:6);
%
% compute stresses, idem
%
s(1:3,i) = Q * e(1:3,i);
s(4:6,i) = Q * e(4:6,i);
%
end,
%
% plots strains and stresses in global laminate axes if nfig > 0
%
if (nfig < 1), return; end,
%
figure(nfig); % activates/creates figure
%
% plots laminate pic at the left of the figures
%
for k = [1,5];
subplot(2,4,k); % 2x4 = 8 plots, activates plot k
for i=1:nlaminas,
%
% color code
%
color = ’w’;
if (abs(laminate(i)) == 0), color = ’c’; end,
if (abs(laminate(i)) ==90), color = ’c’; end,
if (abs(laminate(i)) ==30), color = ’m’; end,
if (abs(laminate(i)) ==45), color = ’y’; end,
if (abs(laminate(i)) ==60), color = ’m’; end,
%
% plots stacking
%
patch([0,1,1,0],[zbot(i),zbot(i),ztop(i),ztop(i)],color);
text(0.5,zmid(i),num2str(laminate(i)),...
’HorizontalAlignment’, ’center’);
end,
set(gca,’XTicklabel’,");
end,
%
% plots strain components
%
for k = 1:3,

- 127-
3 Constitutive Equations for Laminated Composites

subplot(2,4,k+1); % 2x4 = 8 plots, activates plot k+1


for i=1:nlaminas,
patch([0,e(k,i),e(k+3,i),0],[zbot(i),zbot(i),ztop(i),ztop(i)],’c’);
end,
if (k ==3), k = 4; end,
set(gca,’YTicklabel’,"); title([’\epsilon_{’,num2str(k),’}’]);
end,
%
% plots stress components
%
for k = 1:3,
subplot(2,4,k+5); % 2x4 = 8 plots, activates plot k+5
for i=1:nlaminas,
patch([0,s(k,i),s(k+3,i),0],[zbot(i),zbot(i),ztop(i),ztop(i)],’c’);
end,
if (k ==3), k = 4; end,
set(gca,’YTicklabel’,"); title([’\sigma_{’,num2str(k),’}’]);
end,
%
% arranges y-axis of plots to limit values so it plots nicely
%
zmin = min(zbot); zmax = max(ztop);
for k=1:8; subplot(2,4,k); set(gca,’ylim’,[zmin,zmax]); end,
%
suptitle(’Strains and stresses in the laminate (global laminate axes)’);
%
return

The resulting plot using the strains and curvatures (ek) computed just above

ek = laminatedef(’Glass_Epoxy’,[0,45,90,45,0],0.01,1e6*ones(3,1),1e3*ones(3,1));
[e,s] = laminatestress(’Glass_Epoxy’,[0,45,90,45,0],0.01,ek,1);

is

- 128-
3.2 Laminated Plate Theory

Strains and stresses in the laminate (global laminate axes)


ε ε ε
1 2 4

0.02 0

0.01 45

0 90

−0.01 45

−0.02 0

0 2 4 0 5 0 1 2
−4 −4 −3
x 10 x 10 x 10

σ σ σ
1 2 4

0.02 0

0.01 45

0 90

−0.01 45

−0.02 0

0 2 4 0 2 4 0 2 4
7 7 7
x 10 x 10 x 10

Figure of Exercise 14

- 129-
3 Constitutive Equations for Laminated Composites

3.3 Symmetric and antisymmetric laminates

Composite structures have the ability to be tailored to match individual loading


and stiffness requirements by means of crossplying considerations. The foregoing
formulation has been developed for a generic laminate, formed by N perfectly bonded
unidirectionally reinforced laminae, each one of them being particularly defined by
its in-plane elastic properties, fiber orientation relative to the laminate axes, location
through the out-of-plane direction and thickness. In this section, we specialize the
derived force-moment-strain-curvature general relation given in Eq. (409) to some
specific composite layups commonly used in practical engineering applications.

3.3.1 Single-layered orthotropic plate

The simplest geometrical case to be analyzed consists of a laminate formed by one


lamina only (actually it is not a true laminate, but a single-layered laminate or
simply a lamina). If we consider a specially orthotropic lamina, then the lamina
reinforced direction coincides with the laminate x-axis and the reduced stiffness
matrix expressed in the laminate coordinate system X = Oxyz is directly given by
Eq. (171). In this case, N = 1, z (1) = 0 and t(1) = h, so the stiffness matrices Eqs.
(406)–(408) are readily calculated

 
Q11 Q12 0
[A] = [Q] h =  Q12 Q22 0 h (411)
0 0 Q44
[B] = [0] (412)
h3 h2
[D] = [Q] = [A] (413)
12 12

All the coefficients of the extensional-bending stiffness matrix [B] vanish, so the
force equations and the moment equations are fully uncoupled. Accordingly, such a
laminate does never experience coupling between the extensional-shear and bending-
twist behaviors. Since we are considering a specially orthotropic lamina, for which
α = 0o , there is neither extensional-shearing coupling (A14 = A24 = 0) nor bending-
twist coupling (D14 = D24 = 0) effects. However, we remark that these kinds of
coupling will be present for a generally orthotropic lamina (for which 00 6= α 6= 90o )
by virtue of the apparent material constants given in Eqs. (390) and (391). Using
the engineering constants of the orthotropic material, the force-strain-curvature and
the moment-strain-curvature of the single-layered orthotropic laminate under

- 130-
3.3 Symmetric and antisymmetric laminates

study are
 Ex νyx Ex 
0
   1 − νxy νyx 1 − νxy νyx  0 
 Nx     ε1 
N

= h νxy Ey Ey 
ε0 (414)
 y   20 
0 
 
Nxy  1 − νxy νyx 1 − νxy νyx  ε4
0 0 Gxy
Ex νyx Ex 
0
   1 − νxy νyx 1 − νxy νyx  0 
 Mx  h3    κ1 
My =

 νxy Ey Ey 
 κ0 (415)
  12  0   20 
Mxy  1 − νxy νyx 1 − νxy νyx  κ4
0 0 Gxy

which, evidently, reduce to Eqs. (378) and (379) when the special case of a single-
layered isotropic laminate is regarded, namely
    
 Nx  1 ν 0  ε01 
ν 1 0 Eh
ε0
 
N = A A= (416)
 y   20 

1−ν 1 − ν2
Nxy 0 0 ε4
2
    
 Mx  1 ν 0  κ01 
ν 1 0 Eh3
κ0
 
My = D D= (417)
 20 

  1−ν 12 (1 − ν 2 )
Mxy 0 0 κ4
2

3.3.2 Symmetric laminates

As we have mentioned above, a symmetric laminate is formed by a set of lami-


nae which are symmetrically laid up (regarding material systems, fiber orientations
and geometrical properties) relative to the resulting laminate middle surface. Of
course, a symmetric laminate may be irregular and/or hybrid. The main feature of
these symmetric composite plates arises when ones specializes Eq. (407) taking into
account the symmetry condition. This leads to

N N/2  
X (k) (k) (k)
X (k) (k) (k) (k ′ ) (k ′ ) (k ′ )
[B] = [Q] t z = [Q] t z + [Q] t z = [0] (418)
k=1 k=1

where the lamina k′ is the symmetric counterpart of the lamina k within the lam-
′ ′ ′
inate, so [Q](k) = [Q](k ) , t(k) = t(k ) , z (k ) = −z (k) and the coupling stiffnesses
vanish. Note the number of layers has been taken to be even, but the same result
is obtained for an odd-numbered laminate because z (k) = 0 for the central lamina.
Therefore, every material-and-geometry symmetric laminate has no coupling effects
between axial-shear forces and bending-twist moments. For example, a symmetric
flat laminate does not become curved due to thermal contraction effects after the
curing process, which is an important feature to be taken into account.

- 131-
3 Constitutive Equations for Laminated Composites

[0,90]S

x y

Figure 50: Example of a symmetric cross-ply laminate (unbonded view).

Symmetric cross-ply laminate

If the reinforced direction of adjacent layers of the symmetric laminate are alternately
aligned with the reference x-axis (α = 0o ) and y-axis (α = 90o ), then we refer to it
as a symmetric cross-ply laminate. An example is shown in Figure 50. Although a
generic symmetric laminate can contain either an even or an odd number of laminae,
only laminates with an odd number of layers can simultaneously fulfill both the
symmetry condition and the alternation of right angles. For a symmetric cross-ply
laminate conformed by the same type of fiber-matrix layers, the expression of the
reduced stiffness matrix of each pair of symmetric layers k − k′ expressed in laminate
coordinates is
 
Q11 Q12 0

[Q](k) = [Q](k ) =  Q12 Q22 0  if α(k) = 0o (419)
0 0 Q44

or  
Q22 Q12 0

[Q](k) = [Q](k ) =  Q12 Q11 0  if α(k) = 90o (420)
0 0 Q44

where Eqs. (386)–(391) have been used. The structural stiffnesses [A] and [D] result
from the summation of all these contributions, so we obtain
   
A11 A12 0 D11 D12 0
[A] =  A12 A22 0  and [D] =  D12 D22 0  (421)
0 0 A44 0 0 D44

and the extensional-shear and bending-twist couplings are prevented. As a result,


regarding coupling effects, symmetric cross-ply laminates macroscopically behave as
a single-layered specially orthotropic plate, but with the advantage that the former
is reinforced in both the laminate x-axis and y-axis (note that Q11 contributes to
both A11 and A22 and both D11 and D22 ).

- 132-
3.3 Symmetric and antisymmetric laminates

Symmetric angle-ply laminate


If the reinforced direction of adjacent layers of the symmetric laminate are alter-
nately oriented about a specific angle α (0o 6= α 6= 90o ) and −α, then we refer to it
as a symmetric angle-ply laminate. An example is shown in Figure 51. Again, only
laminates with an odd number of layers can simultaneously fulfill both the sym-
metry condition and the alternation of angles. For a symmetric angle-ply laminate
conformed by layers with the same material preferred properties, the expression of
the reduced stiffness matrix of each pair of symmetric layers k − k ′ expressed in
laminate coordinates is
 
Q11 Q12 Q14

[Q](k) = [Q](k ) =  Q12 Q22 Q24  if α(k) = +α (422)
Q14 Q24 Q44
or  
Q11 Q12 −Q14
(k ′ )
[Q](k) = [Q] =  Q12 Q22 −Q24  if α(k) = −α (423)
−Q14 −Q24 Q44
where Eqs. (386)–(391) have been used. In this case
   
A11 A12 A14 D11 D12 D14
[A] =  A12 A22 A24  and [D] =  D12 D22 D24  (424)
A14 A24 A44 D14 D24 D44

and the structural stiffnesses A14 , A24 , D14 and D24 do not vanish in general, so there
may exist extensional-shear and bending-twist coupling behaviors. However, for
example, note that the corresponding summation of the coefficient A14 (analogously,
A24 ) is
  
XN (N −1)/2
X  Q14 t if N = 1, 5, ...
(k) (k)
A14 = Q14 t(k) = Q14 + 2 Q14  t =

k=1 k=1 −Q14 t if N = 3, 7, ...
(425)
where the laminate is assumed to be regular and the central lamina to be oriented
at +α. This last result for the extensional-shear stiffness coupling compares to that
obtained for the extensional stiffness A11 (analogously, A22 , A12 or A44 )
N
X (k)
A11 = Q11 t(k) = N Q11 t = Q11 h (426)
k=1

It is observed that A11 increases with the number of laminate plies, whereas A14
does not. A similar result is obtained for the same components of the bending
stiffness matrix [D]. Thus, for many-layered symmetric angle-ply laminates, both
the extensional-shear (A14 and A24 ) and the bending-twist stiffness couplings (D14
and D24 ) reach values much lower than the other stiffnesses. This fact, together
with the higher structural shear stiffness A44 that these laminates present when
compared to the symmetric cross-ply laminates (because of the fibers orientation),

- 133-
3 Constitutive Equations for Laminated Composites

[-30,30]S

x y

Figure 51: Example of a symmetric angle-ply laminate (unbonded view).

causes that they are used more often in practical applications. Finally, regarding
coupling effects, symmetric angle-ply laminates macroscopically behave as a single-
layered generally orthotropic plate, but with the advantage that the extensional-
shear and the bending-twist stiffness couplings can be comparatively reduced for
many-layered laminated plates (with h fixed).

3.3.3 Antisymmetric laminates


Another special type of laminates for which some stiffness simplifications arise are
the so-called antisymmetric laminates. A laminate is regarded antisymmetric if the
material system (fiber-matrix) and the geometrical properties (z (k) and t(k) ) of a
given pair of laminae are symmetric about the laminate middle surface, but their
fiber orientations present some kind of antisymmetry. In this respect, two main
antisymmetries are commonly defined, one relative to the orientation angle α = 0o
and another relative to the direction α = 45o . The first case applies for layers with
a representative (generic) direction α ∈ (−90o , 90o ], so a given pair of antisymmetric
plies respect to α = 0o is simply denoted as [..., α, ..., −α, ...]. The latter case applies
for layers with a representative (specific) angle α = 0o or α = 90o , so a given pair
of antisymmetric laminae relative to α = 45o are represented by [..., 0, ..., 90, ...] or
[..., 90, ..., 0, ...], respectively. In general, both types of antisymmetry associated to
different angles α(k) may be encountered within the same antisymmetric laminate.
Attending to these definitions, antisymmetric laminates will generally contain an
even number of laminae. However, note that the very special case of a laminate being
antisymmetric with respect to α = 0o containing a single central lamina aligned with
the directions α = 0o or α = 90o is also antisymmetric, although it is odd-numbered.
The first important difference between antisymmetric laminates and symmetric
laminates is that, since for a given pair of layers k − k ′ the material systems are
equal but the reinforced directions are different, their apparent reduced stiffnesses
take also different values and the coupling stiffnesses Bij do not vanish in a general
case. In this case, we obtain
 
N/2 
X  B11 0 B14

[B] = [Q](k) − [Q](k ) t(k) z (k) =  0 −B11 B24  (427)
k=1 B14 B24 0

- 134-
3.3 Symmetric and antisymmetric laminates

so certain extensional-bending coupling effects will appear. The reason for the in-
volved results B22 = −B11 and B12 = B44 = 0 can be seen below. The second
difference with respect to symmetric laminates is that neither extensional-shear nor
bending-twist couplings are present in antisymmetric laminates. These second stiff-
ness simplifications arise from the reduced stiffness matrices of a pair of antisym-
metric laminae k − k′
   
Q11 Q12 0 Q22 Q12 0

[Q](k) =  Q12 Q22 0  and [Q](k ) =  Q12 Q11 0  (428)
0 0 Q44 0 0 Q44

for a given pair [..., 0, ..., 90, ...] (or vice versa for the case [..., 90, ..., 0, ...]) and
   
Q11 Q12 Q14 Q11 Q12 −Q14

[Q](k) =  Q12 Q22 Q24  and [Q](k ) =  Q12 Q22 −Q24  (429)
Q14 Q24 Q44 −Q14 −Q24 Q44

for a given pair [..., α, ..., −α, ...] (or vice versa for the case [..., −α, ..., α, ...]). Hence,
due to the summation of opposite-signed terms, on the one hand, or vanishing terms,
on the other, it is straightforward to obtain
   
A11 A12 0 D11 D12 0
[A] =  A12 A22 0  [D] =  D12 D22 0  (430)
0 0 A44 0 0 D44

and the extensional-shear and bending-twist couplings are always prevented. Note
that these coefficients are not zero, in general, for a symmetric laminate. However,
although these usually undesired couplings vanish for this type of laminates, they are
not used much in practice, essentially due to the always present extensional-bending
coupling stiffnesses Bij .

Antisymmetric cross-ply laminate


The expression for the stiffness coupling matrix [B] can be further simplified if only
one type of the aforementioned antisymmetries is present within a given laminate.
If the reinforced direction of adjacent layers of an antisymmetric laminate are al-
ternately aligned at α = 0o and α = 90o , then we refer to it as an antisymmetric
cross-ply laminate. An example is shown in Figure 52. Then, from Eqs. (427) and
(428), we obtain
 
N/2 
X  B11 0 0

[B] = [Q](k) − [Q](k ) t(k) z (k) =  0 −B11 0  (431)
k=1 0 0 0

where B22 = −B11 because for each term added to B11 we add the same term, but
with the sign changed, to B22 . Furthermore, note that A11 = A22 and D11 = D22
for this type of laminates. Finally, an antisymmetric cross-ply laminate is always
even-numbered.

- 135-
3 Constitutive Equations for Laminated Composites

[0,90,0,90]

x y

Figure 52: Example of an antisymmetric cross-ply laminate (unbonded view).

[-30,30]

x y

Figure 53: Example of an antisymmetric angle-ply laminate (unbonded view).

Exercise 15 Show that the structural coupling coefficients B11 and B22 = −B11 of
an antisymmetric cross-ply laminate go to zero as the number of plies increases for
a fixed laminate total thickness h.
Solution: Use Eq. (431) along with Eq. (428) for alternate pairs [..., 0, ..., 90, ...]
and [..., 90, ..., 0, ...]. Check the results using the MATLAB function laminatestiff.m.

Antisymmetric angle-ply laminate


If the reinforced direction of adjacent layers of an antisymmetric laminate are alter-
nately oriented about a specific angle α (0o 6= α 6= 90o ) and an angle −α, then we
refer to it as an antisymmetric angle-ply laminate. An example is shown in Figure
53. Using Eqs. (427) and (429), the matrix [B] is simplified to
 
N/2 
X  0 0 B14

[B] = [Q](k) − [Q](k ) t(k) z (k) =  0 0 B24  (432)
k=1 B14 B24 0

Again, an antisymmetric angle-ply laminate is always even-numbered. Aside, note


that the same expression for [B] would be obtained even though different orientations

α(k) = −α(k ) were present within the laminate for different pairs of layers.
(k) (k ′ ) (k) (k ′ ) (k) (k ′ ) (k) (k ′ )
The relations Q22 − Q22 = −(Q11 − Q11 ) and Q12 − Q12 = 0 = Q44 − Q44
are always obtained for any pair (and type) of antisymmetric laminae k − k′ , which

- 136-
3.3 Symmetric and antisymmetric laminates

justifies the expression of the matrix [B] shown above for a generic antisymmetric
laminate, Eq. (427).

Exercise 16 Show that the structural coupling coefficients B14 and B24 of an anti-
symmetric angle-ply laminate approach to zero as the number of plies increases for
a fixed laminate total thickness h.
Solution: Use Eq. (432) along with Eq. (429) for alternate pairs [..., α, ..., −α, ...]
and [..., −α, ..., α, ...]. Check the results using the MATLAB function laminatestiff.m.

3.3.4 Other specific ply layups


Quasi-isotropic laminates
A laminate is quasi-isotropic when

• All the plies have the same thickness t (i.e. the laminate is regular) and have
the same preferred in-plane material constants (for example, they are made of
the same fiber-matrix material system).

• The fibers are aligned with the orientations αi = for i = 0, 1, ..., n − 1 and
n
n ≥ 3, with the same number of layers in each direction i.

Then, using Eqs. (386)–(391) and Eq. (406) we obtain the following expression
for the matrix [A]
 
A11 A12 0
[A] =  A12 A11 0
 
 (433)
A11 − A12
0 0
2
which, first, depends on two independent parameters and, second, adopts the same
form than the extensional matrix of a single-layered isotropic laminate, namely Eq.
(416). This equivalence is more clearly seen if we identify A11 = AQI = EQI h/(1 −
2 ) and A
νQI 12 = νQI AQI , so
 
1 νQI 0
EQI h  νQI 1 0 
[A] = 2   (434)
1 − νQI 1 − νQI
0 0
2
where EQI and νQI are the equivalent Young’s modulus and equivalent Poisson’s
ratio of the quasi-isotropic laminate, respectively. Therefore, the expression of [A]
for a quasi-isotropic laminate is independent of the system of representation being
used (rotated within the laminate middle plane). In other words, a quasi-isotropic
laminate behaves isotropically when it is subjected to in-plane forces only. Finally,
note that due to the presence of the through-the-thickness coordinates z (k) in Eqs.
(407) and (408), the structural stiffness matrices [B] and [D] of a quasi-isotropic
laminate do not simplify in the same manner than the matrix [A]. Hence, there may
exist both extensional-bending coupling and anisotropic bending behavior, which
are the differences between a quasi-isotropic laminate and a single-layered isotropic
laminate (for which [B] vanishes and [D] has also an isotropic representation).

- 137-
3 Constitutive Equations for Laminated Composites

Exercise 17 Use the previous MATLAB functions to plot a quasi-isotropic laminate


and to compute its associated matrices [A], [B] and [D].
Solution: A possible example is:

plotlaminate([0,120,240,240,120,0],1,1);
[~,A,B,D,~] = laminatestiff(’Graphite_Epoxy’,[0,120,240,240,120,0],0.01)

The results are

120

240

240

120

Figure of Exercise 17

and

A =
1.0e+009 *
4.8740 1.5707 0
1.5707 4.8740 0
0 0 1.6516

- 138-
3.3 Symmetric and antisymmetric laminates

B =
1.0e-007 *
-0.0745 -0.0262 -0.0047
-0.0262 -0.1420 0
-0.0093 0 -0.0466
D =
1.0e+006 *
2.7230 0.2220 -0.0889
0.2220 0.6999 -0.2616
-0.0889 -0.2616 0.2462

We notice that A11 = A22 , A14 = A24 = 0 and A44 = (A11 − A12 ) /2 (because
the laminate is quasi-isotropic), that [B] vanishes numerically speaking (because the
laminate is symmetric) and that [D] does not have an isotropic representation (be-
cause the laminate is not single-layered isotropic but quasi-isotropic).

Balanced laminates

A laminate is said to be balanced when for every fiber-reinforced lamina oriented


about an angle α, there is an identical layer (same thickness and principal material
properties) oriented about the angle −α. Each laminae pair has its own associated
direction α. For a given pair of laminas k − k′ (not necessarily laid up symmetrically
with respect to the middle plane of the laminate), we obtain
   
Q11 Q12 Q14 Q11 Q12 −Q14

[Q](k) =  Q12 Q22 Q24  and [Q](k ) =  Q12 Q22 −Q24  (435)
Q14 Q24 Q44 −Q14 −Q24 Q44

so
 
N/2 
X  A11 A12 0
(k ′ )
[A] = [Q](k) + [Q] t(k) =  A12 A22 0  (436)
k=1 0 0 A44

and the (in-plane) shear-extension coupling vanishes in the principal laminate axes,
which is the distinctive feature of this type of laminated plates. As for the quasi-
isotropic laminate, no simplifications arise for [B] and [D] in the most general (un-
symmetrical) case, so extension-bending and bending-twist couplings may be present
in a balanced laminate.

Exercise 18 Use the previous MATLAB functions to plot an unsymmetrical bal-


anced laminate and to compute its associated matrices [A], [B] and [D]. Compare
the results (in terms of vanishing entries in the structural stiffness matrices) to those
of an antisymmetric balanced laminate and a symmetric balanced laminate.

Solution: We leave this exercise to the reader.

- 139-
3 Constitutive Equations for Laminated Composites

3.4 Temperature and moisture effects in a laminate


As mentioned in Section 2.5, changes of temperature and changes in humidity may
result in important stresses and strains in the laminate. Composite laminates are
cured at high temperatures and work in very different conditions. In contrast to
isotropic unrestrained solids, in an unrestrained laminate even a homogeneous tem-
perature distribution may cause significant stresses. Hygromechanical stresses are
similar in nature to thermal stresses and may be analyzed altogether as seen in
Section 2.5.2.
Usually laminates are analyzed in an uncoupled manner. First the temperature
(and/or moisture) distribution throughout the laminate is obtained. Then the ther-
mal stresses and strains, as well as their laminate resultants are computed taking
into account the computed temperature distribution. Finally the problem is solved
to obtain displacements and strains.

3.4.1 Temperature distribution in the laminate


In order to obtain the temperature or moisture distribution, the heat equation must
be solved. The heat equation is obtained from thermodynamical principles, see
Section C.1 of Appendix C, which for an anisotropic case is
∂T 1
= K : ∇∇T (437)
∂t c
where K is the tensor of thermal conductivities and c is the heat capacity. This
partial differential equation is of the parabolic type. For the hygroscopic case, the
equation to be solved is Fick’s second identity which results into an identical math-
ematical problem
∂H
= K H : ∇∇H (438)
∂t
where K H is the tensor of moisture (mass) diffusion coefficients. In a complex
structure with different boundary conditions the solution for these equations may be
obtained using any software that solves parabolic problems (heat transfer). However,
in laminates the situation is usually simple. The laminate has two large dimensions
being the thickness (in z − direction) small. Then the heat transfer problem has
one relevant direction and it can be reasonably analyzed just in that direction, the
through-the-thickness direction (in z − direction). In this case T = T (z) so
     
∂T /∂x 0 0 0 0
∇T =  ∂T /∂y  =  0  and ∇∇T =  0 0 0  (439)
∂T /∂z ∂T /∂z 2
0 0 ∂ T /∂z 2

and the heat equation is


∂T 1 ∂2T
= K33 2 (440)
∂t c ∂z
Let us rename K33 /c = K̄. This equation may be solved by separation of variables.
Assuming that T can be written in terms of 1D functions

T (t, z) = τ (t) ζ (z) (441)

- 140-
3.4 Temperature and moisture effects in a laminate

we have
∂T dτ ∂2T d2 ζ
= ζ, = τ (442)
∂t dt ∂z 2 dz 2
and
∂T 1 ∂2T dτ d2 ζ 1 dτ 1 d2 ζ
= K33 2 ⇒ ζ = K̄τ 2 ⇒ = (443)
∂t c ∂z dt dz K̄τ dt ζ dz 2
Of the last identity, the left-hand-side only depends on the variable t, whereas the
right-hand-side only depends on the variable z. Since it must hold for any values of
t and z, this implies that both quantities must be the same constant, i.e.
1 dτ 1 d2 ζ
= = −k2 (444)
K̄τ dt ζ dz 2
where −k2 is an arbitrary constant (the sign is really a consequence of the analysis
of the possible solutions and boundary conditions). Then we have the following set
of 1D ordinary differential equations:

dτ /dt + k2 K̄τ = 0
(445)
d2 ζ/dz 2 + k2 ζ = 0
which are of constant coefficients. The solution for k 6= 0 is obtained from the roots
of the algebraic equations

r + k2 K̄ = 0 and s2 + k 2 = 0 (446)

whereupon
2 K̄t
τ = Ãk e−k (447)
ikz −ikz
ζ = B̃k e + C̃k e = B̄k sin kz + C̄k cos kz (448)

i.e.
2 K̄t
T (t, ζ) = (Bk sin kz + Ck cos kz) e−k
where Bk and Ck are constants to be determined from initial conditions and bound-
ary conditions. For k = 0, we readily obtain from Eqs. (445)

τ = Ā (449)
ζ = B̄z + C̄ (450)

thereby
T (t, ζ) = Bz + C (451)
where, again, B and C are constants to be determined. Since all are solutions to
the differential equation, the general solution is the sum of all of them, i.e.
Z ∞
2
T (t, z) = Bz + C + (Bk sin kz + Ck cos kz) e−k K̄t dk (452)
0

To illustrate the possible solutions, assume that the laminate has a temperature
Tt at z = h and Tb at z = 0 (note8 ). Assume that enough time has passed to consider
8
You can later redefine z ∗ = z − h/2.

- 141-
3 Constitutive Equations for Laminated Composites

thermal equilibrium (t → ∞). In this case, dτ /dt = 0 so k = 0 and the solution is


necessarily of the form (all Bk and Ck must be zero)
T (z) = Bz + C (453)
which is the steady-state solution. In our case

Tb = C Tt − Tb z
⇒B= ⇒ T = (Tt − Tb ) + Tb (454)
Tt = Bh + C h h
So in a laminate the steady-state solution is just linear in z.
Assume now that initially we have a distribution of temperatures T (0, z) =
T0 (z). Take it to be the addition of a linear and a nonlinear contribution
T0 (z) = T̄0 (z) + T̃0 (z) (455)
The linear part is determined as in the previous case
z
T̄0 (z) = (Tt − Tb ) + Tb (456)
h
and for the second part:
Z ∞
T̃0 (z) = (Bk sin kz + Ck cos kz) dk (457)
0

We must prescribe T̃0 (0) = T̃0 (h) = 0 to account for the boundary conditions
T (t, z = h) = Tt and T (t, z = 0) = Tb . Then since cos kz 6= 0 for z = 0, we have
Ck = 0 and if Bk 6= 0, k takes only some discrete values so the T̃0 (z) are zero at
z = 0 and z = h.

sin kh = 0 ⇒ kh = nπ ⇒ k = , n≥1 (458)
h
Thus we obtain the Fourier series

X nπz
T̃0 (z) = Bn sin (459)
n=1
h
If we multiply both sides by sin (mπz/h) and take the integral
Z h ∞ Z h
mπz X nπz mπz
T̃0 (z) sin dz = Bn sin sin dz (460)
0 h n=1 0 h h
Because if m 6= n all second integrals vanish and for m = n the value of that integral
is h/2, we have
Z
2 h nπz
Bn = T̃0 (z) sin dz (461)
h 0 h
so the solution is

z X nπz
Bn e−(n π /h )K̄t sin
2 2 2
T (t, z) = (Tt − Tb ) + Tb + (462)
h h
n=1

For thermal loads in laminates K̄/h2


is large such that thermal equilibrium in the
laminate (steady-state solution) is obtained in hours. For hygroscopic loads K̄H /h2
is a much larger number which also depends strongly on the temperature (higher
for higher temperatures), and to arrive at hygroscopic equilibrium in the laminate
may take months.

- 142-
3.4 Temperature and moisture effects in a laminate

3.4.2 Hygrothermal loads in the laminate


Once the temperature and/or moisture distribution is obtained, the stresses in each
lamina k in an arbitrary system of reference X (i.e. for example that of the laminate)
may be computed using Eq. (197)

{σ}(k) = [Q](k) {ε}(k) − [Q](k) {α}(k) ∆T (k) (463)

where for each point of each lamina in the laminate


 
{ε}(k) = ε0 + z κ0 , z ∈ [zk−1 , zk ] (464)

so  0  
{σ}(k) = [Q](k) ε + z κ0 − [Q](k) {α}(k) ∆T (k) (465)
Then, the stress resultants for the laminate are computed using Eqs (394) and (395),
so for each lamina (k)
Z h/2 N Z
X zk
{N } = {σ} dz = {σ}(k) dz
−h/2 k=1 zk−1
N Z zk
 0  0 X
= [A] ε + [B] κ − [Q](k) {α}(k) ∆T (k) dz
k=1 zk−1

(466)

Z h/2 N Z
X zk
{M } = z {σ} dz = z {σ}(k) dz
−h/2 k=1 zk−1
N Z zk
  X
= [B] ε0 + [D] κ0 − z [Q](k) {α}(k) ∆T (k) dz
k=1 zk−1

(467)

We see that the governing equations are unchanged except for the last terms in
Equations (466) and (467), which depend on the temperature for each lamina. The
terms {N T } and {M T } are known as thermal loads, and are defined as
 
{N } = [A] ε0 + [B] κ0 − {N T } (468)
 0  0
{M } = [B] ε + [D] κ − {M T } (469)
with
N Z
X zk
{N T } = [Q](k) {α}(k) ∆T (k) dz (470)
k=1 zk−1
XN Z zk
{M T } = z [Q](k) {α}(k) ∆T (k) dz (471)
k=1 zk−1

- 143-
3 Constitutive Equations for Laminated Composites

Then, we can write


 
   ..
   0 
{N } {N T }  [A] . [B]  ε
 ···  +  ···  =  ··· · · ·  ···  (472)
   0
{M } {M T } .. κ
[B] . [D]

So we note that temperature effects may be considered simply as artificial loads. In


the case of homogeneous temperature (which is a typical analysis) we have ∆T (k) =
∆T and
 
    ..   0 
{N } {N̂ T }  [A] . [B]  ε
 · · ·  +  · · ·  ∆T =  · · · · · ·  ···  (473)
   0
{M } {M̂ T } . κ
[B] .. [D]

where {N̂ T } and {M̂ T } are the values for a unit temperature increment.
Finally, we mention that in the case that we also want to analyze the effects of
a moisture change in the laminate, we simply modify α(k) and ∆T (k) as commented
in Section 2.5.2.

3.4.3 Recovery of stresses in the laminae


 
Once we have computed the solution using Eq. (472), we know ε0 and κ0 , so
it is immediate to obtain the strains for each lamina using Equation (464) and the
stresses using Equation (463). These stresses may be used to verify the safety factor
of the laminate.

3.5 Piezoelectric effects in the laminate


The Equations to be employed to account for piezoelectric effects depend on the
purpose of performing such analysis and the task for what piezoelectricity is used
for.
For example in morphing applications where the electric field E (k) is given for
each lamina, we can use Eq. (206) (considering plane stress and temperature effects)
to write
(k) (k) (k) (k) (k) (k) (k)
{σ}X = [Q]X {ε}X − [Q]X {α}X ∆T (k) − [e]X {E}X (474)
where (note the systems of representation)
   (k)
0 0 e13k 0 0 e13
(k) (k)
[e]Xk =  0 0 e23k  , [e]X =  0 0 e23  (475)
0 0 0 0 0 e33

and  (k)
 E1 
(k)
{E}X = E (476)
 2 
E3

- 144-
3.5 Piezoelectric effects in the laminate

We usually can just operate using the reduced notation


 (k)
e13
(k) (k)
[e]X ≡  e23  and {E}X ≡ E3 (477)
e33

for the usual case of materials doped in the z−direction. Then, in a similar form as
before we can write (we leave to the reader the steps)
 
      ..   0 
{N } {N T } {N P }  [A] . [B]  ε
 ···  +  ···  +  ···  =  ··· · · ·  · · · 
  (478)

{M } {M T } {M P } . κ0
[B] .. [D]

where
N Z
X zk
{N P } = [e](k) {E}(k) dz (479)
k=1 zk−1
XN Z zk
{M P } = z [e](k) {E}(k) dz (480)
k=1 zk−1

are the piezoelectric loads. Hence, for practical purposes, we just add the computed
piezoelectric loads to the hygrothermal and mechanical loads.
Another application is just to measure the stresses and strains in a laminate (for
example for health monitoring). Then we can use Eq. (209) for constant dielectric
displacements, so
E = −g : σ (481)
then for each lamina the electric field is a direct consequence of the stresses
(k) (k) (k)
{E}X = − [g]X {σ}X (482)

- 145-
3 Constitutive Equations for Laminated Composites

- 146-
4 Mechanical analysis of composite laminates
In the previous chapter we have addressed the basic mechanics of a laminate. In
Chapter 2 we have seen the principles of the mechanical behavior and some simple
different failure criteria for a lamina. The purpose of this chapter is to complete the
analysis of a composite laminate addressing the most important failure mechanisms
of the laminate. In the first part of this chapter we study the possible failure
mechanisms of a lamina inside a laminate and then we perform the strength analysis
of a laminate. This first part studies the failure of a given point of the laminate of a
composite structure. In the second part we will focus on the failure of the composite
structure itself from a structural level.

4.1 Mechanical failure in composites


Composite laminates may fail in many different manners. Any failure mode implies
progressive failure of the different laminae. However, composite laminates work in
a similar manner as hyperstatic structures. In a hyperstatic structure, the failure of
one member does not imply a total collapse, but merely a reduction of the overall
stiffness of the structure. Similarly, a failure of a lamina does not usually imply
the failure of the complete laminate because even reducing totally the load-carrying
capacity of a lamina, it still serves as a redistributing device of the loads to the
adjacent laminae. In a hyperstatic structure, once a member fails, it carries zero
loads, but otherwise the structure continues behaving in a linear elastic manner
until failure of the next member; and so on until a mechanism is formed. In a
similar manner, a composite laminate continues carrying load and behaving in a
linear elastic manner until next ply fails.
However, there is a marked difference with hyperstatic structures. Whereas when
a member fails in a hyperstatic structure, it is usually assumed that it can carry no
load thereafter, when a ply fails it can frequently carry load, although not in the
“direction” it failed. This happens because laminae may fail in very different forms.
For example, if fibers fail, the matrix may still carry load in the transversal direction.
If failure of the lamina is due to matrix failure in the transversal direction, fibers
can still hold the same loading. Furthermore, there are differences between failure
in tension and in compression for both fiber and matrix. Fibers in compression
may buckle. In this case fibers may still carry more load but with a significant
reduction of stiffness (that reduction given by the lateral restraining capabilities of
the matrix). Failure of matrices in compression may be more catastrophic in some
cases. In this case delamination may be a consequence. Delamination is also typical
in laminates because of three-dimensional effects at borders or because of impacts
(the wave creates tensile stresses between laminae). This last case is of special
concern because no apparent damage can be seen from outside the laminate.
From an engineer perspective, it is important to determine the strength char-
acteristics and behavior until collapse of a given type of laminate. Of course the
most accurate way of obtaining such behavior is to test the laminate in a labora-
tory under the assumed load distribution and working conditions. However, this is
an expensive task and, hence, engineers want an inexpensive but accurate-enough

- 147-
4 Mechanical analysis of composite laminates

procedure to determine such behavior. Several procedures have been published and
are implemented in commercial software depending on the degree of accuracy and
level of complexity desired. In next section we will briefly address these procedures.
However we note that care must be exercised9 and that this is a complex topic
currently under research.
In the previous paragraphs we have assumed that the laminate is a somehow
uniformly loaded structure. In other words, we have been looking only to a single
stress point in the structure made of composites. Hence we have been looking at
stress-point failure in an equivalent plate. However, we know that structural failure
may not only be due to simple static loads under linear behavior. Column buckling
is a typical case. In a column under compressive loads, there is a bifurcation point
(buckling load) from which the straight column is no longer able to sustain the
loads and the new equilibrium point is in a new path that may bring too large
displacements to the structure. Then the critical limit load may be well below the
value predicted by the usual static analysis. Furthermore, under dynamic loads,
resonance phenomena may also mean that the collapsing load is even an order of
magnitude less than that of linear static analysis. Then, it is apparent the need for a
study of the composite structure as a whole. This is a task usually performed using
finite elements, because analytical solutions can only be obtained for very simple
cases. However it is instructive to study some of those simple cases in order to
understand the phenomena. For this task we need to establish the field equations
for a composite plate under static, linear buckling and dynamic loads. We dedicate
some sections of this chapter to understand the overall behavior of a composite plate.
Finally, there are many other issues involving the failure in a composite struc-
ture and a proper design of a composite structure. Such issues are, for example,
global fracture and fatigue failure, joint failures, composite beam design, sandwich
composites, etc. An in-depth treatment of such issues is completely out of the scope
of these notes. However, in the last part of this chapter we will briefly address these
issues.

4.2 Strength of laminates


In order to compute the strength of a laminate, we need first to understand the failure
mechanisms of a lamina within a laminate. Then we will introduce the differences
in predictions of failure criteria and finally we will compute the failure behavior of
a laminate in what is known as progressive failure analysis.

4.2.1 Failure mechanisms in a lamina


Unlike homogeneous materials, a fiber reinforced composite lamina may fail in many
different ways. This is simply true because it is composed of different bonded mate-
9
The reason of Prof. Hashin to decline participate in WWFE-I (World-Wide Failure Exercise-I)
was (taken from M.Hinton’s presentation at NAFEMS 2011 Word Conference):
...my only work in this subject relates to failure criteria of unidirectional fiber
composites... I must say to you that I personally do not know how to predict the
failure of a laminate (and furthermore, that I do not believe that anybody else does).

- 148-
4.2 Strength of laminates

rials with very different material properties. However, there are some failure modes
that are the most common ones. These modes are summarized in Figure 54 and
may be classified as follows.

Fiber failure
• Tensile mode. Fiber tensile failure mode happens with traction loads in the
direction of the fibers. If only one fiber fails, stresses are redistributed to
neighbouring ones through the surrounding matrix. We have seen how this re-
distribution takes place in Chapter 2 (a simplified micromechanical approach
to the lamina tensile strength was also presented in that Chapter). However,
once a sufficient amount of fibers fail, they no longer can carry the load and the
lamina fails. The Xt experimental value already takes this effect into account
because it is obtained in macroscopic tests.
• Compression mode. This failure mode occurs with compression loads in the
direction of the fibers. In this failure mode fibers are compressed and micro-
buckling takes place. The limit compression load depends on the bending
stiffness of the fibers and on the supporting material, which determine the
buckling load of the fibers. In this case all fibers buckle at similar loads or
strains, but failure is often more progressive until full matrix failure occurs.
The Xc experimental value already takes all these effects into account because
this value is also obtained in macroscopic tests. However, the failure load also
depends on moisture contents and temperature, being lower for higher temper-
ature and humidity. This happens because the matrix supporting properties
have such dependence. Of course it is more difficult to take this into account
when obtaining Xc . Several values of Xc may be given for different temperature
and moisture contents.

Matrix failure
• Tension mode. Matrix tension failure mode happens when traction loads are
applied in a direction perpendicular to the fibers. Since the lamina is stiffer
in the fiber direction, it is not common that the matrix fails first with cracks
perpendicular to that direction. However no reinforcement exists in the plane
perpendicular to the fibers, so the matrix may fail with cracks parallel to the
fibers. In this case, the matrix failure is more of a sudden. Redistribution of
loads can only take place through neighbouring laminae with different fiber
directions, see Figure 55. Furthermore, the actual maximum load in this case
also depends on the temperature. Polymeric matrices usually get stiffer (carry-
ing relatively more load) and more brittle (less tough) at lower temperatures.
These effects are more difficult to consider in a single Yt experimental value
for the matrix. This fracture mode is often known as failure “mode A” of the
matrix.

• Compression modes. This failure mode occurs with compression loads in a


direction perpendicular to the fibers. There are two compression failure modes

- 149-
4 Mechanical analysis of composite laminates

Figure 54: Main failure modes of a fiber reinforced composite lamina within a lam-
inate.

- 150-
4.2 Strength of laminates

Figure 55: Failure of the matrix in a middle laminate in a cross-ply laminate. Load
can still be carried by surrounding laminae (After [4]).

for the matrix of a lamina which, as for similar reasons as tensile modes,
generate cracks parallel to the fiber directions.
The first one is known as “mode B” and simply corresponds to fracture
in a plane perpendicular to the compressing load with no relevant relative
displacements between both parts of the lamina unless there is an additional
in-plane shear stress. Usually this failure mode is more important in fatigue
situations. The load-deformation curve in this case is highly nonlinear and
failure occurs at a given strain level. For linear analysis it is frequent to
linearize the stress limit as shown in Figure 56: the limit stress is computed
as the stress that corresponds to a linear behavior up to the prescribed limit
strain.
The second mode is more critical and is known as matrix failure “mode C”.
In this mode two cracking planes are generated at a given angle which may
go from 0o to about 50o − 55o . In this case traction pressure is transferred in
a wedge effect in the through-the-thickness direction to neighbouring laminae
favouring delamination and buckling of neighbour laminae.
Both types of matrix compression failure modes are slightly favoured by hot
and wet conditions. Usually it is difficult to take all these effects into account
and consider both failure modes into a single Yc value, so some criteria include
additional material parameters.

Mixed failure
• Mixed shear mode. The shear test to investigate this failure mode is usually
performed in an angle-ply laminate at ±45o . The failure mode is very similar
to the matrix compression mode B, but including fiber debonding. The stress-
strain relation is also highly nonlinear, so it is also frequent to linearize the
stress limit for a given strain as shown in Figure 56. Complete failure under this
mode (as in mode B) is not usual, but may be very important in fatigue. Hot

- 151-
4 Mechanical analysis of composite laminates

Figure 56: Determination of the limit stress to be used in linear analysis for failure
modes that behave in a nonlinear manner.

and wet conditions facilitate matrix deformation and matrix-fiber debonding.


Hence, this mode is negatively affected by these conditions. The limit stress
is given in tests by S.

• Mixed compression (kink) mode. This failure mode occurs with compression
loads in the direction of the fibers but depends mainly on the matrix stiffness
properties. It is a fiber kinking mode similar to the fiber compression one, but
unlike the latter, both matrix and fibers fail simultaneously after the fibers
reach a given tilt angle. Including this failure mode in any failure criteria
usually means including a new material parameter. Most failure criteria do
not include this failure mode explicitly.

Lamina buckling

Lamina buckling within a laminate produces (and it is favoured by) delamination.


Delamination is due to normal tensile stresses in the through-the-thickness direc-
tion or due to interlaminar stresses because of high out-of-plane shear loads. Then
compressive in-plane stresses trigger buckling of the lamina which at the same time
produce more tensile normal stresses. This failure mode is more of a laminate failure
mode (structural mode) than a lamina mode itself because it involves the study of
the composite structure. This failure mode is also frequently present in stringers
in plate and shell composite structures, where the shear and normal stresses in the
bonding between components may be relatively high.

- 152-
4.2 Strength of laminates

4.2.2 Additional considerations on failure criteria for laminae


As already introduced in Chapter 2, there are many failure criteria for laminae.
These failure criteria attempt to macroscopically predict when a lamina is failing
within a laminate.

Types of failure representative scalar measures


There are two types of indices usually associated to failure criteria: the so-called
Failure Index and the Strength Ratio.

Failure Index The Failure Index F I is an index that takes values less or equal
than one. More specifically, a given criterion may be written in terms of F I as
F I (σxx , σyy , ...) ≤ 1 (483)
where
FI = 1 (484)
indicates that the corresponding failure has taken place. Actually, the Failure Index
may take values greater than one, but that situation is not possible in practice (just
theoretically) because a lamina is considered to fracture when a given criterion is
satisfied (i.e. F I = 1) and then it cannot resist larger loads (for which F I > 1).
The Failure Index strongly depends on the criteria being used. For example,
under a maximum stress criterion, the failure index in tension is
σxx
F I1 = (485)
Xt
with 0 ≤ F I1 ≤ 1. In this case there is a failure index for each failure mode con-
sidered: fiber-tension, fiber-compression, matrix-tension, matrix-compression and
shear. For this criterion, one overall F I value might be computed to be the maxi-
mum of all F Ii indices.
Under a combined criteria such as the Tsai-Hill criterion, the failure index is
2 2
σyy 2
σxy
σxx σxx σyy
FI = + − + (486)
X2 Y2 X2 S2
with 0 ≤ F I ≤ 1, which is an inherently quadratic (and hence less intuitive) measure
in terms of stresses.
For the Hoffman and Tsai-Wu failure criteria the center (C) of the failure ellipsoid
may not be located at the origin of the stress space. Hence, F I may take negative
values for some stress states. In this cases, we note that the failure index definition
Eq. (483) could be alternatively defined as
s 
F I (σxx , σyy , ...) − F I σxx C , σ C , ...
yy
FI = C , σ C , ...
 ≤1 (487)
1 − F I σxx yy

which is more intuitive for quadratic criteria because we are taking the square root.
Furthermore, note that 0 ≤ F I ≤ 1.
The actual value of the Failure Index is somewhat arbitrary, so it is frequent to
use the more intuitive Strength Ratio index.

- 153-
4 Mechanical analysis of composite laminates

Strength Ratio The Strength Ratio SR is an index that takes values greater or
equal than one. More specifically, a given criterion may be written in terms of SR
as
SR (σxx , σyy , ...) ≥ 1 (488)
where
SR = 1 (489)
indicates that the corresponding failure has taken place. Again, we can theoretically
define the Strength Ratio to be lower than one, but this also represents that the
lamina has failed. However, this may be useful for design purposes in order to
quantify in what measure the design loads exceed the allowable limits.
Mathematically, the SR is the linear scaling factor λ that must be applied to the
loads to reach failure. Assuming that the load increase is proportional (i.e. there is
no change in load distribution or pattern), the stresses are
   ∗ 

 σxx 
 
 σxx 


 σyy 
 
 σ∗ 

 yy

 
 
 


σzz
 ∗
σzz

λσ = σ ∗ ⇒ λ = ∗ (490)

 σ xy 
 
 σxy 

   
 σyz
 
   σ∗
 


   yz




σzx σzx

where σ is the actual stress and σ ∗ is the specific limit stress that preserves the
actual load pattern. Thus, SR ≡ λ = 1 means that there is no remaining load
capacity according to a given failure criterion. If SR > 1, the loads can be scaled
by SR, i.e., they can be increased a (SR − 1) × 100% (if SR = 1.2, then the load
can be increased a 20% without failure). In a linear criterion the strength ratio is
simply the inverse of the failure index.
For example using the Tsai-Hill criterion
!
∗ 2 ∗ 2 ∗ 2 2 2 2
(σxx ) (σyy ) σxx

σyy

(σxy ) σxx σyy σxx σyy σxy
2
+ 2
− 2
+ 2
= 1 ⇒ λ2T H + − + = 1 (491)
X Y X S X2 Y2 X2 S2

The strength ratio is then computed as


1 1 1
SRT H ≡ λT H = s =√ = (492)
2
σxx 2
σyy σxx σyy 2
σxy F IT H F IT H
+ − +
X2 Y2 X2 S2

Using Hoffman’s criterion and Equation (490) we have


2 2 2
σxx σyy σxx σyy Xt + Xc Yt + Yc σxy
−λ2H − λ2H + λ2H + λH σxx + λH σyy + λ2H 2 = 1 (493)
Xt Xc Yt Yc Xt Xc Xt Xc Yt Yc S

i.e.
!  
2 2
σ2 σyy σxx σyy σxy Xt + Xc Yt + Yc
λ2H − xx − + + 2 + λH σxx + σyy − 1 = 0 (494)
Xt Xc Yt Yc Xt Xc S Xt Xc Yt Yc

- 154-
4.2 Strength of laminates

The strength ratio SRH ≡ λH is then computed solving the quadratic equation and
selecting the positive λH value.
This type of scalar measures are the key piece of the progressive failure analysis.
In these cases it is frequently convenient to keep track of the actual failure mode
through the failure indices (the maximum one corresponds to the failure mode) or
through the strength ratios (the minimum one corresponds to the failure mode).

Types of failure criteria


There are two types of failure criteria: interactive failure criteria and non-interactive
failure criteria.

• Non-interactive failure criteria are criteria which do not consider the effect
of other stresses than the main stress/strain for a failure mode; i.e. do not
produce combinations of stresses or strains. Non-interactive failure criteria
are, for example, the maximum stress criteria and the maximum strain criteria.
These criteria return a failure index for each failure mode.

• Interactive failure criteria. This type of criteria performs combinations of


stress and/or strain components in order to account for the effects of all of
them in each failure mode. Here there are two sub-types:

– Combined failure mode criteria. In this case usually a single equation


is used to assess the failure of the lamina. There is no apparent clear
distinction of the failure mode (although they are already accounted for
somehow in the failure criteria). Then, only one failure index is computed
which simply determines whether the lamina has failed or not. Examples
are the Hoffman and the Tsai-Wu failure criteria.
– Distinctive failure mode criteria. In these type of criteria different failure
modes are considered and an interactive failure index is computed for
each failure mode. In this case there is not only information about when
a lamina fails, but also about the actual failure mode.

Distinctive failure mode criteria are the most advanced and recent ones. They
are mainly phenomenological models obtained through testing with important en-
gineering thinking and understanding about the mechanisms by which failure takes
place. The reason to distinguish the failure mode of a lamina is to improve the
layout of the laminate. Progressive failure analysis also takes advantage of the pos-
sibilities given by the knowledge of the failure mode. For example, some terms of
the stiffness matrices of each laminae can be kept whereas others are degraded, leav-
ing some remanent mechanical properties. Distinctive failure mode criteria are not
only accurate, but also give important insight in how laminae fail. However, the
drawback is that they are more complex than nondistinctive criteria.
Due to the foregoing reasons, distinctive failure mode criteria are the best ones
to perform progressive failure analysis. Before explaining this specific analysis, we
introduce two frequently used distinctive criteria for unidirectionally reinforced lam-
inae, namely those by Hashin and Puck.

- 155-
4 Mechanical analysis of composite laminates

Hashin failure criterion Using Hashin’s criterion, four possible failure modes
are taken into account

1. Fiber-tension mode. This mode may occur if σxx > 0, and happens when the
following failure index for the mode is equal to (or larger than) one
2 2
σxy 2
σxx σxz
F I1 = + + (495)
Xt2 2
Sxy 2
Sxz

This is basically the (three-dimensional) Tsai-Hill criterion Eq. (251) when


only stresses in the x−direction are considered.

2. Fiber-compression mode. This mode failure mode may occur if σxx < 0, and
is activated when the following index for the mode is equal to (or larger than)
one
2
σ2 σxy σ2
F I2 = xx2 + 2 + xz 2
(496)
Xc Sxy Sxz
This equation is again derived from the Tsai-Hill criterion, but considering the
compression strength.

3. Matrix-tension mode
 
σyy + σzz 2 σyy σzz 2
σxy 2
σxz 2
σyz
F I3 = − 2
+ 2
+ 2
+ 2
(497)
Yt Syz Sxy Sxz Syz

4. Matrix-compression mode
    2 2 2
Yc2 σyy + σzz σyy + σzz σyy σzz σxy 2
σxz σyz
F I4 = 2
−1 + − 2
+ 2
+ 2
+ 2
(498)
4Syz Yc 2Syz Syz Sxy Sxz Syz

Sometimes a term (σxx /K)2 is added to take into account the influence of fibers
in compression (i.e. mixed failure mode), where K is a parameter. Note that
shear terms influence all failure modes.

Puck failure criterion Puck’s criterion is currently one of the best and most
regarded failure criterion. The theory behind the criterion is physically sound and
the resulting expressions are still usable in engineering analysis. The main drawback
is that it needs more material parameters (something an engineer always hates!).
The theory of Puck goes back to the work of Coulomb in the eighteen century and
to the work of Mohr in the break of the 20th century. It contains the ideas behind
the Mohr-Coulomb criterion used in solids and concrete.
Puck understood that there are six different possible stressings that result in five
possible failure modes under combined loading, as we will see below. The six possible
stressings are shown in Figure 57. Two of them correspond to fiber failure (F F ) and
four to interfiber-failure (IF F , he avoided to use the word matrix probably because
fiber failure may also imply matrix failure in some cases; we keep for the moment his
nomenclature). It is seen that the failures associated to these stressings are closely
related to the failure modes of Figure 54. The possible stressings are labeled by

- 156-
4.2 Strength of laminates

Figure 57: Puck’s six different stressings and corresponding failure planes. Sub-
scripts k and ⊥ stand for parallel to the fibers and perpendicular to the fibers
respectively. Superscript t stands for tension and superscript c stands for compres-
sion

Puck with subscripts k or ⊥ whether the direction is parallel or perpendicular to


the fiber directions respectively, and with superscripts t or c whether they are in
tension or in compression respectively, see Figure 57. For each of the 6 stressings
Puck assigned a tested strength (resistance)

Rkt , Rkc , R⊥
t c
, R⊥ , R⊥⊥ , R⊥k (499)

The F F modes present no difficulty. It is simply assumed that fracture takes


place when
 σxx
 ≥ 1 if σxx > 0
 Rkt

 σxx
 c ≥1
 if σxx < 0
Rk
where Rkc < 0.
In Figure 57 it is seen that the failure plane depends on the nature of the stresss-
ing (normal or shearing) and on the direction. Moreover, for each stressing the
failure plane is not necessarily the action plane, and this fact must be considered to
correctly account for the actual action surface when computing the strengths. For
example, the failure plane is different for σ⊥t than for σ c . It is also different for τ
⊥ ⊥⊥
c
than for τ⊥k . Furthermore, whereas for σk the mechanism is a fiber buckling and
matrix crushing, for σ⊥ c is a shear failure (because strength in shear is lower than

in compression), basically the same as for τ⊥⊥ but very different from the failure
caused by τ⊥k (a cleavage fracture failure). In this last case the natural failure plane
would imply shearing the fibers. Then, for each loading it is important to identify
the failure plane, which at the same time depends on the stresses themselves.

- 157-
4 Mechanical analysis of composite laminates

Figure 58: Puck’s failure plane, after G. Lutz: The Puck theory of failure in lami-
nates in the context of the new guideline VDI 2014 Part 3.

Then, in order to assess Inter-Fiber Failure in a given plane, we must decompose


the stress tensor on that plane. Since we know that the fiber direction will always be
contained in the failure plane (see IF F modes in Figure 57), we can characterize the
direction of the failure plane by a single angle θf p (see Figure 58), which represents
the angle formed between the normal to the plane under study and the y-axis.
Then, the stresses on that plane are composed of a normal stress σn , a shear stress
in the direction of the fibers τnx and a shear stress perpendicular to the fibers τnt ,
as shown in Figure 58, such that (the reader can verify this expression following a
similar procedure to that explained to obtain Eq. (92) departing from Eq. (89))
 

 σ xx 
     σ 
 

yy
0 c2 s2 0
 
 σn  2sc  0   

2 2 σ zz
τnt = 0 −sc sc 0 c − s
 0  (500)
   σxy 
τnx 0 0 0 c 0 s   

 σ 
 yz 

 

σxz
where c = cos θf p and s = sin θf p .
From the values of R⊥ t , Rc , R
⊥ ⊥⊥ , R⊥k (the ones involving matrix failure), we can
compute the resistance values in the action plane (hence the superscript A)
tA A A
R⊥ , R⊥⊥ , R⊥k (501)

The reason there is no R⊥ cA is because σ will never produce fracture in its own
n
action plane (the material never separates or moves relatively, so we can consider
R⊥cA extremely high). However, we must consider σ in the failure of the acting
n
plane because, like in Mohr-Coulomb theory, a tensile stress favours fracture and a
compressive one makes it more difficult. It is similar to what happens in friction
theory. The friction force is Fr = µN , where µ is the friction resistance and N
is the normal load. Puck then adopted the same expressions to be added to the
shear resistances taking into account that compressive stresses are negative, that

- 158-
4.2 Strength of laminates

is −p⊥⊥ σn and −p⊥k σn , where p⊥k and p⊥⊥ are the friction coefficients in the
A and RA
action plane parallel and perpendicular to the fibers, associated to R⊥k ⊥⊥
respectively. The modulus of the shear stress in the action plane is

τ 2 = τnt
2 2
+ τnx (502)

Weighting each contribution in the action plane by its respective resistance for com-
pressive σn leads to propose the following quadratic criterion

 2 !2
τnt τnx
1= A
+ A
for σn < 0 (compression) (503)
R⊥⊥ − p⊥⊥ σn R⊥k − p⊥k σn

For the case of σn > 0, no extra resistance is applied to the shear components,
tA that has to be
but the contribution of σn is to provide a fracture resistance R⊥
overcome:

 2 !2  2  
τnt τnx σn σn
1= A
+ A
+(1 − c) tA
+c tA
for σn > 0 (tension) (504)
R⊥⊥ R⊥k R⊥ R⊥

where c is a material parameter that controls the slope of the curve at the meeting
between surfaces given by Eqs. (503) and (504) for σn = 0. The failure envelope
given by these equations is depicted in Figure 59.
Of course now the key factor in finding if there is failure according to Puck’s
criterion is to guess the failure plane, a task far from obvious. However, one can
develop a numerical algorithm varying θf p , compute the criterion according to Eqs.
(503) and (504) and select the θf p with the maximum failure index, which will be
the failure plane. If θf p = 0 and σn < 0 we are in failure mode B. If θf p > 0 and
σn < 0 matrix failure mode is mode C.
Since obtaining θf p is a rather cumbersome task, expressions available in the
literature are based on five failure indices, one for each failure mode. One of the
typical sets of equations with the hypothesis of plane stress used in laminates (which
also simplifies the criterion) is

 σxx
 Fiber tension: F I1 = X with σxx > 0

t
σ (505)
 Fiber compres.: F I2 = xx with σxx < 0

Xc

- 159-
4 Mechanical analysis of composite laminates

Figure 59: Puck failure envelope showing inter fiber failure (IFF ) modes of a unidi-
rectional lamina under plane stress, after M. Knops: The Puck theory of failure in
fiber polymer laminates: Fundamentals, verification and applications. Springer 2007.
Stresses in the failure plane σn , τnx and τnt (axes of the figure) relate to the lamina
stresses through Eq. (500). In particular, σn = σyy , τnx = σyx and τnt = σyz = 0
for mode B (θf p = 0o ).

- 160-
4.2 Strength of laminates

 s

  σ 2   
Yt 2 σyy 2
 σ 
 xy yy

 Matrix mode A: F I 3 = + 1 − p ⊥k + p ⊥k


 S S Yt S





 with σyy > 0

 r
  σ 2 


 xy σyy 2 σ 
yy

 Matrix mode B: F I 4 = + p ⊥k + p ⊥k

 S S S

σyy R

 with σyy < 0 ≤ ≤


 σxy S ∗

 "

 2  2 #

 σ xy R σ xy Yc

 Matrix mode C: F I5 = +


 Yc S Yc σ22





 σxy S ∗

 with σyy < 0 ≤ ≤
σyy R
(506)
where
Yc S p
R= , p⊥k = p⊥⊥ , S ∗ = S 1 + 2p⊥⊥
2 (1 + p⊥⊥ ) R
In these expressions we have already substituted the equivalent symbols R⊥k =
t = Y , Rc = Y . We note that for the case of plane stress
S, Rkt = Xt , Rkc = Xc , R⊥ t ⊥ c
a new parameter is needed, p⊥k . A usual value for epoxy matrices is p⊥k = 0.35,
which corresponds with a maximum failure plane angle of about θf p = 53.5o .

4.2.3 Progressive failure analysis


Composite materials are mostly assumed to behave linearly until fracture. However,
a composite laminate does not suddenly fail once a lamina within the laminate has
failed. It is frequent that the laminate may still sustain a larger load after such event
has happened. However, of course, the overall stiffness of the laminate will degrade
after such events occur. In this sense, it is similar to linear-elastic hyperstatic
structures (think of a hyperstatic truss structure).
If one member of a hyperstatic structure fails, the structure may not collapse, and
may even sustain larger loads at the cost of a reduced stiffness. Once a member has
failed, the behavior of the structure is still linear-elastic, so if loads are eliminated,
the structure recovers its original undeformed strain and stress-free configuration.
Of course, the structure remains damaged for further loading events. Furthermore,
it is also possible that after the first member fails, the rest of the structure is not
able to sustain the load and, hence collapse. Whether this happens or not, depends
on the member distribution itself and on the load pattern.
In a similar way, when a lamina in a laminate fails, the laminate may still sustain
further loading or fail depending on whether the remaining laminae is able to do so.
The laminate stacking sequence may be in this case crucial.
On one side, one would like to have a laminate such that all laminae fail at the
same time, hence taking benefit of the strength properties of all and each lamina. In
this case we would probably have large savings both in weight and in cost. However, a
sudden failure, with no remanent resistance capacity is extremely dangerous because

- 161-
4 Mechanical analysis of composite laminates

the structure may fail without any previous warning. This is one of the advantages
of ductile structures: after reaching the yield limit, the material may still sustain
a larger loading at the cost of larger, permanent deformations. Even though the
permanent deformations are in principle not desirable, they serve to warn that the
structure is overloaded and may be near failure.
Composite laminae are brittle in essence. The behavior is almost linear until
failure. However, within a laminate, if a lamina fails, the other laminae may still
carry their share of that load and even carry further more load. The cost of that
lamina failure is a reduced stiffness and, hence, larger displacements. Then a be-
havior similar to that of a ductile structure is obtained... if the laminae stacking is
adequate.
The first task is then to determine the strength ratios of all laminae prior to the
first failure. These strength ratios bring us an idea of how far from failure is each
lamina. Very different strength ratios may mean that we are loosing money and
weight. However, very similar strength ratios may mean that all laminae may fail
simultaneously in an unpredictable moment and way.
Therefore it is of paramount importance to determine the strength ratios of
the laminate. However, it is also important to notice that once a lamina fails, the
behavior of the laminate may differ substantially. First, several laminae may fail
at the same time before the laminate sustains further load. Second, the geometric
distribution of laminae may lose its initial configuration. For instance, a symmetric
laminate may no longer be symmetric, or a balanced laminate may no longer be a
balanced laminate. These facts should also be taken into account in order to predict
the behavior of the laminate after the first failure.
The previous paragraphs show the importance of developing a so-called progres-
sive failure analysis (PFA). In these type of analyses the laminate is subjected to
increasing loads, tracking the laminate failure sequence and the associated loads and
deformations. Loss of specific geometric properties may also be assessed.
The main difference of PFA with a nonlinear elastoplastic analysis is that PFA
may be performed linearly because it is a sort of piecewise linear analysis. More
specifically, it simply consists of several successive linear analyses. In order to un-
derstand the reason, the reader has to recall that laminae behave in a linear manner,
so once a lamina fails, in principle that laminae simply has no more carrying capacity
and no stiffness. If the loading is released, the laminate just returns to an unde-
formed configuration, with no remanent strains. If loaded again, the slope of the
load-deflection curve will have been damaged; i.e. once a lamina fails, the laminate
mechanically behaves as if that lamina has never been there before.
Assume that we have prescribed (fixed) load vectors given by N̂ and M̂ . Pro-
gressive failure analysis assume that loading is scaled in a proportional way; the
loads being of the form
N = λN̂ ; M = λM̂ (507)
where λ is the scaling factor. Conversely, if we want to perform a displacement-based
PFA, then we prescribe some strains and curvatures ε̂0 and κ̂0 , and we assume they
are proportionally increased by means of

ε0 = λε̂0 ; κ0 = λκ̂0 (508)

- 162-
4.2 Strength of laminates

Note that both analyses are not equivalent


 0 0 because after the first lamina fails, pro-
portionality between {N , M } and ε , κ (given by the structural stiffness matri-
ces [A], [B] and [D]) changes. Thus, it is not the same to fix a proportionality for
prescribed {N̂ , M̂ } during all the analysis than for constant {ε̂0 , κ̂0 }.
Stresses are linearly dependent on the loads during each linear analysis, what-
ever the progressive computation being followed (load-driven or deformation-driven).
Hence, they are equally dependent on the λ proportionality parameter, see Section
4.2.2, page 154. The next failure event will occur when the λ parameter reaches the
value given by the lowest Strength Ratio of all laminae. Hence, the procedure to
detect the next failure sequence is to simply compute all SR of the laminae and to
take the minimum, i.e.
λ = min (SR (m, l)) (509)
m,l

where l accounts for the number of laminae and m are the possible failure modes of
each lamina. It is important to note that the moduli of N̂ and M̂ are irrelevant.
Since failures take place at values given by Eqs. (507), only the “direction” is
important. For example, if we change the modulus of the initial load, the values
of λ will change accordingly such that failure takes place exactly at the same load
(limit) level.
In summary, the procedure for progressive failure analysis, assuming proportion-
ality of loads, follows:

1. Compute laminate strains and curvatures of the laminate for the given load
pattern N̂ , M̂ .

2. Compute strains and stresses of each lamina in the lamina principal material
axes.

3. Compute strength ratios for each lamina.

4. Compute the minimum strength ratio, Eq. (509). The failed lamina is the one
with minimum strength ratio. The failure mode of that lamina is that with the
minimum strength ratio (or maximum failure index) within those computed
for that lamina.

5. Degrade the properties of the failed lamina; i.e. take for example Ex = Ey =
Gxy = 0. Usually small numbers are used in order to avoid numerical problems.
Compute the new laminate structural stiffness matrices [A], [B] and [D].

6. If there is remanent stiffness, go to step 1 to compute next failure event.


Otherwise all laminae have failed, so exit the analysis.

We want to make here several remarks

• If the failure criteria is distinctive, we could degrade only the corresponding


stiffness parameters. For example if fibers have failed, we can take Ex = 0 for
that lamina, but still keep Ey in its original value. Other options are possible,
of course.

- 163-
4 Mechanical analysis of composite laminates

• If the criteria is distinctive, a lamina may fail first (at certain event) in one
mode and later (at another event) in other different mode.

• When a lamina fails, several other laminae may abruptly fail immediately
after. The way to check it is to verify the strength ratio Eq. 509 of the
subsequent events. If the λ value of event k + 1 is less than that of event k,
then these failure events occur at the same time. If the λ value of event k + 2
is less than that of the reference event k, then they also take place at the same
time. And so on. It may look surprising that subsequent events may have
less strength ratio than previous ones. This does not mean that the failure
sequence is incorrect since the different SR for each event are computed for
different structures. If the failed lamina at event k had not failed, the one at
event k + 1 would have neither failed.

• After computing the sequence for all lamina failures, as described above, the
full laminate collapse event may be computed as that with maximum associ-
ated strength ratio. After that event occurs, the rest of the laminae will fail
all of a sudden, see Figures 60 (left) and 63.

• In an actual laminate failure sequence, there are jumps in the load-deformation


curves. If the process is load-controlled once a lamina has failed, the load path
will be that of the damaged laminate, so even though loads at failure are kept
constant, deformations abruptly increase to the new equilibrium ones so the
load-deformation curve will have a horizontal line. On the other hand, if the
process is deformation-controlled, there will be a sudden drop on the loading to
the new equilibrium values, so the load-deformation curve will have a vertical
line, see Figure 60.

• It is possible to enhance the PFA using damage models. In these cases damage
is activated for values of failure indices close to one and it progresses usually
following an exponential relaxation equation. However, we note that in this
case, the analysis becomes a fully nonlinear analysis and the λ parameter must
be smoothly increased after the first failure event.

Exercise 19 Describe in a flowchart the procedure to perform  0a progressive


failure
0
analysis using proportionality of kinematic variables; i.e. of ε , κ .
Solution: We leave this task to the reader.

Exercise 20 (Project) Create a set of Matlab functions to perform a Progressive


Failure Analysis for a given set of laminate, load pattern {N̂ , M̂ }, failure criterion
of laminae and degradation scheme for material properties.
Solution: This task is left to the reader. Figures 61 to 63 show the results obtained
for a [0, 90, 0] Glass-Epoxy laminate with total thickness h = 3 mm, under the load
pattern corresponding to a single axial load N̂x = 1 N and using Hashin’s criterion.
Figure 61 shows the failure sequence of the laminae within the laminate as the
load is being increased during the load-controlled analysis.

- 164-
4.3 Governing equations of a laminated plate

Figure 60: Progressive failure analysis. Left: load control case (load is always
increased). Right: displacement control case (deformation is always increased).
Note that the laminate failures occur at different points for the load-controlled case
(at fourth event) and the deformation-controlled case (at fifth event).

Figure 62 shows how the structural stiffnesses of the laminate degrade as the
PFA evolves and the failure events are taking place. It is interesting to observe
that the coupling stiffnesses Bij are zero for event 0 (original laminate) and after
event 1 (central lamina has failed) because in both cases the laminate is symmetric.
Then, after event 2 takes place (one lamina at 0o fails) the coupling coefficients Bij
take values different from zero because the laminate is no longer symmetric. That
is, one has to be aware that the geometrical characteristics of a laminate change
during a PFA, so very different responses may appear for the same load pattern
due to qualitative changes of the laminate structural stiffness matrices [A], [B] and
[D] (i.e. unexpected extension-shear-bending-twist coupling behaviors may appear or
disappear after each lamina failure occurs).
Finally, Figure 63 shows that the maximum SR is obtained at event 2. In other
words, since SR3 < SR2 , event 3 occurs immediately after event 2 (with no incre-
ment of load) and the laminate collapses at event 2.

4.3 Governing equations of a laminated plate


In this section we are going to introduce the governing equations for the classical
plate theory for a composite laminate. We first introduce the equilibrium equations
in terms of stress resultants including dynamic and second order effects for the
z−direction. These equilibrium equations in terms of stress resultants are the same
for any isotropic or anisotropic plate. Then we introduce the equilibrium equations
in terms of displacements, usually referred to as the field equations.

- 165-
4 Mechanical analysis of composite laminates

−3 −3
x 10 x 10
1.5 1.5

1 0º 1 FT

0.5 0.5

Total failure
0 90º 0 MT

−0.5 −0.5

−1 0º −1 FT

−1.5 −1.5
0 1 2 3
Failure event

Figure 61: First figure of Exercise 20. Left: Schematic view of the initial laminate
stacking sequence [0, 90, 0]. Right: Laminae failure sequence using Hashin’s criterion
(M T ≡ M atrix T ension mode, F T ≡ F iber T ension mode). Laminate fails at
event 2 (see Fig. 63).

7 4
x 10 x 10
14 140 6
A D B
11 11 11
A D B
12 22 120 22 22
A D 5
12 12 B12
A D
33 33 B
10 100 33
4
D component
A component

B component

8 80
3
6 60

2
4 40

1
2 20

0 0 0
0 1 2 3 0 1 2 3 0 1 2 3
Failure event Failure event Failure event

Figure 62: Second figure of Exercise 20. Degradation of laminate structural stiff-
nesses.

- 166-
4.3 Governing equations of a laminated plate

6
x 10
2.2

1.8

1.6
Load vector multiplier (SR)

1.4

1.2

0.8

0.6

0.4

0.2

0
1 2 3
Failure event number

Figure 63: Third figure of Exercise 20. Strength Ratios for N̂x = 1 N associated to
each lamina failure. Event 2 corresponds to the full laminate collapse event.

4.3.1 Equilibrium equations in terms of stress resultants


We recall from Eqs. (394) and (395) that the stress resultants are
Z h/2 

Nx (x, y) = σxx (x, y, z) dz 

−h/2 



Z h/2 

Ny (x, y) = σyy (x, y, z) dz (510)
−h/2 



Z h/2 



Nxy (x, y) = σxy (x, y, z) dz 

−h/2

Z h/2


Mx (x, y) = zσxx (x, y, z) dz 

−h/2 



Z h/2 

My (x, y) = zσyy (x, y, z) dz (511)
−h/2 



Z h/2 



Mxy (x, y) = zσxy (x, y, z) dz 

−h/2

where we have indicated explicitly the dependence on the coordinates (x, y). It is
important to emphasize that, until now, we have been analyzing a given point (x, y)
of a laminate plate or, as a very particular case, a whole laminate with uniform
internal forces and moments and middle-plane strains and curvatures (recall the
Matlab exercises of the previous Chapter). In this and the following sections, we will

- 167-
4 Mechanical analysis of composite laminates

study a more general case in which loading and deformation variations throughout
the plate are possible.
Consider the differential plate element of Figure 64. Stress resultants in the
differential element dx × dy × h are due to the stresses throughout the thickness
according to Eqs. (510) and (511). Additional to these stress resultants there are
the following shear stress resultants in the z−direction
Z h/2 

Qx ≡ Qxz (x, y) = σxz (x, y, z) dz 



−h/2
Z h/2 (512)


Qy ≡ Qyz (x, y) = σyz (x, y, z) dz 


−h/2

Until now we have assumed that σxz = σyz = 0 when deriving the Kirchhoff plate
theory. However, we will later see that these out-of-plane stresses indeed do not
vanish in some situations and that they can be approximately computed. We rewrite
Eqs. (510) to (512) in index notation as
Z h/2 Z h/2 Z h/2
Nαβ = σαβ , dz ; Mαβ = zσαβ dz ; Qαz = σαz dz (513)
−h/2 −h/2 −h/2

where both α and β (in-plane indices) can take the values x and y. Then matrix or
Voigt representation may be also applied to Eqs. (513).
Since all the stress resultants in Eqs. (510), (511) and (512) depend on (x, y),
there may be variations across the differential element, as shown in Figure 64c for
the resultant in-plane forces and moments (defined per unit plate width). Then,
assuming there are no in-plane body loads, equilibrium of forces in the x−direction
yields    
∂Nx ∂Nxy
Nx + dx − Nx dy + Nxy + dy − Nxy dx = 0 (514)
∂x ∂y
i.e. dividing by dxdy
∂Nx ∂Nxy
+ =0 (515)
∂x ∂y
Similarly, in the y−direction

∂Nxy ∂Nxy
+ =0 (516)
∂y ∂y

These equations can be generalized in order to account for body loads and dynamic
inertial forces using the equilibrium equations at a point, see Equation (1016)1

 ∂σxx ∂σxy ∂σxz

 + + + bx = ρüx


 ∂x ∂y ∂z

 ∂σ
yx ∂σyy ∂σyz
+ + + by = ρüy (517)

 ∂x ∂y ∂z



 ∂σzx ∂σzy ∂σzz

 + + + bz = ρüz
∂x ∂y ∂z

- 168-
4.3 Governing equations of a laminated plate

Figure 64: Equilibrium in a differential element of a plate: a) Stresses involved.


b) Resultant forces and moments per unit width considering a possible in-plane
variation of the stresses. c) View in the x − y plane.

- 169-
4 Mechanical analysis of composite laminates

For example, for non-vanishing volumetric and inertial terms, b and ρü respectively,
the first equation yields upon integration through the thickness
Z h/2 Z h/2 Z h/2 Z h/2
∂σxx ∂σxy
dz + dz + bx dz = ρüx dz (518)
−h/2 ∂x −h/2 ∂y −h/2 −h/2

i.e. Z h/2
∂Nx ∂Nxy
+ + fx = ρüx dz (519)
∂x ∂y −h/2
where Z h/2
fx (x, y) = bx (x, y, z) dz (520)
−h/2
is the body load in direction x per unit plate surface dxdy, in the case it exists. Note
we have neglected the contribution of the out-of-plane stresses σxz and σyz . For the
inertial term we recall Eq. (366) so
∂ ü0z (x, y, t)
üx (x, y, z, t) = ü0x (x, y, t) − z (521)
∂x
thereby
Z h/2
∂ ü0z
ρüx dz = m0z ü0x − m1z (522)
−h/2 ∂x
where m0z and m1z are the mass and the first mass moment respect to the plate
middle surface, respectively
Z h/2 Z h/2
0 1
mz = ρdz ; mz = ρzdz (523)
−h/2 −h/2

both being defined per unit plate surface. Note zG = m1z /m0z represents the z-
location of the center of mass G of the differential plate element. For ρ constant or
symmetrical with respect to the middle plane, then m1z = 0 and zG 1 = 0. Finally,

the equilibrium equations for the plate for both x and y−directions are

∂Nx ∂Nxy 0

 + + f − m 0 ü0 + m1 ∂ üz = 0

 ∂x x z x z
∂y ∂x
(524)
 ∂Nxy
 ∂Ny 0 0 1 ∂ ü0z

 + + fy − mz üy + mz =0
∂x ∂y ∂y
In order to obtain the equilibrium equation in z−direction including second-order
effects, consider Figure 65a. The slope of the normal stress resultant Nx at a given
point is given by the slope of the middle surface ∂u0z /∂x. However, the slope of the
stress resultant Nx + (∂Nx /∂x) dx is given by ∂u0z /∂x + ∂ 2 u0z /∂x2 dx, the second
term due to the variation of the slope along dx. Then, the projection of these forces
per unit dy in the z−direction is
  0 
∂u0z ∂Nx ∂uz ∂ 2 u0z ∂ 2 u0z ∂Nx ∂u0z
−Nx + Nx + dx + dx ≃ N x dx + dx
∂x ∂x ∂x ∂x2 ∂x2 ∂x ∂x
 
∂ ∂u0z
= Nx dx (525)
∂x ∂x

- 170-
4.3 Governing equations of a laminated plate

Figure 65: Second order effects from “in-plane” stress resultants. a) Normal stress
resultants. b) Shear stress resultants.

where we have neglected terms in (dx)2 . For the case of Ny , we proceed in a similar
way to obtain its force contribution per unit dx in the z−direction
  0 
∂u0z ∂Ny ∂uz ∂ 2 u0z ∂ 2 u0z ∂Ny ∂u0z
−Ny + Ny + dy + dy ≃ Ny dy + dy
∂y ∂y ∂y ∂y 2 ∂y 2 ∂y ∂y
 
∂ ∂u0
= Ny z dy (526)
∂y ∂y
The vertical force component due to Nxy is obtained from Figure 65b. In this case
the Nxy acting along dx has a slope given by ∂u0z /∂x. On the face which normal
points towards positive y, the stress resultant acting along dx is Nxy +(∂Nxy /∂y) dy.
The slope of this stress resultant is ∂u0z /∂x + ∂ 2 u0z / (∂x∂y) dy, being the last addend
the change of slope when traveling along y−direction. Then, the force per unit dx
in z−direction, due to this pair of shear resultants is
  0   
∂u0z ∂Nxy ∂uz ∂ 2 u0z ∂ ∂u0z
−Nxy + Nxy + dy + dy ≃ Nxy dy (527)
∂x ∂y ∂x ∂x∂y ∂y ∂x

where we have again neglected terms in (dy)2 . Taking into account also the other
pair of Nxy acting along dy, the vertical force per unit dy is
  0   
∂u0z ∂Nxy ∂uz ∂ 2 u0z ∂ ∂u0z
−Nxy + Nxy + dx + dx ≃ Nxy dx (528)
∂y ∂x ∂y ∂x∂y ∂x ∂y
Then, the total vertical force due to “in-plane” normal and shear stress resultants
on surface dxdy are
   
∂ ∂u0z ∂u0z ∂ ∂u0z ∂u0z
fN dxdy = Nx + Nxy dxdy + Ny + Nxy dxdy (529)
∂x ∂x ∂y ∂y ∂y ∂x
These forces can be added to the rest of the vertical ones to establish equilibrium
in the z−direction (see Figures 64-b.3 and 64-c.3)
   
∂Qxz ∂Qyz
dx dy + dy dx + fz dxdy + fN dxdy + fzd dxdy = 0 (530)
∂x ∂y

- 171-
4 Mechanical analysis of composite laminates

where fzd = −m0z ü0z is the only non-vanishing dynamic inertial force in z−direction
and Z h/2
fz (x, y) = bz (x, y, z) dz (531)
−h/2

accounts for the applied load in z−direction per unit plate surface. Dividing by
dxdy
∂Qxz ∂Qyz
+ + (fz + fN + fzd ) = 0 (532)
∂x ∂y
which is the equilibrium equation in z−direction. The term in parenthesis is just an
equivalent vertical load per unit plate surface.
Equilibrium of angular moments may be obtained from Figures 64-c.2 and 64-c.3.
Taking moments in x−direction:

∂My ∂Mxy
− dydx − dxdy + Qyz dxdy + myd dxdy = 0 (533)
∂y ∂x

where third-order terms have been neglected. The term myd dxdy is the moment of
the y−direction dynamic inertia forces. For very small thickness this term may also
be neglected. Otherwise we need to include them. The inertia per unit volume in
y−direction is  
0 ∂ ü0z
−ρüy = −ρ üy − z (534)
∂y
which gives the moment in x−direction per unit surface (we integrate it throughout
the thickness)
Z h/2
∂ ü0
myd = − z (−ρüy ) dz = m1z ü0y − m2z z (535)
−h/2 ∂y
where Z h/2
m2z = ρz 2 dz (536)
−h/2

is the second mass moment or moment of inertia respect to the middle plane. Fur-
thermore, (rather unusual) nonuniform body loads bx and by may also produce the
same moments, but we will neglect them here.
Thus, from Eq. (533)

∂Mxy ∂My ∂ ü0


Qyz = + − m1z ü0y + m2z z (537)
∂x ∂y ∂y

Similarly, we obtain

∂Mx ∂Mxy ∂ ü0


Qxz = + − m1z ü0x + m2z z (538)
∂x ∂y ∂x

Then, the equilibrium in z−direction, Equation (532), is expressed in terms of the


internal moments as
∂ 2 Mx ∂ 2 Mxy ∂ 2 My
+ 2 + = −qz (539)
∂x2 ∂x∂y ∂y 2

- 172-
4.3 Governing equations of a laminated plate

where
!  
∂ ü0x ∂ ü0y ∂ 2 ü0z ∂ 2 ü0z
qz = fz + fN + fzd − m1z + + m2z + (540)
∂x ∂y ∂x2 ∂y 2

is the effective load in z−direction including second-order effects due to in-plane


stress resultants and dynamic effects.
Finally we summarize the equilibrium equations for the plate:

∂Nx ∂Nxy


 + = −qx


 ∂x ∂y

 ∂N
xy ∂Ny
+ = −qy (541)

 ∂x ∂y




 ∂ 2 Mx ∂ 2 Mxy ∂ 2 My
 + 2 + = −qz
∂x2 ∂x∂y ∂y 2
where the effective loads are
 0

 q = f − m 0 ü0 + m1 ∂ üz
 x x z x z


 ∂x

 0
0 ü0 + m1 üz∂



 yq = f y − m z y z

 ∂y


∂u0z ∂u0z



∂u0z ∂u0z
 (542)

 q = f + N + N + N + N
 z z x xy y xy


 ∂x ∂x ∂y ∂y ∂y ∂x

 !  
∂ ü0x ∂ ü0y
 2 0 2 0
2 ∂ üz + ∂ üz



 −m 0 ü0 − m1 + + m
z z z z
 ∂x ∂y ∂x2 ∂y 2

The z−equilibrium Equation (541)3 is simply a more handy form of Eq. (532) which,
at the same time, can be derived from Eq. (517)3 assuming σzz ≈ 0 (as it is usually
done in Beam and Thin Plate theories). In summary, Equations (541), written in
terms of internal loads, with second order effects for “in-plane” resultant forces being
included in the z−equilibrium equation and with some underlying hypothesis for the
out-of-plane stress components, are the Strength of Materials approach to Equations
(517).
Alternatively, typical symbolic and index notations for the equilibrium equations
in terms of stress resultants are
(
N · ∇α = −f α + m0z ü0α − m1z ∇α ü0z
 (543)
∇α · M · ∇α = −fz − ∇α · N · ∇α u0z + m0z ü0z + m1z ∇α · ü0α − m2z ∇2α ü0z
where ∇α is the in-plane gradient operator
 T
∂ ∂
[∇α ] = (544)
∂x ∂y

f α and u0α are the plane components


 0  0 T  T
uα = ux u0y , [f α ] = fx fy (545)

- 173-
4 Mechanical analysis of composite laminates

and N and M are the stress resultants in plane-tensor format


   
Nx Nxy Mx Mxy
[N ] = ; [M ] = (546)
Nxy Ny Mxy My

The index notation of Equations (543) is


(
Nαβ,β = −fα + m0z ü0α − m1z ü0z,α
(547)
Mαβ,αβ = −fz − (Nαβ u0z,β ),α + m0z ü0z + m1z ü0α,α − m2z ü0z,αα

where α and β are the in-plane indices and we have applied Einstein’s convention
implying sum on repeated indices.

Exercise 21 Obtain the static equilibrium equations in terms of stress resultants


for the bending case.
Solution: In the case of bending there are no in-plane resultant forces nor in-plane
loads per unit volume, so N = 0, f α = 0. For static equilibrium ü0 = 0. Hence
the first two Eqs. (541) vanish identically and the third equation is
∂ 2 Mx ∂ 2 Mxy ∂ 2 My
2
+2 + = −fz (548)
∂x ∂x∂y ∂y 2
which is known as the bending static equilibrium equation of a plate.

Exercise 22 Obtain the static buckling equilibrium equation of a plate in principal


in-plane loading directions (i.e. in the directions where Nxy = 0) assuming these
directions are constant all over the plate.
Solution: In static cases ü0 = 0. In buckling analysis we assume that in-plane
distributed loads vanish, q α = 0, so with the condition that Nxy (x, y) = 0, the first
two Eqs. (541) are
∂Nx ∂Ny
= 0 and =0 (549)
∂x ∂y
i.e. Nx = N̄x (y) (constant in x) and Ny = N̄y (x) (constant in y) are given by the in-
plane normal loads at the boundaries. Then the equilibrium equation in z−direction
Eq.(541)3 is
∂ 2 Mx ∂ 2 Mxy ∂ 2 My ∂ 2 u0z ∂ 2 u0z
+ 2 + + N̄x + N̄ y = −fz (550)
∂x2 ∂x∂y ∂y 2 ∂x2 ∂y 2
This last equation constitutes the static buckling equilibrium equation of a plate under
normal loadings.

Exercise 23 Obtain the static buckling equilibrium equation of a plate for shear
loading only all over the plate (i.e. when Nx = Ny = 0).
Solution: In static cases ü0 = 0. In buckling analysis we assume that in-plane
distributed loads, q α = 0, so with the condition that Nx = Ny = 0, the first two Eqs.
(541) simplify to
∂Nxy ∂Nxy
= =0 (551)
∂x ∂y

- 174-
4.3 Governing equations of a laminated plate

i.e. Nxy = N̄xy (constant) is given by the shear loading at the boundaries. Then the
equilibrium equation in z−direction, Eq.(541)3 is

∂ 2 Mx ∂ 2 Mxy ∂ 2 My ∂ 2 u0z
+ 2 + + 2N̄xy = −fz (552)
∂x2 ∂x∂y ∂y 2 ∂x∂y

This last equation constitutes the static buckling equilibrium equation of a plate under
shear loading.

Exercise 24 Obtain the dynamic equilibrium equation in z−direction assuming van-


ishing in-plane loads and vanishing in-plane displacements of the middle surface, i.e.
N = 0 and u0α = 0
Solution: In this case the equilibrium Equation (541)3 is
 
∂ 2 Mx ∂ 2 Mxy ∂ 2 My ∂ 2 ü0z ∂ 2 ü0z
+ 2 + = −fz + m0z ü0z − m2z + (553)
∂x2 ∂x∂y ∂y 2 ∂x2 ∂y 2

where the last term of the right-hand-side (the rotary inertial term) is usually ne-
glected for thin plates.

Exercise 25 Defining the stress resultants for each lamina k in a laminate as


Z zk Z zk Z zk
(k) (k)
Nαβ = σαβ dz; Q(k)
α = σαz dz; Mαβ = zσαβ dz (554)
zk−1 zk−1 zk−1

where both α and β can take the values x and y, show that for vanishing body load
of any type we obtain:

(k) (k)

 ∂Nx ∂Nxy (k) (k−1)

 + + σxz − σxz =0


 ∂x ∂y


 (k) (k)
∂Nxy ∂Ny (k) (k−1) (555)
+ + σyz − σyz =0


 ∂x ∂y


 (k) (k)

 ∂Qx ∂Qy (k) (k−1)

 + + σzz − σzz =0
∂x ∂y

(k) (k)

 ∂Mx ∂Mxy (k) (k−1) (k)

 + + zk σxz − zk−1 σxz − Qx = 0
∂x ∂y (556)
(k) (k)

 ∂Mxy ∂My (k) (k−1) (k)

 + + zk σyz − zk−1 σyz − Qy = 0
∂y ∂y
(k)
where σij are the stresses at the top of lamina k.
Solution: In every single point of the laminate the local equilibrium Equations (517)
have to be fulfilled. Hence, for each lamina they also hold. Then, we can integrate

- 175-
4 Mechanical analysis of composite laminates

these equilibrium equations through the thickness of the lamina to yield


 Z zk Z zk Z zk
∂σxx ∂σxy ∂σxz


 dz + dz + dz = 0


 zk−1 ∂x zk−1 ∂y zk−1 ∂z

 Z zk ∂σ
 Z zk
∂σyy
Z zk
∂σyz
yx
dz + dz + dz = 0 (557)

 zk−1 ∂x zk−1 ∂y zk−1 ∂z
 Z
 Z zk Z zk
zk


 ∂σ xz ∂σ yz ∂σzz

 dz + dz + dz = 0
zk−1 ∂x zk−1 ∂y zk−1 ∂z

We now note that, for example


Z zk
∂σxz
dz = σxz |zzkk−1 = σxz
(k) (k−1)
− σxz (558)
zk−1 ∂z

Interchanging integration and differentiation, we readily identify the stress resultants


and obtain Eqs.(555).
The remaining equations in terms of internal moments are obtained multiplying
the first two local equilibrium Equations (517) by z and taking again the integral
throughout the thickness of the lamina
 Z zk Z zk Z zk
 ∂σxx ∂σxy ∂σxz

 z dz + z dz + z dz = 0
 zk−1 ∂x
 zk−1 ∂y zk−1 ∂z
Z zk
∂σyx
Z zk
∂σyy
Z zk
∂σyz (559)

 z dz + z dz + z dz = 0

 zk−1 ∂x
 zk−1 ∂y zk−1 ∂z

Since, for example


Z zk Z zk Z zk
∂σxz ∂ (zσxz )
z dz = dz − σxz dz
zk−1 ∂z zk−1 ∂z zk−1
(k) (k−1)
= zk σxz − zk−1 σxz − Q(k)
x (560)

and
Z zk Z zk (k)
∂σxx ∂ (zσxx ) ∂Mx
z dz = dz = (561)
zk−1 ∂x zk−1 ∂x ∂x

we immediately obtain Eqs. (556).


Note that, since the out-of-plane stresses σxz , σyz and σzz are usually zero at the
top (zN ) and bottom (z0 ) of a N-layer laminate, Eqs. (555) clearly show that the
(k) (k) (k)
interlaminar stresses σxz , σyz and σzz , for 1 ≤ k ≤ (N − 1), are due to in-plane
(k) (k)
variations of the stress resultants Nαβ and Qα ; i.e. these interlaminar stresses
equilibrate those variations. In other words, a homogeneous stress state within each
lamina (with constant resultant forces and moments for each lamina) results into
vanishing interlaminar stresses in the laminate.

- 176-
4.3 Governing equations of a laminated plate

4.3.2 Equilibrium equations in terms of displacements


In order to obtain the equilibrium equation in terms of displacements, we use the
constitutive equation of a laminate, Eq. (478)
 
  ..   0   
{N }  [A] . [B]  ε {N ∗ }
 ···  =  ··· ··· 
 · ·0· −
   ···  (562)

{M } .
. κ {M ∗ }
[B] . [D]

where the asterisk stands for piezo-hygro-thermal loads, if present, and, for example,
{N } is the Voigt notation of the in-plane stress resultants N . Then using symbolic
notation (
N = A : ε0 + B : κ0 − N ∗
(563)
M = B : ε0 + D : κ 0 − M ∗

where, for example, A : ε0 is the symbolic notation of the product [A] ε0 using
Voigt notation. In index notation Eq. (563) is
(
Nαβ = Aαβγδ ε0γδ + Bαβγδ κ0γδ − Nαβ∗
(564)
Mαβ = Bαβγδ ε0γδ + Dαβγδ κ0γδ − Mαβ ∗

where the indices α, β, γ and δ can take in-plane values (x and y) only. Substituting
Eqs. (564) into Eqs.(547)




Aαβγδ ε0γδ,β + Bαβγδ κ0γδ,β = −fα + m0z ü0α − m1z ü0z,α + Nαβ,β

Bαβγδ ε0γδ,αβ + Dαβγδ κ0γδ,αβ = −fz − (Nαβ u0z,β ),α + m0z ü0z + m1z ü0α,α (565)



−m2z ü0z,αα + Mαβ,αβ

Finally, since the middle-plane strains (Eq. (370)) and curvatures (Eq. (371)) in
index notation are

ε0γδ = 12 u0γ,δ + u0δ,γ and κ0γδ = −u0z,γδ (566)

we obtain   
 Aαβγδ 21 u0γ,δβ + 21 u0δ,γβ − Bαβγδ u0z,γδβ = −qα

  (567)
 Bαβγδ 1 u0

+ 1 0 0
2 γ,δαβ 2 δ,γαβ − Dαβγδ uz,γδαβ = −qz
u

where the effective loads


( ∗
qα = fα − m0z ü0α + m1z ü0z,α − Nαβ,β
(568)
qz = fz + (Nαβ u0z,β ),α − m0z ü0z − m1z ü0α,α + m2z ü0z,αα − Mαβ,αβ

include now the effect of the piezo-hygro-thermal loads. Equations (567) are the field
equations (or equilibrium equations in terms of displacements) in index notation for
a general anisotropic plate. We note that the first line of Eqs. (567) is a vector
equation with x and y components, whereas the second line is a scalar equation

- 177-
4 Mechanical analysis of composite laminates

(component z). We also note that we used Einstein’s convention of repeated indices,
so for example
X x,y X
x,y X x,y
x,y X
Dαβγδ u0z,γδαβ ≡ Dαβγδ u0z,γδαβ (569)
α β γ δ

i.e. 16 addends, so the left-hand-side of Eq. (567)2 has 48 addends in total! (some
of them equal). In the next exercise, we write Eqs. (567) in expanded format. We
will consider these equations for some special cases which simplify Eqs. (567) in a
considerable way.

Exercise 26 Particularize the Field Equations (567) for a generally anisotropic


laminated plate in Voigt notation accounting for symmetries
Solution: Collapse the indices of the structural tensors according to Eq. (51) and
the corresponding symmetries (for example: A1 4 = Axx xy = Axx yx = Axy xx =
Ayx xx ). We leave this exercise to the reader and simply quote the result.

∂ 2 u0x ∂ 2 u0x ∂ 2 u0x ∂ 2 u0y ∂ 2 u0y ∂ 2 u0y


A11 + 2A 14 + A44 + A14 + (A12 + A 44 ) + A24 ...
∂x2 ∂x∂y ∂y 2 ∂x2 ∂x∂y ∂y 2
∂ 3 u0z ∂ 3 u0 ∂ 3 u0z ∂ 3 u0z
−B11 3
− 3B14 2 z − (B12 + 2B44 ) 2
− B24 = −qx (570)
∂x ∂x ∂y ∂x∂y ∂y 3

∂ 2 u0x ∂ 2 u0x ∂ 2 u0x ∂ 2 u0y ∂ 2 u0y ∂ 2 u0y


A14 + (A 12 + A 44 ) + A24 + A 44 + 2A 24 + A22 ...
∂x2 ∂x∂y ∂y 2 ∂x2 ∂x∂y ∂y 2
∂ 3 u0z ∂ 3 u0z ∂ 3 u0z ∂ 3 u0z
−B14 − (B 12 + 2B 44 ) − 3B 24 − B 22 = −qy (571)
∂x3 ∂x2 ∂y ∂x∂y 2 ∂y 3

∂ 4 u0z ∂ 4 u0 ∂ 4 u0 ∂ 4 u0z ∂ 4 u0z


D11 4
+ 4D14 3 z + 2 (D12 + 2D44 ) 2 z 2 + 4D24 3
+ D22 ...
∂x ∂x ∂y ∂x ∂y ∂x∂y ∂y 4
∂ 3 u0x ∂ 3 u0 ∂ 3 u0x ∂ 3 u0x
−B11 3
− 3B14 2 x − (B12 + 2B44 ) 2
− B24 ...
∂x ∂x ∂y ∂x∂y ∂y 3
∂ 3 u0y ∂ 3 u0y ∂ 3 u0y ∂ 3 u0y
−B22 − 3B 24 − (B 12 + 2B 44 ) − B 14 = qz (572)
∂y 3 ∂x∂y 2 ∂x2 ∂y ∂x3
where  ∗ 
∂ ü0z ∂Nx∗ ∂Nxy
−qx = −fx + m0z ü0x − m1z + + (573)
∂x ∂x ∂y
 ∗ 
∂ ü0 ∂Nxy ∂Ny∗
−qy = −fy + m0z ü0y − m1z z + + (574)
∂y ∂x ∂y
   
∂ ∂u0z ∂u0 ∂ ∂u0 ∂u0
qz = fz + Nx + Nxy z + Nxy z + Ny z
∂x ∂x ∂y ∂y ∂x ∂y
!  
∂ ü0x ∂ ü0y ∂ 2 ü0z ∂ 2 ü0z
− m0z ü0z − m1z + + m2z +
∂x ∂y ∂x2 ∂y 2
!
∂ 2 Mx∗ ∂ 2 Mxy ∗ ∂ 2 My∗
− + 2 + (575)
∂x2 ∂x∂y ∂y 2

- 178-
4.3 Governing equations of a laminated plate

Exercise 27 Obtain the field equations for a symmetric orthotropic laminate (i.e.
what is frequently named a “specially orthotropic” plate, or simply an “orthotropic”
plate).
Solution: For symmetric laminates [B] = [0]. For symmetric cross-ply laminates
A14 = A24 = D14 = D24 = 0. Using tensor index notation and accounting for
symmetries
Axxxy = Axxyx = Axyxx = Ayxxx = Ayyxy = Ayyyx = Ayxyy = Axyyy = 0 (576)
Dxxxy = Dxxyx = Dxyxx = Dyxxx = Dyyxy = Dyyyx = Dyxyy = Dxyyy = 0 (577)
The terms different from zero are
A11 , A22 , A12 , A44 and D11 , D22 , D12 , D44 (578)
i.e. in tensor index notation
Axxxx , Ayyyy , (Axxyy , Ayyxx ) , (Axyxy , Axyyx , Ayxxy , Ayxyx ) (579)
and
Dxxxx , Dyyyy , (Dxxyy , Dyyxx ) , (Dxyxy , Dxyyx , Dyxxy , Dyxyx ) (580)
Then Eq. (567)1 for α = x has only four terms in the left-hand side:
 
Axxxx u0x,xx + Axxyy u0y,yx + Axyxy 21 u0x,yy + u0y,xy + Axyyx 21 u0y,xy + u0x,yy = −qx (581)

In Voigt notation, accounting for symmetries of A, this last equation results into

A11 u0x,xx + A12 u0y,yx + A44 u0x,yy + u0y,xy = −qx (582)
i.e.
∂ 2 u0x ∂ 2 u0y ∂ 2 u0x
A11 + (A12 + A44 ) + A44 = −qx (583)
∂x2 ∂x∂y ∂y 2
Alternatively, we could have started from Eq. (570) taking only the terms given in
Eq. (578).
In a similar way for α = y
∂ 2 u0y ∂ 2 u0x ∂ 2 u0y
A22 + (A12 + A44 ) + A44 = −qy (584)
∂y 2 ∂x∂y ∂x2
On the other hand, for the z−direction
Dαβγδ u0z,δγαβ = qz (585)
Taking only the nonvanishing terms given in (580), it results in
Dxxxx u0z,xxxx + 2Dxxyy u0z,xxyy + 4Dxyxy u0z,xyxy + Dyyyy u0z,yyyy = qz (586)
which in usual Voigt notation is
∂ 4 u0z ∂ 4 u0z ∂ 4 u0z
D11 + 2 (D12 + 2D44 ) + D22 = qz (587)
∂x4 ∂y 2 ∂x2 ∂y 4
(compare with Eq. (572)). In these equations, the effective loads are given by Eqs.
(573) to (575). Note that the out-of-plane behavior, Eq.(587), is uncoupled from the
in-plane behavior, Eqs. (583) and (584).

- 179-
4 Mechanical analysis of composite laminates

Exercise 28 Obtain the z−field equation for an isotropic, homogeneous plate under
bending loads.
Solution: For a single-layered isotropic laminate, Eq. (417), we have

1−ν
D11 = D22 = D D12 = νD D44 = D (588)
2
where
Eh3
D= (589)
12(1 − ν 2 )
Then Eq.(587) reduces to

∂ 4 u0z ∂ 4 u0z ∂ 4 u0z qz


4
+ 2 2 2
+ 4
= (590)
∂x ∂y ∂x ∂y D

which is the well-known form of the bending equation of an isotropic plate. Using
symbolic notation, it reads
qz
∇2 ∇2 u0z = (591)
D
where qz and u0z have the same direction for positive values.

Exercise 29 Obtain the field equations for a symmetric angle-ply laminate.


Solution: For symmetric angle-ply laminates, [B] = [0], but all terms in A and D
are nonzero in general (see Section 3.3.2). Then we can use Eqs. (570) to (572) to
arrive at
∂ 2 u0x ∂ 2 u0x ∂ 2 u0x ∂ 2 u0y ∂ 2 u0y ∂ 2 u0y
A11 + 2A14 + A44 + A14 + (A12 + A44 ) + A24 = −qx (592)
∂x2 ∂x∂y ∂y 2 ∂x2 ∂x∂y ∂y 2

∂ 2 u0x ∂ 2 u0x ∂ 2 u0x ∂ 2 u0y ∂ 2 u0y ∂ 2 u0y


A14 + (A12 + A44 ) + A24 + A44 + 2A24 + A22 = −qy (593)
∂x2 ∂x∂y ∂y 2 ∂x2 ∂x∂y ∂y 2
∂ 4 u0z ∂ 4 u0z ∂ 4 u0z ∂ 4 u0z ∂ 4 u0z
D11 + 4D 14 + 2 (D 12 + 2D 44 ) + 4D 24 + D 22 = qz (594)
∂x4 ∂x3 ∂y ∂x2 ∂y 2 ∂x∂y 3 ∂y 4
We note that in the case of many plies, the ratio between the ply and the laminate
thicknesses t/h is small and the ratio between coupling and non-coupling coefficients
is of that order, i.e. A14 /A11 = O (t/h), D14 /D11 = O (t/h), see Equations (425)
and (426). Then, with a sufficient number of plies, the resulting field equations tend
to adopt the form of those of Exercise 27 and the laminate approximately behaves as
a specially orthotropic one.

Exercise 30 Obtain the field equations for an antisymmetric angle-ply laminate.


Solution: It is left to the reader to verify that the field equations for this case are:

∂ 2 u0x ∂ 2 u0x ∂ 2 u0y ∂ 3 u0z ∂ 3 u0z


A11 + A44 + (A12 + A44 ) − 3B 14 − B 24 = −qx (595)
∂x2 ∂y 2 ∂x∂y ∂x2 ∂y ∂y 3

∂ 2 u0x ∂ 2 u0y ∂ 2 u0y ∂ 3 u0z ∂ 3 u0z


(A12 + A44 ) + A44 + A22 − B 14 − 3B 24 = −qy (596)
∂x∂y ∂x2 ∂y 2 ∂x3 ∂x∂y 2

- 180-
4.3 Governing equations of a laminated plate

∂ 4 u0z ∂ 4 u0 ∂ 4 u0z
D11 4
+ 2 (D12 + 2D44 ) 2 z 2 + D22 ...
∂x ∂x ∂y ∂y 4
! !
∂ 3 u0x ∂ 3 u0y ∂ 3 u0x ∂ 3 u0y
−B14 3 2 + − B24 +3 = qz (597)
∂x ∂y ∂x3 ∂y 3 ∂x∂y 2
Recall that for this type of laminates the coupling terms B14 and B24 also approach
to zero when the number of plies is increased for fixed h.
Exercise 31 Obtain the expression of the stress resultants in terms of the displace-
ment field and simplify the expressions for the case of isotropic plates under static
equilibrium.
Solution: From Eqs. (564), substituting Eqs. (566) we arrive at
  
 Nαβ = 12 Aαβγδ u0γ,δ + u0δ,γ − Bαβγδ u0z,γδ − Nαβ
 ∗

  (598)
 Mαβ = 1 Bαβγδ u0 + u0 − Dαβγδ u0 − M ∗

2 γ,δ δ,γ z,γδ αβ

For the case of isotropic plates, [B] = [0], and the other coefficients are given by
Eqs. (588)–(589) along with
1−ν
A11 = A22 = A A12 = νA A44 = A (599)
2
and
Eh
A= (600)
1 − ν2
whereupon  !
0 ∂u 0

 ∂u x y
 N =A +ν − Nx∗
 x


 ∂x ∂y



 !
 ∂u0y ∂u0x
Ny = A +ν − Ny∗ (601)

 ∂y ∂x



 !


 1 − ν ∂u 0
x ∂u0y ∗
 Nxy = A 2 + − Nxy


∂y ∂x
  2 0 
 ∂ uz ∂ 2 u0z

 Mx = −D 2
+ν 2
− Mx∗


 ∂x ∂y

  2 0 
∂ 2 u0z

∂ uz
My = −D 2
+ν 2
− My∗ (602)

 ∂y ∂x


∂ 2 u0z


 ∗
 Mxy = −D (1 − ν)
 − Mxy
∂x∂y
Finally from Eqs. (537) and (538) and considering static equilibrium
  3 0  ∗
 ∂ uz ∂ 3 u0z Mx∗ Mxy
 Qxz = −D
 + 2 − −
 ∂x3 ∂y ∂x ∂x ∂y
 3 0  ∗
(603)

 ∂ uz ∂ 3 u0z Mxy My∗
 Qyz = −D
 + 2 − −
∂y 3 ∂x ∂y ∂x ∂y
We note that the in-plane behavior is fully uncoupled from the out-of-plane behavior.

- 181-
4 Mechanical analysis of composite laminates

4.3.3 Matrix-operator form of the field equation


The field equations of the anisotropic laminate may be written in a handy familiar
matrix operator format, specially useful for the Navier solution shown below. The
matrix operator format of the general equation given in Exercise 26 is
  0    0   
K11 K12 K13  ux M11 0 M13  üx  −(fx + Fx )

 
  
  

u0y ü0y
   
 K12 K22 K23  + 0 M22 M23  = −(fy + Fy∗ )

 0 
 
 0 
   

K13 K23 K33 uz M13 M23 M33 üz fz + Fz∗
(604)

where


 K11 = A11 ∂xx + 2A14 ∂xy + A44 ∂yy




 K12 = A14 ∂xx + (A12 + A44 ) ∂xy + A24 ∂yy


 K13 = − [B11 ∂xxx + 3B14 ∂xxy + (B12 + 2B44 ) ∂xyy + B24 ∂yyy ]


K22 = A44 ∂xx + 2A24 ∂xy + A22 ∂yy (605)


 K23

 = − [B14 ∂xxx + (B12 + 2B44 ) ∂xxy + 3B24 ∂xyy + B22 ∂yyy ]




 K33 = D11 ∂xxxx + 4D14 ∂xxxy + 2 (D12 + 2D44 ) ∂xxyy + 4D24 ∂xyyy


+D22 ∂yyyy − (Nx ∂xx + 2Nxy ∂xy + Ny ∂yy )


 M11 = −m0z

 0
 M22 = −mz


M13 = m1z ∂x (606)

M23 = m1z ∂y





M33 = m0z − m2z (∂xx + ∂yy )

and  ∗ 

 Fx = − ∂Nx∗ /∂x + ∂Nxy ∗ /∂y
 
Fy∗ = − ∂Nxy∗ /∂x + ∂N ∗ /∂y
y (607)

 
 ∗
Fz = − ∂ 2 Mx∗ /∂x2 + 2∂ 2 Mxy
∗ /∂x∂y + ∂ 2 M ∗ /∂y 2
y

The differential operators in compact form are


∂ (·) ∂ (·) ∂ 2 (·)
∂x = , ∂y (·) = , ∂xy (·) = , ... (608)
∂x ∂y ∂x∂y
Note that the resultant forces Nx , Ny and Nxy have been assumed as in exercises
22 and 23, so the corresponding spatial derivatives are not included in the second-
order contribution to K33 . This will be useful below, within the Section devoted to
bucking analysis.

4.3.4 Boundary conditions for the field equations


Essential boundary conditions
The field equations of the laminate, i.e. Eqs. (567) are second order differential
equations in terms of in-plane displacement components (u0x and u0y ) and fourth order

- 182-
4.3 Governing equations of a laminated plate

differential equations in terms of the out-of-plane displacement u0z . This implies


that we need one boundary condition for each in-plane displacement (for example
prescribed u0x and u0y values at the boundaries) and two boundary conditions for the
out-of-plane component10 (for example prescribed u0z and ∂u0z /∂x values established
along a boundary parallel to the y−direction). Since these boundary conditions have
to be prescribed along the boundary Γ, which may have arbitrary shape, see Figure
66, one possible set is
 0
 un = ū0n on Γ


 u0 = ū0 on Γ
s s
0 = ū0 on Γ
(609)


 u z z

∂u0z /∂n = −ϕ̄s on Γ
where ū0n , ū0s and ū0z are prescribed displacements at the boundaries and ϕ̄s is
a prescribed rotation at the boundaries (for example they may take zero values
ū0n = ū0s = ū0z = ϕ̄s = 0). The coordinate n is normal to the surface Γ of the
boundary and s is the coordinate along the boundary as shown in Figure 66. The
displacements u0n and u0s may be obtained from u0x , u0y using the usual change of
system of representation (c.f. Eq. (959), Appendix B)
( ) " #( )
u0n n̂ · ex n̂ · ey u0x
= (610)
u0s ŝ · ex ŝ · ey u0y

where n̂ is the unit vector perpendicular to the surface and ŝ is the unit vector
tangent to the surface, see Figure 66. The same transformation rule applies to the
derivatives (i.e. rotation angles) ∂u0z /∂n and ∂u0z /∂s, which are obtained through
the gradient vector transformation
   
 ∂u0z  " # ∂u0z 
   

∂n
 n̂ · ex n̂ · ey  ∂x 
0 
= 0 
(611)
 ∂uz  ŝ · ex ŝ · ey  ∂uz 

 
 
∂s ∂y

Note that prescribing for example ū0z along the whole boundary implies at the same
time that
u0z = ū0z on Γ ⇒ ∂u0z /∂s = ϕ̄n on Γ (612)
Hence, the displacement u0z and the rotation ∂u0z /∂s are directly related at bound-
aries and they cannot be independently prescribed.
Kinematical boundary conditions of the type of Eq. (609) (certainly, they are just
one choice) are named essential boundary conditions. In this case Eqs. (609) corre-
spond to fully clamped boundaries (not necessarily with vanishing values). However,
it is possible to prescribe work-conjugate boundary loads instead of displacements.
10
For example, a second order ordinary differential equation could be d2 u/dx2 = 2a for x ∈ [0, 1].
Then by integration u (x) = ax2 + bx + c, where a is given. Two boundary conditions are needed
to obtain b and c (for example u (0) = c and u (1) = a + b + c); i.e. one single boundary condition
established at the boundaries u (x̄) = ū (x̄), x̄ = 0, 1, where the overbar implies value at the
boundary.

- 183-
4 Mechanical analysis of composite laminates

Figure 66: Boundary conditions on a plate. Positive senses are drawn for all the
variables being shown.

Natural boundary conditions


Load-type boundary conditions are usually named natural boundary conditions. In
this case, one possible (partial) set of boundary conditions is given by
 0 0
 Nn = N̄n on Γ

Nns0 = N̄ 0 on Γ
ns (613)

 0 0
Mn = M̄n on Γ
These prescribed values may be regarded the load-type counterpart of the first,
second and fourth kinematical boundary conditions given in Eq. (609).
However a note is needed regarding the remaining natural boundary condition
in the set of Eq. (613), i.e. the work-conjugate of u0z . The work performed by the
(out-of-plane) boundary conditions is
Z

WΓ = Qn u0z + Mn ϕs − Mns ϕn ds (614)
Γ

where the minus sign on the last term is due to the sign criteria used and shown in
Figure 66. Substituting the rotations by their values in terms of the displacements
u0z and integrating the last addend by parts along a boundary line sA → sB , where
A and B are two points on Γ, we obtain
Z  
∂u0 ∂u0
WΓ = Qn u0z − Mn z − Mns z ds
Γ ∂n ∂s
Z  0

0 ∂uz ∂Mns 0  B
= Qn uz − Mn + uz ds − Mns u0z A
Γ ∂n ∂s
Z  
∂Mns 0 ∂u0z  B
= (Qn + )uz − Mn − Mns u0z A (615)
Γ ∂s ∂n

- 184-
4.3 Governing equations of a laminated plate

Obviously if A and B are the same point (i.e. we travelled all the boundary Γ closing
the loop) the term in brackets must vanish and we have that the work-conjugate of
u0z is a new quantity Vn defined as
∂Mns
Vn = Q n + (616)
∂s
which receive the name of Kirchhoff plate (boundary shear) equivalent force. This
means that the Neumann substitute of u0z in a boundary condition is Vn = V̄n , where
V̄n is the actual shear force being prescribed at the boundary (both the shear stress-
resultant and the twist moment stress-resultant variation contribute to the vertical
shear force). Furthermore, since Qn and ∂Mns /∂s are directly related by means of
an angular moments equilibrium equation of the type of Eq. (538) (simply change
x by n and y by s and consider the static case), one can infer something similar
to what is stated in Eq. (612); i.e. that Qn , ∂Mns /∂s and Vn are not independent
variables at the boundaries. Accordingly, only one of them can take part in the
selected set of boundary conditions.
Finally, the components of force and moment stress-resultants in an arbitrary
plane (see Figure 67) may be readily obtained from the values Nx , Ny , Nxy and Mx ,
My , Mxy . It is apparent from Eqs. (510) and (511) that the procedure to obtain such
through-the-thickness components is exactly the same as that to obtain the normal
σnn , σss and shear τns stresses on that plane from components σxx , σyy and σxy ,
i.e. they change of system of representation as plane tensors. Thus, from Appendix
B, the in-plane resultant forces per unit width change of system of representation
through
" # " #" #" #T
N n N ns n̂ · ex n̂ · ey N x N xy n̂ · ex n̂ · ey
= (617)
N ns N s ŝ · ex ŝ · ey N xy N y ŝ · ex ŝ · ey

The same transformation rule is to be used for the resultant moments. The 2D Mohr
circle is also directly applicable as shown in Figure 68. We note that there are two
criteria when using Mohr circles; Figure 68 shows one of them. Also note that arrow
directions for stresses and resultant forces are distinct from those of the associated
moments (i.e. although both Nn and Mn are obtained from σnn , the Nn −arrow is
parallel to the direction of σnn and the Mn −arrow is perpendicular to the direction
of σnn just because Nn represents a force and Mn is a moment, see Figures 66 to
68).

Kirchhoff plate equivalent forces


The bracket in Eq. (615) vanishes between two infinitesimally nearby points located
in smooth boundaries as the one shown in Figure 66. In these boundaries the normals
in two infinitesimally nearby points A and B are coincident and so the Mns values.
However, when the boundary has a corner, the normals at the nearby points A and
B, which are on different planes, are also different, see Figure 68. Then at the
corner, the bracket takes the value
 B  0
Mns u0z A = Mns
B A
− Mns uz = RAB u0z (618)

- 185-
4 Mechanical analysis of composite laminates

Figure 67: Stress resultants in an arbitrary transversal plane n̂ of a plate. Positive


senses are drawn.

Figure 68: Mohr circles for plate stress resultants.

- 186-
4.3 Governing equations of a laminated plate

y
a)
P

P
b)

My , Mxy RAB < 0


Mxy > 0

B
Mns < 0
A
Mx , Mxy Mns > 0

Figure 69: a) Rectangular plate, with contact restriction at boundaries, being loaded
downwards. b) Same plate, but with the vertical displacement completely restricted
at boundaries (simply supported); equivalent interpretations of the Kirchhoff reac-
tion force at the shown corner.

so the boundary energy Eq. (615) implies that there is a concentrated force at the
corner with value RAB = Mns B − M A . Note that the dimensions of M
ns ns are those of
moment per unit length, i.e. those of force (N m / m = N), and that Mns > 0 if it
has the sense indicated in Figures 66 and 67.
For example, for the simply supported, loaded downwards rectangular plate of
Figure 69b, one observes that the values taken by Mns at both sides of the depicted
right-angled corner are the same but have opposite signs (simply take α = 90o in
Figure 68). Since Mxy > 0 at that corner (recall Figures 44 or 64-b.2), the force at
that corner is
B A
RAB = Mns − Mns = −Mxy − Mxy = −2Mxy < 0 (619)
|{z} |{z} | {z } |{z}
<0 >0 <0 >0

i.e. it is negative, so has the sense of negative u0z . The same occurs for the other
three corners. These are the well known Kirchhoff corner loads of a plate. They are
concentrated reaction forces acting at the corners so that the boundary conditions
are fulfilled; i.e. they prevent the corners raising up (note that if positive vertical
displacements were not restricted throughout the boundary, then the corners tend

- 187-
4 Mechanical analysis of composite laminates

to raise up, see Figure 69a). A possible physical interpretation of these corner loads
is to consider moments as force-couples as shown also in Figure 69b. At smooth
surfaces (like the lateral faces of a rectangular plate, for example) adjacent couples
cancel each other, but at corners the couples are not cancelled-out, originating the
corner forces. Finally, note that the Kirchhoff vertical reaction forces Vn acting at
the lateral right ends of the plate are not shown in Figure 69b; obviously they will
have the opposite sense of both the external load P and the four reaction corner
forces R, so that the plate will be under global static equilibrium.

Typical sets of boundary conditions


Possible boundary conditions are combinations of {u0n , u0s , u0z , ∂u0z /∂n} and {Nn ,
Nns , Vn , Mn }. When u0z = ū0z = 0 and ∂u0z /∂n = −ϕ̄s = 0, the plate is considered
as clamped. If u0z = ū0z = 0 and Mn = M̄n = 0, the plate is considered simply
supported. However note that there are several combinations for both cases; typical
ones follow:


 C1: u0z = 0 ϕs = 0 u0n = ū0n u0s = ū0s
 0
C2: uz = 0 ϕs = 0 Nn = N̄n u0s = ū0s
0 0 0 (620)

 C3: uz = 0 ϕs = 0 un = ūn Nns = N̄ns

C4: u0z = 0 ϕs = 0 Nn = N̄n Nns = N̄ns


 S1: u0z = 0 Mn = 0 u0n = ū0n u0s = ū0s
 0
S2: uz = 0 Mn = 0 Nn = N̄n u0s = ū0s
(621)

 S3: u0z = 0 Mn = 0 u0n = ū0n Nns = N̄ns

S4: u0z = 0 Mn = 0 Nn = N̄n Nns = N̄ns

4.3.5 Initial conditions for the field equations


For the case of initial conditions, we note that the time dependence is present in
Eqs. (568) which constitute the right-hand-sides of Equations (567). Since the
differential equation is of second order in time, two initial conditions are needed for
each displacement component. For example it is typical to prescribe
( 0
ux (t = 0) = u00x , u0y (t = 0) = u00y , u0z (t = 0) = u00z
(622)
0 , u̇0 (t = 0) = v 0 , u̇0 (t = 0) = v 0
u̇0x (t = 0) = v0x y 0y z 0z

where u00x , u00y , u00z are initial (at t = 0) prescribed displacements in the middle
surface of the plate and v0x 0 , v 0 , v 0 are initial prescribed velocities in the middle
0y 0z
surface of the plate.

4.3.6 The Airy stress function


The analytical solution of the partial differential equations for plane elasticity prob-
lems has been frequently obtained using the Beltrami-Michell form and the Airy
stress function. The use of the Airy stress function is specially interesting for the
case of symmetric specially orthotropic laminates because the resulting in-plane and
out-of-plane differential equations result to be the same, and hence are the solutions.

- 188-
4.3 Governing equations of a laminated plate

The Airy stress function is a scalar function φ (x, y) to be determined but such
that
∂2φ ∂2φ ∂2φ
σxx = + V ; σ yy = + V ; σ xy = − (623)
∂y 2 ∂x2 ∂x∂y

where V (x, y) is another function such that the in-plane body loads are bα = −∇α V ,
i.e. bx = −∂V /∂x and by = −∂V /∂y. Then, it is immediate (and left to the reader)
to verify that the Local Equilibrium Equations (517) are automatically fulfilled for
the in-plane, static case. This is basically the point of using the Airy stress function,
because if equilibrium is automatically fulfilled, we simply have to worry about
compatibility, which in the thin plate case is

∂ 2 ε0xx ∂ 2 ε0yy ∂ 2 γxy


0
+ = (624)
∂y 2 ∂x2 ∂x∂y

For the case of stress resultants, the observation of Eqs. (541) prompts us to get

∂2Φ ∂2Φ ∂2Φ


Nx = + Υ ; N x = +Υ ; Nxy = − (625)
∂y 2 ∂x2 ∂x∂y

where
Z h/2 Z h/2
Φ= φ dz = hφ ; Υ = V dz = hV (626)
−h/2 −h/2

and
f α = −∇α Υ (627)

Then, the in-plane equilibrium equations are automatically fulfilled for any Airy
stress-resultant function Φ. The definite solution of the in-plane behavior will be
given by the specific function Φ that makes Eq. (624) to be satisfied.
The strains and the moment stress resultants can be written in terms of the
curvatures and the force stress resultants using Eqs. (563)
(
ε0 = A−1 : N − A−1 : B : κ0
 (628)
M = B : A−1 : N + D − B : A−1 : B : κ0

where for simplicity we do not consider here piezo-hygro-thermal loads. These equa-
tions may be re-written in compact form as
(
ε0 = A # : N + B # : κ 0
(629)
M = C# : N + D# : κ0

where the definitions of A# , B# , C# and D# are apparent from comparison of Eqs.


(628) and (629). Note that B# and C# are not symmetric. Then, substituting Eqs.

- 189-
4 Mechanical analysis of composite laminates

(625) into the first of Eqs. (629)


 
 ∂2Φ 

 +Υ 

 0   
 ∂y 2 

 εx
  A# A# A# 
 

 11 12 14
  
∂2Φ
ε0y =  A#
12 A#
22
# 
A24  +Υ + ...
   ∂x2 
 0
γxy
 A#
14 A#
24 A#
44








 ∂2Φ 


 − 
∂y∂x
  0

# # # κx 
B11 B12 B14 
 
 # # # 
+  B21 B22 B24  κ0y (630)
# # #  
B41 B42 B44  0
κxy

Now Eq. (624) upon substitution of the strains of Eq. (630) gives a partial differen-
tial equation in terms of the Airy stress function. We leave this task to the reader as
#
an exercise, but simply quote the result for symmetric (Bij = 0 = Bij ; A# −1
ij = Aij ;
#
Dij = Dij ) specially orthotropic laminates (A# # # #
14 = A24 = 0; D14 = D24 = 0) under

constant body loads ∇α · f α = −∇2 Υ = 0 :
!
#∂ Φ
4
# A# ∂4Φ #∂ Φ
4
A22 4 + 2 A12 + 44 + A11 =0 (631)
∂x 2 ∂x2 ∂y 2 ∂y 4

This Equation is to be compared to Eq. (587) for the same static case which we
reproduce here for the reader comfort:

∂ 4 u0z ∂ 4 u0z ∂ 4 u0z


D11 + 2 (D12 + 2D44 ) + D22 = fz (632)
∂x4 ∂x2 ∂y 2 ∂y 4
As it can be observed, both equations have each a single scalar unknown field (Φ and
u0z ) and are identical except for the value of the different coefficients. Then, the so-
lution for both in-plane and out-of-plane equations have the same form, one in terms
of Φ and the other in terms of u0z . This is an important fact because this means that
all the procedures and solutions employed for orthotropic bidimensional elasticity
may be employed to obtain the solutions of this type of symmetric laminates.
Finally we note that once the Airy stress function is obtained solving Eq. (631),
the force stress-resultants N are directly obtained from Eqs. (625). Then, after
solving Eq. (632), the curvatures are obtained through κ0αβ = −∂ 2 u0z /∂α∂β and
the middle surface strains ε0 can be obtained using Eq. (629)1 . Thus, strains and
stresses are known throughout the plate for the given loads f and the static problem
is solved.

4.3.7 The energy approach


Another approach to solve continuum mechanics problems is the energy one. In
this approach energy theorems are used. Usually ad-hoc solution forms fulfilling
the essential boundary conditions are proposed which include some coefficients to
be determined (i.e. degrees of freedom). Then an energy principle is claimed in

- 190-
4.3 Governing equations of a laminated plate

order to obtain the unknown coefficients. The Rayleigh method and the finite ele-
ment method are two examples which simply use different approximations for the
displacement field. Of course in these cases, the solution is only approximate. The
accuracy of the solution depends on the selected functions and the number of degrees
of freedom allowed.
We state below the Principle of Minimum Potential Energy and the Principle of
Virtual Work, which in linear elasticity yield identical results.

Principle of Minimum Potential Energy


Recall that the total (elastic) stored energy may be written for linear elasticity as
Z Z
1
W= w dV = 2 σ : ε dV (633)
V V

In symbolic notation the in-plane strains are

ε = ε0 + zκ0 (634)

so
w = 21 σ : ε = 12 σ : ε0 + 12 (zσ) : κ0
which after integrating in the z−direction and using Eqs. (513)1 and (513)2 , may
be written in terms of stress resultants
Z
1 
W= N : ε0 + M : κ0 dS (635)
2 S
Using the constitutive Equations (563)
Z
1 
W= ε0 : A : ε0 + 2ε0 : B : κ0 + κ0 : D : κ0 − N ∗ : ε0 − M ∗ : κ0 dS (636)
2 S
Using the symbolic notations for strains and curvatures

ε0 = ∇sα u0α and κ0 = −∇α ∇α u0z (637)

we obtain the stored energy in terms of displacements


Z
1
W= (∇s u0 : A : ∇sα u0α − 2∇sα u0α : B : ∇α ∇α u0z + ...
2 S α α
+ ∇α ∇α u0z : D : ∇α ∇α u0z − N ∗ : ∇sα u0α + M ∗ : ∇α ∇α u0z ) dS (638)

where ∇sα is the in-plane symmetric gradient operator, so


"  #
1
 0  s 0  ∂u0x /∂x ∂u 0 /∂y + ∂u0 /∂x
x y
ε = ∇ α uα = 1 0 /∂y + ∂u0 /∂x
 2 (639)
2 ∂ux y ∂u0y /∂y

The work of the external loads is obtained as usual


Z Z Z Z
W ext = −V = q · u0 dS = qx u0x dxdy + qy u0y dxdy + qz u0z dxdy (640)
S S S S

- 191-
4 Mechanical analysis of composite laminates

where V is the potential of the external loads, which have been assumed to remain
constant between the undeformed configuration and the actual configuration for each
(x, y) point of the plate.
The Principle of Minimum Potential Energy (PMPE ) for laminated plates states
that
δΠ = δW + δV = 0 (641)

where Π = W + V is the Potential Energy and δ (·) implies an infinitesimal variation.


If {Ξ 1 ,Ξ 2 ,Ξ 3 , ...} is a set of coefficients to be determined (the degrees of freedom),
then the PMPE implies that

∂Π
= 0 for all possible Ξ i . (642)
∂Ξ i

Principle of Virtual Works


The Principle of Virtual Works states that the work of the external loads over any
virtual displacement field is the same as the internal work of the stresses over the vir-
tual strains compatible with (i.e. derived from) the previous virtual displacements.
Mathematically
Z Z
ext 0 1
W = q · δu dS = σ : δε dV = W int (643)
S 2 V

where δu0 are the middle-plane virtual displacements


   T
δu0 = δu0x δu0y δu0z (644)

and δε are the through-the-thickness virtual strains

δε = δε0 + zδκ0 (645)

with
 
δε0 = ∇sα δu0α and δκ0 = −∇α ∇α δu0z (646)

Then Z
ext

W = qx δu0x + qy δu0y + qz δu0z dxdy (647)
S

and Z
int 1 
W = N : δε0 + M : δκ0 dxdy
2 S

yielding, after some little algebra


Z
1
W int = (∇s δu0 : A : ∇sα u0α − ∇sα δu0α : B : ∇α ∇α u0z − ∇α ∇α δu0z : B : ∇sα u0α + ...
2 S α α
+ ∇α ∇α δu0z : D : ∇α ∇α u0z − N ∗ : ∇sα δu0α + M ∗ : ∇α ∇α δu0z )dxdy (648)

- 192-
4.3 Governing equations of a laminated plate

Exercise 32 Obtain the potential energy expression for the out-of-plane behavior of
specially orthotropic plates without piezo-hygrothermal effects.
Solution: For specially orthotropic laminates ([B] = [0]) the in-plane and out-of-
plane behaviors are uncoupled (see Eq. (409)). Hence, the out-of-plane contribution
to W is given by the term that includes the plate curvatures only
Z
1
W= (∇α ∇α u0z : D : ∇α ∇α u0z )dxdy (649)
2 S
which in Voigt notation is
 T  

 ∂ 2 u0z 
 
 ∂ 2 u0z 

   
 ∂x2
 
   ∂x2 

Z 

 ∂ 2 u0

 D11 D12 D14 

 

1 z
 ∂ 2 u0z 
W= 2
 D12 D22 D24  dxdy (650)
2 S ∂y   ∂y 2 





 D14 D24 D44  




2 0 ∂ 2 u0z
 2 ∂ uz

 
 
 

 
 
 2 

∂x∂y ∂x∂y
Since for specially orthotropic plates we also have D14 = D24 = 0, then
Z "  2 0 2  2 0 2  2 0 2 #
1 ∂ uz ∂ 2 u0z ∂ 2 u0z ∂ uz ∂ uz
W= D11 2
+ 2D12 2 ∂y 2
+ D22 2
+ 4D44 dxdy
2 S ∂x ∂x ∂y ∂x∂y

We also have qx = qy = 0 and qz 6= 0, so


Z "  2 0 2
1 ∂ uz ∂ 2 u0z ∂ 2 u0z
Π= D11 + 2D 12
2 S ∂x2 ∂x2 ∂y 2
 2 0 2  2 0 2 #
∂ uz ∂ uz 0
+D22 + 4D44 − 2qz uz dxdy (651)
∂y 2 ∂x∂y

Exercise 33 Particularize the potential energy expression of the preceding exercise


for an isotropic plate.
Solution: For isotropic plates

(1 − ν) Eh3
D11 = D22 = D ; D12 = νD ; D44 = D with D= (652)
2 12 (1 − ν 2 )
so
Z (" 2  2  2 # )
D ∂ 2 u0z ∂ 2 u0z ∂ 2 u0z ∂ 2 u0z ∂ 2 u0z 0
Π= + + 2 (1 − ν) +ν − 2qz uz dxdy
2 S ∂x2 ∂y 2 ∂x∂y ∂x2 ∂y 2

Exercise 34 Obtain the potential energy expression for the case of in-plane behavior
of isotropic plates.
Solution: We leave this exercise to the reader (the solution may be guessed from
the results of the previous ones). We note that this is the potential energy of a 2D
isotropic plate under the plane stress assumption.

- 193-
4 Mechanical analysis of composite laminates

Exercise 35 Obtain the out-of-plane contribution to the Potential Energy of a spe-


cially orthotropic plate when the only non-vanishing term of qz is that of second-order
given by the in-plane stress resultants. Assume constant force stress resultants all
over the plate, simply supported boundary conditions and that the plate is rectangu-
lar with in-plane dimensions a × b.
Solution: Equation (542)3 considering only constant resultant forces N is

∂ 2 u0z ∂ 2 u0z ∂ 2 u0z


q z = Nx + 2N xy + N y (653)
∂x2 ∂x∂y ∂y 2

Taking into account that the vertical displacement varies from u0z = 0 (undeflected
configuration) to u0z (actual deflected configuration), it can be deduced that this
second-order contribution to qz increases linearly with u0z during the plate deflec-
tion. Hence the potential of the “external” loads is obtained through
Z Z 0 !
uz
V=− qz (u0z )du0z dxdy =
S 0
Z  
1 ∂ 2 u0z ∂ 2 u0z ∂ 2 u0z
=− Nx + 2N xy + N y u0z dxdy
S 2 ∂x2 ∂x∂y ∂y 2

Integrating by parts over x, we obtain for example


Z Z  Z a Z  2 !
∂ 2 u0 ∂u0 ∂u0z
z
u0z dx dy = u0z z − dx dy (654)
y x ∂x2 y ∂x 0 x ∂x

Since u0z vanishes at the boundaries (x = 0, x = a, y = 0, y = b), the first term of the
right-hand side of the previous equation vanishes. Then, using similar reasonings,
we arrive at
Z "  0 2  0 0  0 2 #
1 ∂uz ∂uz ∂uz ∂uz
V= Nx + 2Nxy + Ny dxdy (655)
2 S ∂x ∂x ∂y ∂y

and the out-of-plane potential energy results into (c.f. Exercise 32)
Z "  2 0 2  2 0 2  2 0 2
1 ∂ uz ∂ 2 u0z ∂ 2 u0z ∂ uz ∂ uz
Π= D11 2
+ 2D12 2 2
+ D22 2
+ 4D44
2 S ∂x ∂x ∂y ∂y ∂x∂y
 0 2  0 0  0 2 #
∂uz ∂uz ∂uz ∂uz
+ Nx + 2Nxy + Ny dxdy (656)
∂x ∂x ∂y ∂y

4.4 Deformations in composite laminated plates: some simple cases


The analytical solution of the governing equations for a general anisotropic laminated
plate is a extremely difficult task, and very few solutions exist for some simple,
particular cases, usually symmetric laminates with very simple geometries. Here is
where numerical procedures as finite elements are important and the only possible
way to obtain accurate solutions to structures with arbitrary geometries, loads and

- 194-
4.4 Deformations in composite laminated plates: some simple cases

Figure 70: Laminated beam

boundary conditions. However, a good understanding of the analytical solutions


for simple problems is also very important in order to gain insight into the physical
problem and to be able to interpret the numerical results of more complex problems.
In fact, analytical solutions to simple problems are often used to test numerical
procedures and to obtain the mandatory preliminary insight into the problem at
hand.
In this section we will simply study the analytical solution of some simple cases
with the objective to gain insight in the behaviour of anisotropic laminates. For
more comprehensive solutions we refer to the listed References.

4.4.1 Laminated beams


Laminated beams are the simplest problem case. In this case the width b of the
laminated plate (length in y−direction) is much smaller that the length a of the
laminated plate (length in x−direction), see Figure 70. Furthermore, the loads can
be considered only a function of the x coordinate and the boundary conditions are
specified only in the x−direction; i.e. all the problem can be characterized by x
alone, as in the classical beam theory, although variables may also depend on y as
seen below.

Displacements and section stress resultants


We will consider only the case of symmetric laminates, so [B] = [0]. Then, the
constitutive Equations (563) without piezo-hygro-thermal loads reduce to
( 
{N } = [A] ε0
 (657)
{M } = [D] κ0

Since both equations are uncoupled we can consider the bending case first and
then, if necessary, apply superposition (because we are performing linear analyses).
Equation (657)2 can be inverted to obtain
 0
κ = [D]−1 {M } = [D⋄ ] {M } (658)

- 195-
4 Mechanical analysis of composite laminates

where we renamed [D⋄ ] = [D]−1 for notation comfort. In Voigt notation this equation
is  
 ∂ 2 u0z 

 
 

 ∂x2   ⋄ ⋄ ⋄
 M 

 ∂ 2 u0   D 11 D 12 D 14

 x 
z ⋄ ⋄ ⋄
− =  D 12 D 22 D 24
 M y (659)
 ∂y 2  ⋄ ⋄ ⋄  

 2 0 

 D 14 D 24 D 44

Mxy

 2 ∂ uz 

 
 
∂x∂y
In beam theory it is assumed that My = Mxy = 0, so

∂ 2 u0z ⋄ ∂ 2 u0z ⋄ ∂ 2 u0z 1 ⋄


= −D11 Mx ; = −D12 Mx ; = − D14 Mx (660)
∂x2 ∂y 2 ∂x∂y 2
which implies that u0z is a function of both x and y (this last dependence due to the
Poisson and twisting couplings). If the dimension b is small respect to the dimension
a, we can neglect the dependency of u0z on y and consider only the variations on x,
so
d2 u0z ⋄
= −D11 Mx (661)
dx2
where since D⋄ = D−1 ,
⋄ D22 D44 − 2D24
D11 = (662)
det ([D⋄ ])
We note that D11⋄ in Eq. (661) plays the role of 1/(Ē I¯ ) in isotropic beams, where
x y
Ēx is the equivalent Young’s modulus and I¯y = h3 /12 is the beam cross-section
moment of inertia per unit width (i.e. as if b = 1) with respect to the y−axis
(usually written as Iy ). Note also that Mx in Eq. (661) is the flexural moment on
the beam per unit width. This equivalence may be established then as
1 12
Ēx = ⋄ I¯
= ⋄ (663)
D11 y h3 D11
Equation (661) is a very simple differential equation which may be directly in-
tegrated once to obtain the rotation
Z x
du0z (x) ⋄
= −ϕy (x) = − D11 Mx (ξ) dξ + C1 (664)
dx 0

and twice to obtain the displacement


Z x Z x Z η 
u0z (x) = − ϕy (η) dη + C2 = − ⋄
D11 Mx (ξ) dξ dη + C1 x + C2 (665)
0 0 0
| {z }
ϕy (η)

where C1 and C2 are constants to be determined from the boundary conditions. The
moment Mx (ξ) in Eq. (665) may be obtained from the equilibrium Equations (532)
and (539) assuming that My = Qyz = Mxy = 0 and that Mx (x) and Qxz (x) are
functions of x alone:
d2 Mx dQxz
2
= = −qz (666)
dx dx

- 196-
4.4 Deformations in composite laminated plates: some simple cases

Taking qz = fz for the case of static linear analysis and integrating once
Z x
Qxz (x) = − fz (ζ) dζ + C3 (667)
0

and integrating twice


Z x Z x Z τ 
Mx (x) = Qxz (τ ) dτ + C4 = − fz (ζ) dζ dτ + C3 x + C4 (668)
0 0 0

Depending on the boundary conditions it is convenient to use Eqs. (667) and (668)
to obtain first the shear and moment distributions (this procedure is exactly the
same as in isotropic beams) and then to integrate the moment distribution using
Eqs. (664) and (665) to obtain the rotation and the vertical displacement.

Strains and stresses


Strains in a beam can be obtained directly from the displacement field, see Eq.
(372), page 103
 2 0 
   0
ε
 ∂ uz
ε  
 xx  x

 
 
 
 
 

   ∂x2 
εyy = ε0y −z ≃0 (669)

 
 
 
 
 

   0   
γxy X
γxy X

≃0

X

Considering bending loads only (the tensile case may be added by superposition due
to the uncoupling present in Eq. (657)) we obtain
 2 0 
  d uz (x)
ε (x)  
 xx

 
 
 

  dx2 
εyy = −z 0 (670)

 
 
 

   
γxy X

0

X

We emphasize that this is only an approximation for the bending load case. The
transversal strains εyy are certainly distinct from zero due to the Poisson effect and
the shear γxy is generally nonvanishing in presence of the corresponding coupling
term in D⋄ (see Eq. (660)).
The stresses in the lamina k may be obtained from Eq. (393), page 109 and Eq.
(658)

(k) (k)  0 (k)


{σ}X = z [Q]X κ X = z [Q]X [D⋄ ]X {M }X z ∈ [zk−1 , zk ] (671)

i.e.
 (k)  (k)  ⋄   M 
 σxx
 
 Q11 Q12 Q14 ⋄
D11 D12 ⋄
D14  x 
 
σyy = z  Q12 Q22 Q24   D12⋄ ⋄
D22 ⋄ 
D24 0 (672)
  ⋄ ⋄ ⋄  

σxy
 Q14 Q24 Q44 X D14 D24 D44 X

0

X X

- 197-
4 Mechanical analysis of composite laminates

so  (k)  (k)  ⋄ 
 σxx (x, z)
 
 Q11 Q12 Q14 D11
σyy (x, z) = zMx (x) Q12 Q22 Q24
   ⋄ 
D12 (673)
  ⋄

σxy (x, z)
 Q14 Q24 Q44 X D14 X
X
i.e.  (k)  (k)
 σxx (x, z)
 
  Gxx
 

σyy (x, z) = zMx (x) Gyy (674)

 
 
 

σxy (x, z) X
Gxy X
where
 (k)  (k)  ⋄ 
 Gxx
 
 Q11 Q12 Q14 D11
(k) ⋄ 
{G}X = Gyy =  Q12 Q22 Q24   D12
  ⋄

Gxy
 Q14 Q24 Q44 X D14 X
X
 (k) (k) (k)

⋄ Q
D11 ⋄ ⋄
11 + D12 Q12 + D14 Q14
 (k) (k)

(k) 
= ⋄ ⋄ ⋄ (675)
 D11 Q12 + D12 Q22 + D14 Q24 
⋄ Q(k) + D ⋄ Q(k) + D ⋄ Q(k)
D11 14 12 24 14 44 X

In the last expressions we have explicitly stated the system of representation in


which all the entities are represented, i.e. the global laminate axes X, in order to
emphasize that all tensors must be operated in the same system of representation.
Finally, we note that in contrast to isotropic beams, the larger stresses σxx are
not necessary those at the top and bottom faces of the beam because they depend
on the properties and orientation of each lamina. Generally, the larger σxx stresses
appear at laminae very stiff in the x direction and with a large z−coordinate. In
order to check the failure of each lamina of the laminate, the stresses given in Eq.
(673) must be previously rotated to the kth lamina principal material directions.

Exercise 36 Show that for isotropic beams Eq. (674) reads


   
 σxx 
   1 
Mx  
σyy =z ¯ 0 (676)

 
 Iy 
  
σxy 0

where I¯y = h3 /12.


Solution: We leave this exercise to the reader.

Interlaminar stresses
Recall that in the Kirchhoff plate theory we have considered vanishing σxz , σyz and
σzz , i.e. a plane stress condition throughout the plate. Hence, we have done so
for each lamina within the laminate. This assumptions is also typically considered
when developing the isotropic beam equations. However, as for the case of isotropic
beams, we know that these stresses, known as out-of-plane stresses, are nonvanishing
in general, see Figure 71.

- 198-
4.4 Deformations in composite laminated plates: some simple cases

Figure 71: Interlaminar stresses σxz and σzz in a laminated beam.

The value of these interlaminar stresses may be obtained integrating in z ∈


[zk−1 , zk ] the Local Equilibrium Equations (517) for each lamina, which yield re-
spectively
 !
Z z (k) (k)

 (k) ∂σ xx ∂σ xy (k)

 σxz = − + + ρüx − bx dz + Cxz


 zk−1 ∂x ∂y



 !
 Z z (k) (k)
(k) ∂σyy ∂σxy (k)
σyz = − + + ρüy − by dz + Cyz (677)

 zk−1 ∂y ∂x


 !

 Z z (k) (k)

 (k) ∂σ xz ∂σ yz (k)
 σzz = −
 + + ρüz − bz dz + Czz

zk−1 ∂x ∂y

(k) (k) (k)


where Cxz , Cyz and Czz are constants to be determined from the values at the
boundaries. Since for beams under bending load we assume that there is no depen-
dence on y, that σxy can be neglected and that by = üy = 0, the second equation
(k)
results in σyz = 0 and the first and third equations reduce to
 !
Z z (k)

 (k) ∂σ xx (k)
 σ =− + ρüx − bx dz + Cxz
 xz


zk−1 ∂x
! (678)
 Z z (k)

 (k) ∂σ xz (k)
 σzz = −
 + ρüz − bz dz + Czz

zk−1 ∂x

Assume now for simplicity the static case with no body loads. Then substituting
Eq. (674), noticing that the only dependence on x is in Mx (x) and considering the
relations given in Eq. (666), we obtain first
Z (k)
(k)
z
∂σxx (k) z 2 − zk−1
2
σxz =− dz + Cxz = −Qxz (x) G(k)
xx
(k)
+ Cxz (679)
zk−1 ∂x 2

- 199-
4 Mechanical analysis of composite laminates

and second
Z (k) Z
(k)
z
∂σxz (k)
z z 2 − zk−1
2
σzz =− dz + Czz = −fz (x) G(k)
xx
(k)
dz + Czz
zk−1 ∂x zk−1 2
z3 − 2 z
3zk−1 + 3
2zk−1
= −fz (x) G(k)
xx
(k)
+ Czz (680)
6
which show that the σxz distribution inside a lamina is piecewise parabolic and
the σzz stresses vary in a cubic manner (this result is parallel to that obtained in
isotropic beams, see Exercise 37). In these last expressions only the values of the
constants are to be determined from the boundary conditions. Assuming the case
in which there are no loads at the bottom of the beam at a given x−coordinate, by
equilibrium σxz (x, −h/2) = 0, so for the first, bottom layer, at the bottom face of
(1)
that layer (z = z0 = −h/2) we obtain Cxz = 0. Then, at the upper surface of that
bottom lamina we have
(1) z12 − z02
σxz (x, z1 ) = −Qxz (x) G(1)
xx
(1)
+ Cxz (681)
2 |{z}
=0

At the bottom face of the second lamina,


(2) (2)
σxz (x, z1 ) = Cxz (682)

By local equilibrium, this stress has to be the same as that at the top of the first
lamina (that is, interlaminar stresses are continuous in value throughout the thick-
ness; note the difference with in-plane stresses, which may be discontinuous in the
interfaces between layers) so

(2) z12 − z02


Cxz = −Qxz (x) G(1)
xx (683)
2
Then proceeding the same way to the following lamina and so on, we obtain the
recurrence relation for σxz
 2 2
 (1) (1) z − z0
 σ
 xz (x, z) = −Q xz (x) Gxx
2
2 2
(684)
 (k)
 σxz (x, z) = −Qxz (x) G(k)
 z − zk−1 (k−1)
xx + σxz (x, zk−1 ) ; k ≥ 2
2
In a similar way we can determine the following recurrence formula for σzz
 3 2 3
 (1) (1) z − 3z0 z + 2z0
 σ
 zz (x, z) = −f z (x) Gxx
6
3 2 3
(685)
(k) z − 3zk−1 z + 2zk−1

 (k) (k−1)
 σzz (x, z) = −fz (x) Gxx + σzz (x, zk−1 )
6
Exercise 37 Show that the out-of-plane stress components of Eqs. (679) and (680)
particularize for an isotropic beam to
Qxz (h/2)2 − z 2
σxz = ¯ (686)
Iy 2

- 200-
4.4 Deformations in composite laminated plates: some simple cases

Figure 72: Cylindrical bending

and
fz 2(h/2)3 + 3(h/2)2 z − z 3
σzz = ¯ (687)
Iy 6

Solution: We leave this exercise to the reader and simply note that σzz ∼ fz ,
σxz ∼ fz (L/h) and σxx ∼ fz (L/h)2 . Hence σxx ≫ σxz ≫ σzz if L/h ≫ 1, as usual.

Finally we note that we can now use the constitutive equations to obtain nonva-
(k) (k)
nishing values of γxz and εzz which contradict the classical (Kirchhoff) plate and
(Bernoulli) beam theory assumptions (γxz = εzz = 0). But this contradiction is ac-
cepted as the trade-off of the simplest ”strength of materials” theory when compared
to the ”elasticity theory”. However, we must also note that the obtained interlami-
nar stress distributions are a mean over those over the y−direction. Actually, there
is a variation on the y−direction which is most important at the edges; i.e. near
y = ±b/2. In the proximity of these locations (say close to the edges to the order of
the thickness h) the obtained values are not valid and the theory of elasticity must
be used. We will study these edge effects below.

4.4.2 Cylindrical bending

The second type of problem that can be treated unidimensionally is that of cylindri-
cal bending. In this case, in contrast to the beam theory, we consider the “width”
(length in y−direction) much larger than the “length” (in x−direction); i.e. b ≫ a,
see Figure 72.
In the cylindrical bending case, the displacements u0x , u0y , u0z can be considered
a function of x alone. Then the field equations may be considerably simplified.
Departing for example from Eqs. (570) to (572) from Exercise 26, page 178, and

- 201-
4 Mechanical analysis of composite laminates

considering only variations with respect to x:



 d2 u0x d2 u0y d3 u0z

 A 11 + A 14 − B 11 = −qx



 dx2 dx2 dx3



d2 u0x d2 u0y d3 u0z
A 14 + A 44 − B 14 = −qy (688)


 dx2 dx2 dx3



 4 0 3 0 d3 u0y
 D11 d uz − B11 d ux − B14


= qz
dx4 dx3 dx3
where for the linear static case qi = fi . These equations may be re-written in a more
comfortable way if we factor-out the second derivatives of u0x and u0y from the first
two equations: 
 d2 u0x d3 u0z
 A dx2 = B̄x dx3 + A14 qy − A44 qx


(689)
 d 2 u0 3 u0


 A y d z
= B̄y 3 − A11 qy + A14 qx
dx2 dx
where
A = A11 A44 − A214 ; B̄x = B11 A44 − A14 B14 ; B̄y = A11 B14 − B11 A14 (690)
Substituting in the third equation
 
d4 u0z B11 d4 u0z dqy dqx
D11 4 − B̄x 4 + A14 − A44 ...
dx A dx dx dx
 
B14 d4 u0z dqy dqx
− B̄y 4 − A11 + A14 = qz (691)
A dx dx dx
i.e.
d4 u0z
D̄ = q̄z (692)
dx4
where
B11 B̄x + B14 B̄y
D̄ = D11 − (693)
A
and
B̄x dqx B̄y dqy
q̄z = qz − − (694)
A dx A dx
Equation (692) may be integrated to obtain
Z x
d3 u0
D̄ 3z = q̄z (ξ) dξ + C1 (695)
dx 0

where C1 is the constant of integration to be obtained from boundary conditions.


This equation may be substituted in Eqs. (689) to obtain
Z x 
d2 u0x B̄x
A 2 = A14 qy − A44 qx + q̄z (ξ) dξ + C1
dx D 0
Z x  (696)
d2 u0y B̄y
A 2 = A14 qx − A11 qy + q̄z (ξ) dξ + C1
dx D 0

- 202-
4.4 Deformations in composite laminated plates: some simple cases

Equations (692) and (696) are uncoupled and can be independently integrated to
obtain the displacements. In the case q̄z is constant, the solution of the Eq. (692) is
polynomial in x and it is straightforwardly obtained. In the general case in which
q̄z (x) depends on the x−coordinate, it must be decomposed using Fourier series
following a similar procedure as that to obtain the solution of the heat equation (see
Eq. (462))
X∞  mπx 
q̄z (x) = q̄zm sin (697)
a
m=1
where Z a  mπx 
q̄z (x) sin dx Z
a 2 a  mπx 
q̄zm = 0Z a   = q̄ z (x) sin dx (698)
2 mπx a 0 a
sin dx
0 a
Then the solution to Eq. (692) is obtained following the same procedure, using
Fourier series
∞  mπx 
X q̄zm  a 4
u0z (x) = u0zm sin with u0zm = (699)
m=1
a π 4 D̄ m

Of course only a sufficient but limited number of terms in the series are considered.
Once the displacements are obtained, the strains and curvatures are deduced
from the kinematic equations for this particular case where there is no y-coordinate
dependency
du0 1 du0y
ε0xx = x ε0yy = 0 ε0xy = (700)
dx 2 dx
d2 u0
κ0xx = − 2z κ0yy = 0 κ0xy = 0 (701)
dx
and the stress resultants from Eqs. (564)

 du0x du0y d2 u0z

 N x = A 11 + A 14 − B 11 2
− Nx∗



 dx dx dx

du0x du0y d2 u0z
N y = A21 + A 24 − B 21 − Ny∗ (702)

 dx dx dx 2


 0 0 2 0
 Nxy = A41 dux + A44 duy − B41 d uz − N ∗


xy
dx dx dx2
and 
 du0x du0y d2 u0z

 Mx = B 11 + B 14 − D 11 − Mx∗


 dx dx dx2


du0x du0y d2 u0z
My = B 21 + B 24 − D21 − My∗ (703)

 dx dx dx 2


 0 0 2 0
 Mxy = B41 dux + B44 duy − D41 d uz − M ∗


xy
dx dx dx2
Finally, the stresses may also be computed in each lamina in a similar way as we
did for the laminated beam (i.e. by means of Eq. (393), page 109).

- 203-
4 Mechanical analysis of composite laminates

Exercise 38 Obtain the equations for cylindrical bending for the case of isotropy.
Solution: In the case of isotropy [B] = [0] and using Eqs. (693) and (694) we
obtain D̄ = D11 and q̄z = qz . Aside, we already know that

(1 − ν)
D11 = D22 = D̄ = D ; D12 = νD ; D44 = D; D14 = D24 = 0
2
D being the bending stiffness of the isotropic plate. Then, the solution Eq. (699)
is valid simply substituting D̄ by D and q̄zm by qzm . The resultant moments, Eqs.
(703), reduce to 
d2 u0z
− Mx∗


 M x = −D


 dx2
d2 u0z (704)
 M y = −νD 2
− My∗


 dx


Mxy = 0
∗ = 0 due to the isotropic nature of the hygro-thermal expansion coeffi-
where Mxy
cients.

4.4.3 Laminated simply supported rectangular plates: the Navier solu-


tion
For simply supported rectangular plates the typical analytical solutions are obtained
using bidimensional Fourier series. In this case the load must be decomposed in such
series, for example
∞ X

X mπx nπy
qz (x, y) = qzmn sin sin (705)
n=1 m=1
a b

where qzmn are the components of the load


Z bZ a
mπx nπy
sin
qz (x, y) sin dxdy
0 0 a b
qzmn = Z bZ a 
mπx nπy 2
sin sin dxdy
0 0 a b
Z Z
4 b a mπx nπy
= qz (x, y) sin sin dxdy (706)
ab 0 0 a b

Four terms of these series are shown in Figure 73.


The solution is also written in terms of such series that are already built to
automatically fulfill the field equilibrium equations and the corresponding boundary
conditions. For example, the following expansion of u0z satisfy the field equation for
the out-of-plane behavior of symmetric laminates (see Exercise 27, page 179) and
also the simply supported boundary conditions
∞ X

X mπx nπy
u0z (x, y) = u0zmn sin sin (707)
n=1 m=1
a b

- 204-
4.4 Deformations in composite laminated plates: some simple cases

Figure 73: Fourier components in a simply supported plate. Clockwise from left up-
per corner, cases {m = 1, n = 1}, {m = 1, n = 2}, {m = 2, n = 2}, {m = 2, n = 1}.

where the coefficients u0zmn are to be obtained from the particular partial differential
equation, i.e. Eq. (587). This is known as the Navier solution, and is the same
solution he employed to solve 2D elasticity problems (in-plane behavior). The form
of the solution for u0x and u0y also depends on the laminate layup and the boundary
conditions.
Obviously, solving the equations for the general laminated plate is a rather
lengthy task. Then we will only consider here explicitly the solution of special
cases of symmetric and antisymmetric laminates. We also introduce a rather gen-
eral procedure for other cases. The solution for many other particular cases may be
found, for example, in Reference [17].

Specially orthotropic symmetric laminates


For the case of specially orthotropic laminates we can use the simplified field equation
obtained in Exercise 27, i.e. Eq. (587), which is uncoupled with the in-plane behavior
and which we reproduce here for the reader comfort:
∂ 4 u0z ∂ 4 u0z ∂ 4 u0z
D11 + 2 (D12 + 2D44 ) + D22 = qz (x, y) (708)
∂x4 ∂y 2 ∂x2 ∂y 4
Then solution Eq. (707) and load representation Eq. (705) may be substituted in
this differential equation to obtain (we leave the algebra to the reader)
qzmn
u0zmn = 4 (709)
π dmn
with  m 4  m 2  n 2  n 4
dmn = D11 + 2 (D12 + 2D44 ) + D22 (710)
a a b b

- 205-
4 Mechanical analysis of composite laminates

We note that the obtained solution for u0z (x, y) has exactly the same components
that have been taken to approximate the external load qz (x, y) by means of Fourier
series (for example, those shown in Figure 73), but the coefficients for the load qzmn
and for the displacement u0zmn are different.
Once the displacements are obtained, the bending and twist moments can be
calculated using Eq. (563)2 and

∞ ∞
∂ 2 u0z (x, y) X X  m 2 2 0 mπx nπy
κ0x = − = π uzmn sin sin (711)
∂x2 a a b
n=1 m=1
∞ X
∞  
∂ 2 u0z (x, y) X n 2 mπx nπy
κ0y = − = π 2 u0zmn sin sin (712)
∂y 2 b a b
n=1 m=1
∞ X∞
0 ∂ 2 u0z (x, y) X m n mπx nπy
κxy = −2 = −2 π 2 u0zmn cos cos (713)
∂x∂y n=1 m=1
a b a b

Using [B] = [0] and D14 = D24 = 0 in Eq. (563)2 and substituting the preceding
results:
∞ 
∞ X  n 2  q
 m 2
X zmn mπx nπy
Mx = D11 + D12 2d
sin sin − Mx∗ (714)
n=1 m=1
a b π mn a b
∞ X ∞   m 2  n 2  q
X zmn mπx nπy
My = D12 + D22 2d
sin sin − My∗ (715)
n=1 m=1
a b π mn a b
∞ X
∞ m n q
X zmn mπx nπy
Mxy = −2D44 2
cos cos − Mxy

(716)
n=1 m=1
a b π dmn a b

The stresses may also be obtained in a similar way using the plane stress constitutive
equation for each lamina. Considering only the bending strains (the in-plane stresses
may be added just applying superposition)

(k) (k)  0
{σ}X = z [Q]X κ X
(717)

(k) (k)
so taking into account that we are in the case of special orthotropy (Q14 = Q24 = 0)
∞ X ∞   2  2  q
(k) m (k) n mπx nπy
X zmn
(k)
σxx (x, y, z) = z Q11 + Q12 2d
sin sin (718)
n=1 m=1
a b π mn a b
∞ X ∞   2 
 2 q
(k) m (k) n mπx nπy
X zmn
(k)
σyy (x, y, z) = z Q12 + Q22 2d
sin sin (719)
n=1 m=1
a b π mn a b
∞ X
∞ m n q
(k)
X (k) zmn mπx nπy
σxy (x, y, z) = −2zQ44 2
cos cos (720)
n=1 m=1
a b π dmn a b

Since qz include the effect of piezo-hygro-thermal loads (if present, see Eq. (568)2 ),
then to these last stresses we must subtract the piezo-hygro-thermal contribution,
see Sections 3.4.2 to 3.5. Finally, the interlaminar stresses may be obtained as in
the beam case integrating the Local Equilibrium Equations, see Eqs. (677), arriving

- 206-
4.4 Deformations in composite laminated plates: some simple cases

at (we leave this algebra to the reader)


 X
(k) z 2 − zk−1
2 ∞ X

(k) mπx nπy (k−1)
σxz =− σxzmn cos sin + σxz (x, y, zk−1 ) (721)
2 n=1 m=1
a b
 X
(k) z 2 − zk−1
2 ∞ X ∞
(k) mπx nπy (k−1)
σyz =− σyzmn sin cos + σyz (x, y, zk−1 ) (722)
2 n=1 m=1
a b
 X
(k) z3 − 2
3zk−1 3
z + 2zk−1 ∞ X

(k) mπx nπy (k−1)
σzz =− σzzmn sin sin + σzz (x, y, zk−1 )
6 n=1 m=1
a b
(723)

where
  q
(k) m 3 (k)  m   n 2  (k) (k) zmn
σxzmn = Q11 + Q12 + 2Q44 (724)
a a b πdmn
   q
(k) n 3 (k)  m 2  n   (k) (k) zmn
σyzmn = Q22 + Q12 + 2Q44 (725)
b a b πdmn
   m 2  n 2    n 4 
(k) m 4 (k) (k) (k) (k) qzmn
σzzmn = Q11 + 2 Q12 + 2Q44 + Q22 (726)
a a b b dmn

Exercise 39 Obtain the coefficients qzmn for a constant load q̂z .

Solution: The coefficients are obtained directly from Eq. (706) as


Z bZ a
4q̂z mπx nπy 16q̂z
qzmn = sin sin dxdy = 2 (727)
ab 0 0 a b π mn

for the cases when m, n are odd. Otherwise qzmn = 0. In Figure 73 only the first
component is nonzero. Note that these terms converge fast because as m and n get
large, qzmn → 0. For example for m = n = 1, qz11 = 1.62 q̂z , and for m = n = 3,
qz33 = 0.18 q̂z (one order of magnitude less). Furthermore, since the term qzmn /dmn
is of the order of the inverse of {m, n} to the fifth, all the preceding (displacement,
curvatures, resultant moments and stress) coefficients go faster to zero than those of
the load, except for those of the σzz interlaminar stress component (which go at the
same speed to zero as the load coefficients). Thus, only very few terms are frequently
needed (except for some type of loads like concentrated loads) to obtain a solution
within the engineering accuracy.

General procedure

In the previous subsection we have seen the Navier solution for the particular case
of specially orthotropic laminated plates. A generalization of this procedure may
be applied to more general cases of simply supported rectangular plates. However,
we want to emphasize that an exact solution of this type might not exist for some
combinations of boundary conditions. For the most general case, we can assume

- 207-
4 Mechanical analysis of composite laminates

displacement functions of the type


∞ X
∞ ∞ ∞
X mπx nπy X X 0sc mπx nπy
u0x = u0ss
xmn sin sin + uxmn sin cos +
m=1 n=1
a b m=1 n=1
a b
∞ X∞ ∞ ∞
X mπx nπy X X 0cc mπx nπy
+ u0cs
xmn cos sin + uxmn cos cos (728)
a b a b
m=1 n=1 m=1 n=1

∞ X
∞ ∞ ∞
X mπx nπy X X 0sc mπx nπy
u0y = u0ss
ymn sin sin + uymn sin cos +
m=1 n=1
a b m=1 n=1
a b
∞ X ∞ ∞ ∞
X mπx nπy X X 0cc mπx nπy
+ u0cs
ymn cos sin + uymn cos cos (729)
a b a b
m=1 n=1 m=1 n=1

∞ X
∞ ∞ ∞
X mπx nπy X X 0sc mπx nπy
u0z = u0ss
zmn sin sin + uzmn sin cos +
m=1 n=1
a b m=1 n=1
a b
∞ X
∞ ∞ ∞
X mπx nπy X X 0cc mπx nπy
+ u0cs
zmn cos sin + uzmn cos cos (730)
a b a b
m=1 n=1 m=1 n=1

each one having four coefficients per {m, n} pair. These twelve coefficients are
restricted in their values. First, the proposed solution in terms of Fourier series
have to fulfill the boundary conditions for the particular case at hand, and here
there are many possibilities. For example, one possibility is (case S2 in Eqs. (621))
(
at x = 0 and x = a : u0z = 0, Mx = 0, Nx = 0, u0y = 0
(731)
at y = 0 and y = b : u0z = 0, My = 0, Ny = 0, u0x = 0

In the absence of thermal, piezoelectric and moisture effects, from Eqs. (598) we
have
!
∂u0 ∂u0y ∂u0x ∂u0y ∂ 2 u0z ∂ 2 u0z ∂ 2 u0z
Nx = A11 x + A12 + A14 + − B11 − B 12 − 2B 14 (732)
∂x ∂y ∂y ∂x ∂x2 ∂y 2 ∂x∂y

!
∂u0 ∂u0y ∂u0x ∂u0y ∂ 2 u0z ∂ 2 u0z ∂ 2 u0z
Ny = A21 x + A22 + A24 + − B21 2
− B22 2
− 2B24 (733)
∂x ∂y ∂y ∂x ∂x ∂y ∂x∂y
!
∂u0 ∂u0y ∂u0x ∂u0y ∂ 2 u0z ∂ 2 u0z ∂ 2 u0z
Mx = B11 x + B12 + B14 + − D11 − D 12 − 2D 14 (734)
∂x ∂y ∂y ∂x ∂x2 ∂y 2 ∂x∂y
!
∂u0 ∂u0y ∂u0x ∂u0y ∂ 2 u0z ∂ 2 u0z ∂ 2 u0z
My = B21 x + B22 + B24 + − D21 − D 22 − 2D 24 (735)
∂x ∂y ∂y ∂x ∂x2 ∂y 2 ∂x∂y

which are to be particularized at the corresponding boundaries. Hence, for each


{m, n} pair Eqs. (731) establish eight conditions for the mentioned coefficients since

- 208-
4.4 Deformations in composite laminated plates: some simple cases

the solution is harmonic with fundamental wavelengths 2a and 2b. Three more
conditions are obtained from the equilibrium equations in terms of displacements,
Eqs. (570), (571) and (572), page 178. The last condition is given by the in-plane
compatibility condition, namely

∂ 2 ε0xx ∂ 2 ε0yy ∂ 2 γxy


0
+ = (736)
∂y 2 ∂x2 ∂x∂y

which, expressed in terms of displacements, completes the system of equations to


be solved. An analogous procedure is to be followed for other type of boundary
conditions.

Matrix form of the Navier solution for antisymmetric laminates


For antisymmetric laminates the Navier-type solution gives a solution of the dif-
ferential equation under some boundary conditions. For the case of antisymmetric
cross-ply laminates, the following solution is valid for S2 boundary conditions (just
consider the corresponding vanishing structural stiffness coefficients in Eqs. (732)
to (735) and perform the derivatives)
 ∞ X∞  mπx   nπy 
 X

 u 0 = u 0 cos sin
x xmn


 a b

 n=1m=1

 ∞ X∞

 0 X  mπx   nπy 
uy = u0ymn sin cos (737)
a b

 n=1m=1



 X∞ X∞  mπx   nπy 

 0 = 0

 u z u zmn sin sin

 a b
n=1m=1

For the case of angle-ply laminates, the following solution is valid for S3 boundary
conditions 
X∞ X∞  mπx   nπy 
 0 = 0

 u x u xmn sin cos



 n=1m=1
a b



 ∞ X∞  mπx   nπy 
 0 X
uy = u0ymn cos sin (738)
a b

 n=1m=1



 X∞ X∞  mπx   nπy 

 0 = 0

 u z u zmn sin sin

 a b
n=1m=1

The preceding two displacement fields also satisfy the field Equations (570), (571)
and (572) under certain conditions (for example, we must enforce Nxy = 0 in order
to fulfill the third equation because the term Nxy ∂ 2 u0z /∂x∂y gives contributions of
the form cos(mπx/a) cos(nπy/b)). Since Fourier functions are orthogonal, we obtain
three equations for each set of (possibly time-dependent) coefficients u0xmn , u0ymn and
u0zmn , where the spatial dependence of the problem has been removed. Each set of
three equations may be written in a handy matrix format (analogous to Eq. (604)).

- 209-
4 Mechanical analysis of composite laminates

We will omit the details here since they are straightforward but just lengthy:
  0    0   
K11 K12 K13  uxmn
 
 M11 0 M13  üxmn
 
  qxmn 
 
u0ymn ü0ymn
   
 K12 K22 K23  + 0 M22 M23  = qymn

 0 
 
 0 
   

K13 K23 K33 + G33 uzmn M13 M23 M33 üzmn qzmn
(739)
where for each {m, n} pair


 K11 = A11 cxx + A44 cyy




 K12 = (A12 + A44 ) cxy


 K13 = − (B11 cxxx + 3B14 cxxy + B24 cyyy )


K22 = A44 cxx + A22 cyy (740)


 K23 = − (B14 cxxx + 3B24 cxyy + B22 cyyy )






 K33 = D11 cxxxx + 2 (D12 + 2D44 ) cxxyy + D22 cyyyy


G33 = Nx cxx + Ny cyy

with
 mπ 2  mπ   nπ   mπ 2
cxx = , cxy = , cyy =
a a b b
 mπ 3  mπ   nπ 2  mπ   nπ 3
cxxx = , cxyy = , cxyyy = (741)
a a b a b
and so on. The other terms in Eq. (739) are

 M11 = M22 = m0z


 M
13 = −m1z cx
(742)


 M23 = −m1z cy

M33 = m0z + m2z (cxx + cyy )

and, for vanishing piezo-hygro-thermal loads


   
 qxmn   fxmn 
q = f (743)
 ymn   ymn 
qzmn fzmn

Equation (739) gives the displacement modal solution for the case under study.
For example, for static analysis Eq. (739) may be written as
  0   
K11 K12 K13  uxmn   qxmn 
 K12 K22 K23  u0 = qymn (744)
 ymn   
K13 K23 K33 + G33 u0zmn qzmn

or, more compactly, in terms of matrix blocks


  0   
K αβ K α3 uαmn q αmn
= (745)
K Tα3 K33 + G33 u0zmn qzmn

- 210-
4.4 Deformations in composite laminated plates: some simple cases

Figure 74: Load and solution form of the Levy-Nadai solution. Note that the plate
is simply supported along the y−direction, but any boundary condition is possible
along the x−direction.

If we want to directly compute the vertical displacement, from the first block

K αβ u0αmn + K α3 u0zmn = q αmn ⇒ u0αmn = K −1 0
αβ q αmn − K α3 uzmn (746)

which substituted in the second block yields for u0zmn

qzmn − K Tα3 K −1
αβ q αmn
u0zmn = (747)
K33 − K Tα3 K −1
αβ K α3 + G33

Hence, if the loads qzmn and qαmn are known, the quantities u0zmn can be determined
and then the solution is computed as

N X
X M  mπx   nπy 
u0z = u0zmn sin sin (748)
n=1m=1
a b

where the number of terms M and N in the summations are selected so that con-
vergence is attained to the desired tolerance (i.e. new terms over M and N do not
change significantly the computed response).

4.4.4 Laminated rectangular plates with two edges simply supported:


the Levy-Nadai solution
Another typical solution for isotropic plates that can also be used in the solution of
laminated composite plates is the Levy-Nadai solution. For simplicity we will also
only deal with specially orthotropic symmetric plates. However, the procedure can
also be applied to other type of laminated plates.
The Levy-Nadai solution is used for rectangular plates which are simply sup-
ported at two opposite ends and with arbitrary support (or free) conditions at the
other two ends, see Figure 74. In this case, the variables are expanded by means

- 211-
4 Mechanical analysis of composite laminates

of Fourier series over the direction perpendicular to the simply supported ends; the
x−axis in Figure 74. That is, the solution proposed by Levy is of the type

X mπx
u0z (x, y) = u0zm (y) sin (749)
a
m=1

and the load



X mπx
qz (x, y) = qzm (y) sin (750)
a
m=1
where u0zm (y) and qzm (y) are Fourier coefficient functions being defined for each
y−coordinate. Note that the sine function is already built such that it fulfills the
kinematical boundary conditions u0z = 0 at x = 0 and x = a for every m. Then,
inserting these expansions in the partial differential equation for the vertical dis-
placement of specially orthotropic plates, Equation (708), we obtain for each m
 mπ 4  mπ 2 d2 u0 (y) d4 u0zm (y)
D11 u0zm (y) − 2 (D12 + 2D44 ) zm
+ D 22 = qzm (y) (751)
a a dy 2 dy 4
which represents an ordinary differential equation for u0zm (y). The solution for this
equation may be obtained dividing the solution in an homogeneous part and a partic-
ular one. Substituting in the homogeneous part of Eq. (751), for which qzm (y) = 0,
a solution of the type
uH
zm (y) = Ce
λy
(752)
we eliminate the dependence on y, obtaining an algebraic equation for λ
 mπ 4  mπ 2
D11 − 2 (D12 + 2D44 ) λ2 + D22 λ4 = 0 (753)
a a
Focusing on Eq. (752), it is apparent that the actual solution form depends on the
nature (real or imaginary) of λ, which at the same time depends on the sign of
the discriminant of Eq. (753), ∆ ≡ (D12 + 2D44 )2 − D11 D22 , i.e. on the bending
structural stiffnesses of the laminate. If the discriminant is negative (the usual case
for cross-ply laminates), there are four complex values for λ, conjugate two by two.
If the discriminant is zero, there are real double roots λ = λ1 = λ2 = −λ3 = −λ4 ,
and if it is positive, there are four distinct real roots λ1 , λ2 , λ3 and λ4 . In summary
• ∆>0  r
 1  mπ 2 h √ i


 λ1 = D12 + 2D44 − ∆


 D22 a
 r


 1  mπ 2 h √ i
 λ2 = −λ1 = − D D12 + 2D44 − ∆


22 a
r (754)

 1  
mπ 2 h √ i

 λ3 = D12 + 2D44 + ∆


 D22 a

 r


 1  mπ 2 h √ i
 λ4 = −λ3 = −
 D12 + 2D44 + ∆
D22 a
a
since e = cosh(a) + sinh(a), cosh(−a) = cosh(a) and − sinh(−a) = sinh(a),
then uHzm (y) is given by
uH
zm (y) = C1m cosh (λ1 y) + C2m sinh (λ1 y) + C3m cosh (λ3 y) + C4m sinh (λ3 y) (755)

- 212-
4.4 Deformations in composite laminated plates: some simple cases

• ∆=0  r
 1  mπ 2
 λ1 = λ2 = D [D12 + 2D44 ]


22 a
r (756)

 1  mπ 2
 λ3 = λ4 = −
 [D12 + 2D44 ]
D22 a
Since there are double roots, and λ = λ1 = λ2 = −λ3 = −λ4 the solution is

uH
zm (y) = (C5m + C6m y) cosh λy + (C5m + C6m y) sinh λy (757)

• ∆ < 0, in this case λ = ±λR ± iλI , where i is the imaginary unit and λR > 0
and λI > 0 are the real and imaginary parts of the complex roots:

2 1  mπ 2 √ 
 λR = D D11 D22 + (D12 + 2D44 )


22 a
 (758)
 2 1 mπ 2 √ 
 λI =
 D11 D22 − (D12 + 2D44 )
D22 a
so since e(a+ib) = ea eib = [cosh(a) + sinh(a)][cos(a) + i sin(a)] we have
uH
zm (y) = [C7m cosh (λR y) + C8m sinh (λR y)] [C9m cos (λI y) + C10m sin (λI y)] (759)

Subsequently, the particular solution is obtained using the method of the unde-
termined coefficients, proposing a particular solution similar to the form of qzm (y)
(for example a quadratic function on y if qzm (y) is quadratic on y). For example
for the simplest case in which qzm (y) = qzm , a known constant, we propose

uPzm (y) = uPzm (constant) (760)

so upon substitution on Eq. (751)


 mπ 4 qzm
D11 uPzm = qzm ⇒ uPzm =  mπ 4 (761)
a
D11
a
For the linear case qzm (y) = Mm y + Nm , with Mm and Nm known constants, we
propose
uPzm (y) = Am y + Bm (762)
so upon substitution on Eq. (751)
 mπ 4
D11 (Am y + Bm ) = Mm y + Nm (763)
a
and identifying coefficients
Mm Nm
uPzm (y) =  mπ 4 y +  mπ 4 (764)
D11 D11
a a
Finally, in any case, the solution is

X  mπx
u0z (x, y) = uH P
zm (y) + uzm (y) sin (765)
m=1
a

- 213-
4 Mechanical analysis of composite laminates

The constants Ckm included in the homogeneous part uH


zm (y) of this last solution
can then be computed from the boundary conditions.
Once the displacement field has been obtained, strains, curvatures, stress resul-
tants and stresses are obtained as in the previous sections.

4.5 Buckling analysis

When a slender structure is subjected to compressive loads, there is a point at


which the initially small strain/displacements linear behavior bifurcates to a non-
linear, large displacements behavior. The structure searches a new configuration in
equilibrium because equilibrium is no longer possible for the infinitesimal displace-
ments state (i.e. the “undistorted configuration”). Then, the effective stiffness of
the structure decreases suddenly to a much smaller value because compressive loads
generate bending moments. A typical example is an office ruler under compressive
forces. Depending on the structure, collapse is possible at that buckling load or
otherwise the structure may sustain a significantly larger amount of load at the cost
of larger displacements and decreasing stiffness. The first case is typical of columns,
whereas the second case is typical of plates. In any of these cases, it is important to
determine the level of load at which buckling takes place (the buckling critical load)
and which will be the deformed shape after reaching that load level.
Buckling analysis is basically an eigenvalue analysis very similar to a dynamic
analysis but in place of the mass matrix we have a geometric stiffness which accounts
for linear second order effects of the compressive loads. A systematic and formal
determination of the geometric terms involves the use of the large strain and stress
quadratic measures, i.e. the nonlinear Green-Lagrange strains and second Piola-
Kirchhoff stresses. However, we can avoid that more complex analysis just including
the linearized second order effects of the force stress resultants in the z-direction,
i.e. Equation (529), page 171. This way we include large displacements but still
work with small (linear) strains.
Here we will only analyze some simple cases in order to gain some insight into
the buckling behavior of laminated beams and plates. In general, in a finite element
program, a buckling analysis involves a linear eigenvalue analysis. The postbuckling
analysis is nonlinear and is frequently performed using a combination of a Newton-
Raphson method with a special continuation method such as the arc-length (Riks)
method.

4.5.1 Buckling of symmetric laminated beams

In the case of laminated beams, only the Nx stress resultant is present. Then,
considering the beam internal equilibrium Equation (666) along with the beam con-
stitutive Equation (661) with only the geometric (second order) terms in Eq. (540)

- 214-
4.5 Buckling analysis

due to Nx , we arrive at

d2 Mx 
(666): = −qz 

dx2 



d2 u0z ⋄ M
 1 d4 u0z d2 u0z
(661): = −D11 x ⇒ ⋄ dx4 + N̄ =0 (766)
dx2 
 D11 dx2
  
du0

d 
Nx z

(540): qz = 

dx dx

where we have assumed a constant compressive force per unit width Nx = −N̄ in
the beam (note the change of sign so that N̄ > 0 means compression) and, see Eq.
(663)
1 det ([D⋄ ]) Ēx h3
⋄ = = (767)
D11 D22 D44 − 2D24 12
where Ēx is the effective Young modulus of the beam.
Equation (766) is the same equation than that obtained for isotropic beams
except for the computation of the effective value Ēx . The solution of this differential
equation is obtained the same way as for other equations we have previously solved.
First, assume a solution of the type

u0z = U eλx (768)

and insert it in the differential equation:


 
1 4 2
⋄ λ + N̄ λ U eλx = 0 (769)
D11
Then, if U 6= 0 and λ 6= −∞, the term in parenthesis has to vanish, which happens
for
λ2 = 0 ⇔ two equal roots λ1 = λ2 = 0
and
q
λ2 = −N̄ D11

<0 ⇔ ⋄
two imaginary roots λ3 = −λ4 = iλ̄ = i N̄ D11

Since e0 = 1, xe0 = x and eib = cos(b) + i sin(b), the general expression of the
solution is
u0z = U1 + U2 x + U3 sin λ̄x + U4 cos λ̄x (770)
where U1 , U2 , U3 , U4 are constants to be determined from the boundary conditions
of the beam and where
λ̄2 Ēx h3
q
λ̄ = N̄ D11⋄ ⇔ N̄ = ⋄ = λ̄2 (771)
D11 12
For example, if the beam is simply supported at both ends we have the following
kinematical boundary conditions
( 0
uz (x = 0) = 0
(772)
u0z (x = a) = 0

- 215-
4 Mechanical analysis of composite laminates

and also the moment boundary conditions Mx (x = 0) = 0 and Mx (x = a) = 0,


which if there is no extensional-bending coupling provide

 d2 u0z
 Mx (x = 0) = 0 ⇒ =0


dx2 x=0
(773)

 d2 u0z
 Mx (x = a) = 0
 ⇒ =0
dx2 x=a

yielding the following system of equations




 U1 + U4 = 0



 U1 + U2 a + U3 sin λ̄a + U4 cos λ̄a = 0
(774)


 U4 = 0



−U3 sin λ̄a − U4 cos λ̄a = 0

so U1 = U2 = U4 = 0 6= U3 and

sin λ̄a = 0 ⇒ λ̄n a = nπ ⇒ λ̄n = with n = 1, 2, 3, ... (775)
a

i.e. by Eq. (771)

Ēx h3  nπ 2 det ([D⋄ ])  nπ 2


N̄n = = (776)
12 a D22 D44 − 2D24 a

The critical load N̄cr will be the lowest one among all the possible buckling loads,
hence n = 1 gives

Ēx h3  π 2 det ([D⋄ ])  π 2


N̄cr = = (777)
12 a D22 D44 − 2D24 a

which is the generalization of the Euler critical load of an isotropic column N̄cr =
Ex Iy π 2 /L2 to a composite beam (simply consider Iy = h3 /12, Ex = Ēx and L = a).
The buckling mode adopts also the same shape of deformation than for the isotropic
case
πx
u0z = U3 sin(λ̄1 x) = U3 sin (778)
a
where U3 remains undetermined. To determine this value, a nonlinear large-deformation
strain-driven postbuckling analysis is needed. Buckling modes for n = 1, 2 and 3 are
shown in Figure 75. Note that since the buckling load of mode 1 is lower, the other
ones never appear in practice. Buckling modes for different support conditions are
shown in Figure 76.
We note that the critical buckling load is a non-conservative value. Usually
imperfections decrease that value significantly. Moreover, additional transversal
deflections due to external loads fz also decrease the critical load to a significant
extent. However, accounting for such effects is out of the scope of these notes.

- 216-
4.5 Buckling analysis

Figure 75: Buckling modes of a simply supported beam.

Figure 76: Fundamental (first) buckling modes for different support conditions.

- 217-
4 Mechanical analysis of composite laminates

Exercise 40 Determine the critical buckling load of a symmetric-layered laminated


beam clamped at both ends.
Solution: The solution is obtained using the boundary conditions
 0

 uz (x = 0) = 0





 u0z (x = a) = 0


du0z (779)
ϕy (x = 0) ⇒ =0


 dx x=0


du0z



 ϕy (x = a) ⇒
 =0
dx x=a

We leave the reader to verify that the following solution is obtained



λ̄n = n with n ≥ 1 (780)
a
i.e. for n = 1
 2  2
Ēx h3 2π det ([D⋄ ]) 2π
N̄cr = = (781)
12 a D22 D44 − 2D24 a
which is four times greater than the buckling load of a simply supported beam of the
same length (a) and material properties. Aside, this solution is equivalent to take
the Euler Solution (776) with half the length. Accordingly, and taking the simply
supported beam as the reference case, it is said the Buckling Equivalent Length of a
beam clamped at both ends is a/2.

Exercise 41 Determine the critical buckling load of a symmetric-layered laminated


beam clamped at one end and free at the other end.
Solution: We leave this exercise to the reader and simply quote the solution

(2n − 1) π  π 2 Ē h3 π 2 Ēx h3
x
λ̄2n−1 = ⇒ N̄cr = = 2 (782)
2a 2a 12 4a 12
with
u0z = −U + U cos λ̄x (783)

Exercise 42 Determine the critical buckling load of a symmetric-layered laminated


beam clamped at one end and simply supported at the other end.
Solution: We leave this exercise to the reader and simply quote the solution
 4.4934 π
λ̄n a = tan λ̄n a ⇒ λ̄1 = = (784)
a 0.6992 a
The reader should also obtain the general expression for u0z .

Exercise 43 Show that the buckling solutions for a beam are essentially valid for
cylindrical bending.

- 218-
4.5 Buckling analysis

Solution: We recall that for cylindrical bending the equilibrium equations are given
by Eqs. (689) and (692). For vanishing in-plane loads and homogeneous compressing
Nx = − N̄ , they reduce to

 d2 u0x d3 u0z

 A = B̄ x



 dx2 dx3

d2 u0y d3 u0z
A = B̄ y
(785)

 dx 2 dx 3


 4 0 2 0
 D̄ d uz = qz = −N̄ d uz


dx4 dx2

The third equation is uncoupled and equivalent to Eq. (766) with D̄ playing the role
of 1/D11⋄ . Then, the solutions are also essentially the same upon such substitution.

Once u0z is obtained, the other two equations may be used to obtain the form of u0x
and u0y .

4.5.2 Buckling of specially orthotropic laminated plates


Buckling of plates is more complex than for beams because in this case buckling
modes are bidimensional and the buckling loads may be a combination of Nx , Ny
and/or Nxy . For most cases an analytical exact solution is not possible even for
the simple geometry of rectangular plates. Here we will only analyze the most
simple cases of symmetric specially orthotropic plates and antisymmetric cross-ply
and angle ply laminated plates just in order to gain some insight in the buckling
behavior of plates.

Buckling of specially orthotropic laminated plates under uniaxial and


biaxial compression
For simplicity assume that there are no Nxy shear loads, and take

Nx = −N̄ ; Ny = kNx = −k N̄ ; N̄ = constant (786)

where k represents the biaxial proportionality factor. Then, the plate equilibrium
equation in terms of stress resultants and including second order effects is, see Ex-
ercise 22, Equation (550), page 174

∂ 2 Mx ∂ 2 Mxy ∂ 2 My ∂ 2 u0z ∂ 2 u0z


+ 2 + − N̄ − k N̄ =0 (787)
∂x2 ∂x∂y ∂y 2 ∂x2 ∂y 2

where we have assumed vanishing body and out-of-plane loads (i.e. only boundary
loads are considered). Consider now the case of symmetric specially orthotropic
plates, i.e. [B] = [0] and A14 = A24 = D14 = D24 = 0. Then, {M } = [D]{κ} and
the equilibrium equation in terms of displacements is (compare to Eq. (587))

∂ 4 u0z ∂ 4 u0 ∂ 4 u0z ∂ 2 u0z ∂ 2 u0z


D11 4
+ 2 (D12 + 2D44 ) 2 z 2 + D22 4
+ N̄ 2
+ kN̄ =0 (788)
∂x ∂y ∂x ∂y ∂x ∂y 2

- 219-
4 Mechanical analysis of composite laminates

The analytical exact solution of this last equation can also be obtained using the
Navier procedure; i.e. a solution of the form
mπx nπy
u0z = u0zmn sin sin (789)
a b
is proposed, with m and n being integer numbers, so the simply supported boundary
conditions are directly fulfilled. Substituting as always the proposed solution Eq.
(789) into Eq. (788) we obtain that if u0zmn 6= 0 (the case u0zmn = 0, under axial
loads, corresponds to the unbuckled solution)
  mπ 4  mπ 2  nπ 2  nπ 4 
D11 + 2 (D12 + 2D44 ) + D22 ...
a a b b
 mπ 2  nπ 2 (790)
− N̄ − kN̄ = 0
a b
If we define the plate aspect ratio by r = a/b, then
D11 m4 + 2 (D12 + 2D44 ) m2 (nr)2 + D22 (nr)4  π 2
N̄ = (791)
m2 + (nr)2 k a
The critical load is
( )
 D11 m4 + 2 (D12 + 2D44 ) m2 (nr)2 + D22 (nr)4  π 2
N̄cr = min N̄ = min 2 (792)
m,n m,n m2 + (nr) k a
It is interesting to note that in contrast to the beam buckling, the critical load does no
necessarily occur for m = n = 1 and that, in fact, it may depend on the plate aspect
ratio through nr and on the biaxial proportionality factor k. The next exercises
analyse this fact. In order to facilitate the analysis it is usual to adimensionalize the
buckling load. A typical example for k = 0 to account for different b values for fixed
a dimension is  
N̄ N/m
Nad =  2 units: (793)
π (1/ m2 ) N m
D22
b
because Nx is force per unit width (b).
Another important difference between buckling of plates and beams is that when
the critical load is reached in a beam, the beam does not longer support larger loads
and collapse is imminent if the loads are increased, see Figure 77a (however, note that
displacement-driven postbuckling analyses could still be performed after the critical
load is reached). Hence relevant safety factors must be considered. In contrast, in
plates when the critical load is reached, the plate just reduces its stiffness to some
extent, as shown in Figure 77b and further loading can be resisted. This different
behavior of plates from beams is due to the existing transversal stiffness in plates
which is not present in beams. Hence, the postbuckling behavior depends on these
transversal stiffnesses, a fact that can be intuitively deduced comparing the field
equations for buckling of plates (under uniaxial loading) and beams:
∂ 4 u0z ∂ 4 u0z ∂ 4 u0z ∂ 2 u0z
D11 + 2 (D12 + 2D44 ) + D22 + N̄ =0
∂x4 ∂y 2 ∂x2 ∂y 4 ∂x2
1 d4 u0z d2 u0z
⋄ dx4 + N̄ =0
D11 d2 x

- 220-
4.5 Buckling analysis

Figure 77: Post-buckling behavior of a beam (a) and a plate (b).

Exercise 44 Write a Matlab code to obtain the buckling loads for different m, n, r
and k.
Solution: The code follows

function [N,Nad] = Biaxialplatebuckling(D,k,a,r,m,n)


%
% function [N,Nad] = Biaxialplatebuckling(D,k,a,r,m,n)
% This is a function to compute the Buckling loads for a simply
% supported specially orthotropic laminated plate under biaxial
% compressive loading
%
% Input:
% D = Moment-curvature constitutive matrix of plate
% k = proportionality factor for biaxial loading
% a = dimension in x of the plate
% r = aspect ratio (may be vector to compute for several ratios)
% m = half wavelength in x (may be vector to compute for several ones)
% n = half wavelength in y (may be vector to compute for several ones)
%
% Output:
% N(m,n,r) = Array of buckling loads for values m, n, and r
% Nad(m,n,r) = adimensional buckling load; N*[(b/pi)^2 / D(2,2)]
%

nm = length(m); nn = length(n); nr = length(r); % num. of terms


N = zeros(nm,nn,nr); % initializes array to zero
Nad = zeros(nm,nn,nr); % initializes array to zero
%

- 221-
4 Mechanical analysis of composite laminates

for im = 1:nm, % for each x-halfwave


for in = 1:nn, % for each y-halfwave
for ir = 1:nr, % for each aspect ratio
ntr = n(in)*r(ir); % n times r
N(im,in,ir) = D(1,1)*m(im)^4 + ...
2*(D(1,2)+2*D(3,3))*m(im)^2*ntr^2 + D(2,2)*ntr^4;
N(im,in,ir) = N(im,in,ir) * (pi/a)^2 / (m(im)^2 + k * ntr^2);
Nad(im,in,ir) = N(im,in,ir) * (a/r(ir))^2 / pi^2 / D(2,2);
end
end
end
%
return

Exercise 45 Use the previous Matlab code to plot the results for a Graphite-Epoxy
[0, 90, 0̄]s composite with a = 1 and ply thickness of t = a/1000. Use adimensional
buckling loads. Comment the results.
Solution: We leave to the reader the task of writing the Matlab script to yield
the plot of Figure 78 and to verify that it is insensitive to the value of a and to
the thickness t of the plies because it is adimensionalized. In fact only the ratios
D11 /D22 and (D12 + 2D44 ) /D22 are relevant for the plots. In this plot it is seen
that the minimum values are obtained for n = 1; i.e. only one half wave in the y-
direction. Then, in the same figure it can be observed that, depending on the aspect
ratio r of the plate, the minimum buckling load (i.e. the critical buckling load) is
obtained for different values of m. The physical interpretation of this observation is
given in Figure 79. When the length of the plate doubles, the same mode wavelength
is obtained after doubling the value of m. This observation is clearly obtained for
isotropic plates.

Exercise 46 Obtain the curves of Exercise 45 for the case of isotropic plates.
Solution: In the case of isotropic plates, using Eqs. (588), Exercise 28, page 180,
and taking k = 0, we have
 2 2 
m b 2 n 4 a2 π2
N̄ = + 2n + D (794)
a2 m 2 b2 b2

where D = Eh3 / 12 1 − ν 2 . Obviously, since n ≥ 1, the minimum buckling loads
are obtained for n = 1, then
m r 2 π 2
N̄ = + D 2 (795)
r m b
The adimensional buckling load is then
m r 2
Nad = + (796)
r m
The plot shown in Figure 80 may be done using Eq. (796) or obtained using the
function of Exercise 45 taking D11 = D22 = D, D12 = νD, D44 = (1 − ν) D/2.

- 222-
4.5 Buckling analysis

Graphite−Epoxy [0,90,0,90,0]; t=a/1000; k=0


n=1 n=2
25 25
Adimensional buckling load Nad = N (b / π) / D22

20 20
2

15 15

m=1
m=2
10 10 m=3
m=4
m=5
m=6
5 5 m=7
Minimum

0 0
1 2 3 4 5 1 2 3 4 5
Aspect ratio r = a / b Aspect ratio r = a / b

Figure 78: Figure of Exercise 45. Buckling loads of a simply supported, specially
orthotropic Graphite-Epoxy laminated plate.

- 223-
4 Mechanical analysis of composite laminates

Figure 79: Figure of Exercise 45. Buckling modes of a plate. Note that the plate
tries to preserve similar wavelengths in both directions (the ratio depends on the
plate stiffnesses D), hence, roughly speaking, doubling r = a/b implies that the value
of m for the minimum critical load also doubles.

- 224-
4.5 Buckling analysis

Isotropic plate
10

Adimensional buckling load Nad = N (b / π) D


m=1

2
9 m=2
8 m=3
m=4
7 m=5
Minimum
6

0
0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Aspect ratio r = a / b

Figure 80: Figure of Exercise 46. Adimensional buckling loads for an isotropic plate.

Exercise 47 Plot the adimensional buckling loads for the case of equibiaxial com-
pression; i.e. k = 1 for a Graphite-Epoxy [0, 90, 0̄]s
Solution: Figure 81 shows the result. In the case of equibiaxial compression it is
seen that the critical load is typically obtained for n = 1 and that the value of m = 1
is significative; i.e. the same critical buckling load is obtained for large values of r
(with m > 1), but the value remains almost constant, close to one in this particu-
lar case. However, next exercise shows that this is not always the case, and that it
depends on the flexure properties of the laminate.

Exercise 48 Plot the adimensional buckling loads for the case of equibiaxial com-
pression; i.e. k = 1 for a Graphite-Epoxy [0, 0, 90]s composite laminate.
Solution: Figure 82 shows the result. It is seen that in this case, for low r values
the mode with m = 1, n = 2 yields the critical buckling load.

Exercise 49 Plot the adimensional buckling loads for the case of equibiaxial com-
pression; i.e. k = 1 for an isotropic plate.
Solution: Figure 83 shows the result. As it should be intuitive to the reader now,
the critical load is obtained for m = 1, n = 1.

Exercise 50 Plot the buckling loads (dimensional) for the case of uniaxial compres-
sion for stacking sequences [0, 90, 0̄]s and [0, 0, 0̄]s . Comment the results.
Solution: Figure 84 shows the result. It can be seen that for r > 1, the buckling
load is greater for a cross-ply laminate than for a similar unidirectional laminate.
The difference is more marked when b becomes smaller (r becomes larger for fixed
a) because the transversal stiffness becomes increasingly relevant.

- 225-
4 Mechanical analysis of composite laminates

n=1 n=2
5 5
22
= N (b / π) / D

4.5 4.5
2

4 4

3.5 3.5
ad

3 3
Adimensional buckling load N

2.5 2.5

2 2

1.5 1.5

1 1

0.5 0.5

0 0
2 4 6 8 2 4 6 8
Aspect ratio r = a / b Aspect ratio r = a / b

Figure 81: Figure of Exercise 47. Buckling loads for equibiaxial compression of a
Graphite-Epoxy laminate with stacking sequence [0, 90, 0̄]s

Graphite−Epoxy [0,0,90,0,0]; k=1


12
Adimensional buckling load Nad = N (b / π)2 / D22

m=1, n=1
m=2, n=1
10 m=3, n=1
m=4, n=1
Minimum n=1
8 m=1, n=2
m=2, n=2
m=3, n=2
6 m=4, n=2
Minimum n=2
Infimum
4

0
1 2 3 4 5 6 7 8
Aspect ratio r = a / b

Figure 82: Figure of Exercise 48. Buckling loads of a Graphite-Epoxy laminate


under equibiaxial compression with ply distribution [0, 0, 90]s .

- 226-
4.5 Buckling analysis

Isotropic plate; k=1


12

Adimensional buckling load Nad = N (b / π) / D


m=1, n=1
2
m=2, n=1
10 m=3, n=1
m=4, n=1
Minimum n=1
8 m=1, n=2
m=2, n=2
m=3, n=2
6 m=4, n=2
Minimum n=2
Infimum
4

0
1 2 3 4 5 6 7 8
Aspect ratio r = a / b

Figure 83: Figure of Exercise 49. Buckling loads of an isotropic plate under equibi-
axial compression.

4
x 10 Graphite−Epoxy; a = 1; t=a/1000; k=0
9

8
Stacking seq. [0,90,0,90,0]
7 Stacking seq. [0,0,0,0,0]
Buckling load N

1
0.5 1 1.5 2
Aspect ratio r = a / b

Figure 84: Figure of Exercise 50. Comparison of buckling loads for unidirectional
and cross-ply laminates.

- 227-
4 Mechanical analysis of composite laminates

Exercise 51 Show that Eq. (791) may be obtained from Eq. (739).
Solution: In the case of buckling, Eq. (739) is
  0   
K11 K12 K13  uxmn   0 
 K12 K22 K23  u0 = 0 (797)
 ymn   
K13 K23 K33 + G33 u0zmn 0
where for each m, n pair the third equation is uncoupled because K13 = K23 = 0 and
 mπ 4  mπ 2  nπ 2  nπ 4
K33 = D11 + 2 (D12 + 2D44 ) + D22 (798)
a a b b
 mπ 2  nπ 4
G33 = Nx + Ny (799)
a b
Then, Eqs. (790) and (791) immediately follow.

Buckling of specially orthotropic laminated plates under shear loading


Another interesting and important case of instability is the case of buckling under
shear load. This is a typical case of web panels, where instability is obtained due
to compressive stresses at 45. In this case an exact solution has not been obtained
because the Navier proposal does not solve the associated buckling differential equa-
tion
∂ 4 u0z ∂ 4 u0z ∂ 4 u0z ∂ 2 u0z
D11 + 2 (D12 + 2D44 ) + D22 − 2Nxy =0 (800)
∂x4 ∂y 2 ∂x2 ∂y 4 ∂x∂y
However, an approximate solution can be found using a handy procedure: the
Rayleigh-Ritz approach. In this case an approximation of the type of Navier can
be proposed (note that in this case it is merely an approximation, not a solution).
Then we need to take enough terms up to achieve the desired tolerance
N X
M N X
M
X X mπx nπy
u0z = u0zmn φmn (x, y) = u0zmn sin sin (801)
a b
n=1 m=1 n=1 m=1

where N and M are to be determined from a convergence check and φmn (x, y)
are the approximation functions. These functions are complete (i.e. they approxi-
mate any possible function taking enough terms) and linearly independent (in fact,
orthogonal), which are two properties being required by the procedure to obtain
convergence. Note that the proposed functions satisfy the essential boundary con-
ditions (automatically, by construction) and also the natural boundary conditions
(for certain, very common, ply layups).
Then the proposed approximation may be used to minimize the potential energy,
see Eq. (656), Exercise 35
Z "  2 0 2  2 0 2
1 ∂ uz ∂ 2 u0z ∂ 2 u0z ∂ uz
Π= D11 2
+ 2D12 2 2
+ D22 +
2 S ∂x ∂x ∂y ∂y 2
 2 0 2  0   0 #
∂ uz ∂uz ∂uz
+ 4D44 + 2Nxy dxdy (802)
∂x∂y ∂x ∂y

- 228-
4.5 Buckling analysis

Inserting the proposed approach and minimizing the functional Π respect to an


arbitrary coefficient u0zpq we arrive at the following equation after some lengthy
algebra
  pπ 4  pπ 2  qπ 2  qπ 4  ab
D11 + 2 (D12 + 2D44 ) + D22 u0
a a b b 4 zpq
N X M 
X mπ   nπ 
−2Nxy Pmnpq u0zmn = 0 (803)
a b
n=1 m=1
where
Z b Z a
mπx nπy pπx qπy 4ab pq
Pmnpq = cos cos sin sin dxdy = 2 2 (804)
0 0 a b a b π (p − m2 ) (q 2 − n2 )
for p 6= m and q 6= n with p ± m and q ± n being odd; otherwise Pmnpq = 0.
Equation (803) constitutes a eigenvalue equation, where Nxy is the eigenvalue
to be found. Usually we may take N = M , so N 2 = M 2 coefficients (degrees of
freedom) are to be obtained. This equation may be written in matrix form as
([K] − Nxy [G]) {U } = 0 (805)
where
   4  2  qπ 2  qπ 4  ab
 K(pq)(pq) = D11 pπ + 2 (D12 + 2D44 ) pπ

+ D22
a a b b 4 (806)


K(pq)(rs) = 0 for (pq) 6= (rs)
and  rπ   sπ 
G(pq)(rs) = 2 Prspq ; U(pq) = u0zpq (807)
a b
with the parentheses in the subscripts meaning a collapse of two indices to one index
according to any proper custom rule. Note that [K] is a diagonal matrix but [G] is
not a positive-definite matrix, so a proper eigenvalue algorithm must be employed.
Of course the lowest eigenvalue found is the critical load.
Exercise 52 Write a Matlab code to compute the critical shear load for a laminated
plate and the corresponding buckling mode.
Solution: We leave this exercise to the reader. The Matlab function to find the
generalized eigenvalues and eigenvectors is eigs.

4.5.3 Buckling of antisymmetric laminated plates


In the case of antisymmetric laminated plates there are nonvanishing extension-
bending coupling coefficients. In these cases the middle-plane displacement compo-
nents u0x , u0y , u0z are coupled, with the following implications. First, proposals for all
three components must be given (in contrast to the previous cases where u0z was suf-
ficient) and the corresponding three equations must be simultaneously solved, so the
problem is algebraically much more involved. Then, boundary conditions must also
be prescribed for these components. This means that for example we have several
options of “simply supported” laminates, see Eq. (621). Since algebra is lengthy
but similar procedures to the already studied examples follow, we simply quote the
results.

- 229-
4 Mechanical analysis of composite laminates

Buckling of antisymmetric cross-ply laminates


In antisymmetric cross-ply laminates A11 = A22 , A14 = A24 = 0, D11 = D22 , D14 =
D24 = 0, and there are two nonvanishing bending-extension coupling B22 = −B11 .
Then, the buckling equilibrium equations for the case of compressive biaxial loads
Nx = −N̄ and Ny = −kN̄ are (just particularize the field equations given in Exercise
26, page 178):

∂ 2 u0x ∂ 2 u0x ∂ 2 u0y ∂ 3 u0z


A11 + A44 + (A12 + A44 ) − B 11 =0 (808)
∂x2 ∂y 2 ∂x∂y ∂x3

∂ 2 u0x ∂ 2 u0y ∂ 2 u0y ∂ 3 u0z


(A12 + A44 ) + A44 + A11 + B 11 =0 (809)
∂x∂y ∂x2 ∂y 2 ∂y 3

∂ 4 u0z ∂ 4 u0z ∂ 4 u0z


D11 +2 (D12 + 2D 44 ) + D 11
∂x4 ∂x2 ∂y 2 ∂y 4
3 0
∂ 3 u0x ∂ uy ∂ 2 u0z ∂ 2 u0z
− B11 + B 11 + N̄ + k N̄ =0 (810)
∂x3 ∂y 3 ∂x2 ∂y 2
A solution can be proposed in order to satisfy the field equations and the boundary
conditions. The following solution fulfills simply supported boundary conditions S2
in Eq. (621), see the respective case within Section 4.4.3
  mπx   nπy 
 u 0 = u0 cos sin
 x xmn


 a b

 0  mπx   nπy 
0
uy = uymn sin cos
a b (811)

  mπx   nπy 

 0 0
 uz = uzmn sin sin


a b

and yields, upon substitution in the previous equilibrium equations, the following
buckling load associated to the (m, n) buckling mode
 2 − K K2

1 2K12 K23 K13 − K22 K13 11 23
N̄ (m, n) = K33 + 2 (812)
cxx + kcyy K11 K22 − K12
where 
 K11 = A11 cxx + A44 cyy




 K12 = (A12 + A44 ) cxy


 K = −B11 cxxx
13
(813)


 K22 = A44 cxx + A11 cyy




 K23 = B11 cyyy

K33 = D11 (cxxxx + cyyyy ) + 2 (D12 + 2D44 ) cxxyy
Compare with Equations (739) and (740). Obviously the critical load is the smallest
buckling load among all m, n values.

N̄cr = min N̄ (m, n) (814)
m,n

- 230-
4.5 Buckling analysis

Exercise 53 Write a Matlab code to compute the critical buckling load for antisym-
metric cross-ply laminates
Solution: The code follows for the case k = 0.

function [N,Nad] = AntisymCrossplyBuckling(A,B,D,a,r)


%
% function [N,Nad] = AntisymCrossplyBuckling(A,B,D,a,r)
% This is a function to compute the Buckling loads for a simply
% supported antisymmetric cross-ply laminate under uniaxial (k=0)
% compressive loading
%
% Input:
% A = Extension constitutive matrix of plate
% B = Extension-bending coupling constitutive matrix of plate
% D = Moment-curvature constitutive matrix of plate
% a = dimension in x of the plate
% r = aspect ratio (may be vector to compute for several ratios)
%
% Output:
% N(r) = Array of buckling loads for values r
% Nad(r) = adimensional buckling load; N*[(b/pi)^2 / D(2,2)]
%

nm = 10; nn = 10; % number of halfwaves to use in computation


nr = length(r); % num. of terms in array
N = zeros(nr); % initializes arrays to zero
Nad = zeros(nr); % initializes arrays to zero
Nmn = zeros(nm,nn); % initializes arrays to zero
%
for ir = 1:nr % for each aspect ratio
b = a / r(ir); % b (width) dimension for r aspect ratio
for m = 1:nm % for each x-halfwave
for n = 1:nn % for each y-halfwave
am = m * pi / a; % some coefficients, see formulae
bn = n * pi / b;
K11= A(1,1)*am^2 + A(3,3)*bn^2;
K22= A(1,1)*bn^2 + A(3,3)*am^2;
K33= D(1,1)*(am^4+bn^4) + 2*(D(1,2)+2*D(3,3))*am^2*bn^2;
K12= (A(1,2) + A(3,3))*am*bn;
K13= -B(1,1)*am^3;
K23= B(1,1)*bn^3;
K = 2*K12*K23*K13 - K22*K13^2 - K11*K23^2;
K = K33 + K / (K11*K22 - K12^2);
Nmn(m,n) = K / am^2; % buckling load for m, n halfwaves
end

- 231-
4 Mechanical analysis of composite laminates

Graphite−Epoxy antisymmetric cross−ply laminate buckling loads


4.5
Adimensional buckling load Nad = N (b / π) / D22

4 1x[0,90]
2

2x[0,90]
3.5 Specially orthotropic solution 4x[0,90]
∞x[0,90]
3

2.5

1.5

0.5

0
0.5 1 1.5 2 2.5 3
Aspect ratio r = a / b

Figure 85: Figure 1 of Exercise 54. Adimensional critical buckling loads for several
antisymmetric stacking sequences for cross-ply Graphite-Epoxy laminates.

end
N(ir) = min(min(Nmn)); % critical = minimum in n and m
Nad(ir) = N(ir)*(b/pi)^2 /D(2,2); % adimensional critical load
end

Exercise 54 Use the code of the previous example to plot the adimensional buckling
loads for Graphite-Epoxy and Glass-Epoxy with stacking sequences [0, 90], [(0, 90)2 ],
[(0, 90)4 ], [(0, 90)n ], where n is a large number. Note this last stacking sequence
represents, essentially, the associated specially orthotropic limit (with B11 → 0 for
a fixed total plate thickness h).

Solution: The solution is shown in Figure 85. It can be seen that after a rea-
sonable number of plies, the curves coalesce to an asymptotic curve, i.e. to what
we have called the specially orthotropic limit. For these cases, the simpler specially
orthotropic solution (for the laminate [(0, 90)n , 0]) may be considered as an excel-
lent approximation. Figure 86 shows the case of Glass-Epoxy, where it is seen that
the specially orthotropic limit is approached faster. The reader can check that these
asymptotic behaviors are also obtained using the code of Exercise 44 for the respective
many-layered specially orthotropic (symmetric) laminates.

Buckling of antisymmetric angle-ply laminated plates

In angle-ply antisymmetric laminated plates the stiffnesses A14 = A24 = D14 =


D24 = 0, but there are nonvanishing bending-extension coupling terms B14 , B24 .

- 232-
4.5 Buckling analysis

Glass−Epoxy antisymmetric laminate buckling loads


6

22
= N (b / π)2 / D
5.5 1x[0,90]
2x[0,90]
5 4x[0,90]
∞x[0,90]
ad
Adimensional buckling load N

4.5

3.5

2.5
0.5 1 1.5 2 2.5 3
Aspect ratio r = a / b

Figure 86: Figure 2 of Exercise 54. Adimensional critical buckling loads for several
antisymmetric stacking sequences for cross-ply Glass-Epoxy laminates

Then, for uniaxial Nx = −N̄ buckling load, the field equations are, see Exercise 30

∂ 2 u0x ∂ 2 u0x ∂ 2 u0y ∂ 3 u0z ∂ 3 u0z


A11 + A44 + (A12 + A44 ) − 3B 14 − B 24 =0 (815)
∂x2 ∂y 2 ∂x∂y ∂x2 ∂y ∂y 3

∂ 2 u0x ∂ 2 u0y ∂ 2 u0y ∂ 3 u0z ∂ 3 u0z


(A12 + A44 ) + A44 + A22 − B 14 − 3B 24 =0 (816)
∂x∂y ∂x2 ∂y 2 ∂x3 ∂x∂y 2

∂ 4 u0z ∂ 4 u0 ∂ 4 u0z
D11 4
+ 2 (D12 + 2D44 ) 2 z 2 + D22 ...
∂x ∂x ∂y ∂y 4
! !
∂ 3 u0x ∂ 3 u0y ∂ 3 u0x ∂ 3 u0y ∂ 2 u0z
−B14 3 2 + − B 24 + 3 + N̄ =0 (817)
∂x ∂y ∂x3 ∂y 3 ∂x∂y 2 ∂x2

In this case also the displacement components u0x , u0y , u0z are coupled. The
solution for simply supported case S3 in Eq. (621) is —see Section 4.4.3
  mπx   nπx 
 u0 = u0 sin cos
 x xmn


 a b

 0  mπx   nπx 
uy = u0ymn cos sin (818)
 a b

  mπx   nπx 
 0 0
 uz = uzmn sin sin


a b

The buckling loads are also given by Eq. (812), but considering now the following

- 233-
4 Mechanical analysis of composite laminates

Graphite−Epoxy antisymmetric Angle−ply laminate buckling loads


12
Adimensional buckling load Nad = N (b / π)2 / D22

1x[45,−45]
2x[45,−45]
10 4x[45,−45]
∞x[45,−45]

0
0.5 1 1.5 2 2.5 3
Aspect ratio r = a / b

Figure 87: Figure of Exercise 56. Adimensional critical buckling loads for several
antisymmetric stacking sequences for angle-ply Graphite-Epoxy laminates

coefficients

 K11 = A11 cxx + A44 cyy




 K12 = (A12 + A44 ) cxy


 K = −3B14 cxxy − B24 cyyy
13
(819)


 K22 = A44 cxx + A22 cyy




 K23 = −B14 cxxx − 3B24 cxyy

K33 = D11 cxxxx + 2 (D12 + 2D44 ) cxxyy + D22 cyyyy

Exercise 55 Write a Matlab code to compute the critical buckling load for antisym-
metric angle-ply laminates.
Solution: The code is a minimum straightforward change over that of Exercise 53.
In fact, they can be readily generalized using the equations of Section 4.4.3.

Exercise 56 Use the code of Exercise 55 to plot the adimensional buckling loads for
Graphite-Epoxy with stacking sequences [45, −45], [(45, −45) 2 ], [(45, −45)4 ], [(45, −45)n ],
where n is a large number. This last stacking sequence is the specially orthotropic
limit (with B14 , B24 → 0 for a fixed total plate thickness h).
Solution: The solution is shown in Figure 87.

Exercise 57 Use the code of Exercise 55 to plot the influence of the fiber angles on
the buckling load. Use a square plate of a = b = 1 and a total thickness of h = a/500.
Solution: The solution is shown in Figure 88.

- 234-
4.6 Vibration analysis

Graphite−Epoxy antisymmetric angle−ply laminate buckling loads


3000
1x[α,−α]
2x[α,−α]
2500 4x[α,−α]
∞x[α,−α]
2000
Buckling load N

1500

1000

500

0
0 10 20 30 40 50 60 70 80 90
Fiber angle respect to load direction α

Figure 88: Figure of Exercise 57. Influence of the angle and number of plies on the
critical buckling load of angle-ply laminates.

4.6 Vibration analysis


The (pure) vibration analysis of a plate is very similar to the (pure) buckling analysis
of a plate. In this case, instead of considering second-order geometric terms in the
equilibrium equations, we consider the inertia terms; i.e. instead of including in the
analysis the second-order geometric term present in K33 in Eq. (604), we include the
mass matrix. This difference is directly motivated by the different effective loads to
be considered in both cases: recall the effective vertical load expression of the field
equations, Eq. (575)3

 ∂ 2 u0z ∂ 2 u0z ∂ 2 u0z

 Buckling: q z = N x + 2N xy + N y

 ∂x2 ∂y∂x ∂y 2
!  2 0  (820)

 0 0 1 ∂ ü0x ∂ ü0y 2 ∂ üz ∂ 2 ü0z
 Vibration: qz = −mz üz − mz ∂x + ∂y + mz ∂x2 + ∂y 2

where for vibration analysis of thin plates (h small) it is frequent to consider only
relevant the term −m0z ü0z . Obviously, if the in-plane loads are important, then the
second-order “in-plane” terms should also be included in the vibration analysis. For
example, it’s a well-known fact that in a guitar string, a substantial tensile load
increases the vibration frequencies of the instrument. We discuss this effect below.
The proposed u0z (x, y) solutions (modes) are frequently the same obtained for
buckling, but the coefficients are allowed to vary with time and an additional separa-
tion of variables is needed. The buckling loads are substituted by natural frequencies
in the eigenvalue problem. However in this case we are not only interested on the
lowest eigenvalue (i.e. the critical buckling load), but on many frequencies. In this

- 235-
4 Mechanical analysis of composite laminates

case it is also important to know the vibration mode associated to each frequency,
because depending on the shape of such mode, we may be exciting it or not. The
number of frequencies and modes of interest depend on the frequency content of the
load exciting the structure. Usually the lowest modes are of major interest for most
analysis, but if the load has a given clear frequency, then the modes close to that
frequency are the ones being excited and, hence, of major interest. If a given mode
is excited, the structure may suffer resonance and, hence, sustain stresses several
orders of magnitude greater than the static ones.
The purpose of this section is to analyze modes and frequencies of a laminated
structure in order to gain some insight in the dynamic behavior of such structures.
However, the implications of the frequency content of the laminate in the dynamic
behavior of the structure is out of the scope of these notes. For such studies the
reader may refer to notes and books on structural dynamics.
The examples to be considered here are basically the same as those studied
for buckling behavior. Since the solutions and the mathematical procedure are
essentially the same, fewer details are given.

4.6.1 Vibration of symmetric laminated beams


In the case of symmetric laminated beams the equilibrium equation may be obtained
from Eq. (766) including the inertia effects
1 d4 u0z d2 u0z 0 0
2 0
2 d üz
⋄ dx4 − N x + m ü
z z − m z =0 (821)
D11 dx2 dx2

where dependencies on y have been neglected and only u0z terms have been considered
relevant (thin beams). We have kept the rotary inertia term and the second order
effects of the axial load in this example to study their influence in the response.
In order to compare with isotropic beams, we can substitute
1 Ēx h3
⋄ = = Ēx I¯y (822)
D11 12

where (recall from previous sections) Ēx I¯y represents the equivalent flexural stiffness
of the beam per unit width. Then, applying separation of variables, a solution of
the type
u0z (x, t) = W (x) T (t) (823)
is proposed for every mode of vibration, where W (x) characterizes the mode shape
and T (t) characterizes the modal frequency. Inserting this expression in Eq. (821)
we obtain
Ēx I¯y T W iv) − Nx T W ′′ + m0z T̈ W − m2z T̈ W ′′ = 0 (824)
This equation may be fulfilled under some conditions if we assume for T (t)

T (t) = eiωt = cos (ωt) + i sin (ωt) (825)

where ω represents the angular frequency of the mode of vibration. Therefore


 
Ēx I¯y W iv) − Nx W ′′ − m0z W ω 2 + m2z W ′′ ω 2 eiωt = 0 (826)

- 236-
4.6 Vibration analysis

i.e.
d4 W  d2 W
Ēx I¯y + m 2 2
z ω − Nx − m0z ω 2 W = 0 (827)
dx4 dx2
or
d4 W 2
2d W
K + Jω − M ω2W = 0 (828)
dx4 dx2
where
K = Ēx I¯y ; J = m2z − Nx /ω 2 ; M = m0z (829)
We notice that this differential equation has the same form as the homogeneous
part of Eq. (751) for the Levy-Nadai problem. Using the same procedure as above,
four conjugate roots λ of the characteristic homogeneous equation are found. The
following general solution is devised
X
W (x) = Ai eλi x = U1 sin(λ1 x)+U2 cos(λ1 x)+U3 sinh(λ2 x)+U4 cosh(λ2 x) (830)
i

where the wavenumbers or spatial frequencies of the modes, namely λ1 and λ2 , are
given in terms of the angular frequency ω through
 ω  √ 
 λ21 =
 Jω + J 2 ω 2 + 4KM
2K
 √  (831)
 λ22 =
 ω
−Jω + J 2 ω 2 + 4KM
2K
Then, note that for each frequency ω to be considered in the temporal function
Eq. (825) there exist two associated wavenumbers λ1 and λ2 to be considered in its
spatial counterpart Eq. (830). The values λ1 and λ2 (and, hence, the frequency)
relative to each mode are to be determined from the boundary conditions, as we
show below in an example.
Factoring out ω from the first relation, for example, of Eqs. (831) we obtain
s   
K 1 + ρN
ω = λ21 (832)
M 1 + ρR

where the ratios


1 Nx m2z 2
ρN = and ρR = λ (833)
λ21 Ēx I¯y m0z 1
show the influence of the axial load and the rotary inertia on the vibrating frequency
ω for each mode characterized by λ1 (analogously, λ2 ). Thus, mode shapes with large
wavenumbers have high vibrating frequencies, and vice versa. Aside, it is readily
seen that if Nx and m2z are neglected, then J = 0, λ1 = λ2 = λ in Eq. (830) and
ρN = ρR = 0, so s
r
K Ēx I¯y
ω = λ2 = λ2 (834)
M m0z
Finally, it can also be deduced from Eq. (832) that the application of a tensile load
Nx increases the vibration frequency ω of a given mode. This influence is inversely
proportional to the squared wavenumber λ21 , so for a given ratio Nx /K the influence

- 237-
4 Mechanical analysis of composite laminates

of Nx in ω by means of ρN has a greater effect in modes with small wavenumbers


(i.e. large wavelengths). Thus, the influence of Nx is mainly relevant at the first,
lower frequencies. On the contrary, the expression for ρR shows that the rotary
inertia effect over the frequency of a given mode is to decrease its value. This is the
idea behind the usage of inertia wheels in rotating axes, for example. Since ρR is
directly proportional to λ21 , the importance of inertia effects on ω is significant for
modes with high frequencies (small wavelengths) and small at low frequencies.
The frequencies and mode shapes depend on the boundary conditions. For ex-
ample, if the symmetric-layered beam is simply supported, then

d2 u0z
at x = 0 ⇒ u0z = 0 and =0 (835)
dx2
d2 u0z
at x = a ⇒ u0z = 0 and =0 (836)
dx2
where, taking

W = U1 sin λ1 x + U2 cos λ1 x + U3 sinh λ2 x + U4 cosh λ2 x (837)


′′
W = −λ21 U1 sin λ1 x − λ21 U2 cos λ1 x + λ22 U3 sinh λ2 x + λ22 U4 cosh λ2 x (838)

we obtain the relations

W (0) = U2 + U4 = 0 and W ′′ (0) = −λ21 U2 + λ22 U4 = 0 (839)

i.e. U2 (λ21 + λ22 ) = 0, so U2 = 0 = U4 . On the other beam end

W (a) = U1 sin λ1 a + U3 sinh λ2 a = 0 (840)


W ′′ (a) = −λ21 U1 sin λ1 a + λ22 U3 sinh λ2 a = 0 (841)

i.e. U3 sinh(λ2 a)(λ21 + λ22 ) = 0, so U3 = 0 (because we are interested in wavenumbers


and frequencies greater than zero), thereby

U1 sin λ1 a = 0 ⇒ λ1n = , n = 1, 2, 3, ... (842)
a
The possible mode shapes are then
nπx
Wn (x) = U sin (843)
a
where the amplitude U = U1 remains to be determined. Note the vibration mode
for n = 1 is the usual vibration mode of a simply supported beam (a half sine wave).
The natural frequency of the mode n is
s
 nπ 2  Ē I¯   1 + ρ 
x y N
ωn = (844)
a m0z 1 + ρR

Obviously, for other boundary conditions, other different modes and frequencies may
be obtained.

- 238-
4.6 Vibration analysis

Exercise 58 Obtain the natural frequencies and mode shapes of a bi-clamped beam
neglecting Nx and m2z .
Solution: We leave this exercise to the reader and simply quote the resulting non-
linear equation to be solved (you may use, for example, the Matlab function fzero to
solve it numerically)

cos(λa) cosh(λa) − 1 = 0 ⇒ λn a = 4.730, 7.853, 10.996, 14.137, ... (845)

with
1 1
U1 = −U3 = ; − U2 = U4 = (846)
sin (λa) − sinh (λa) cos (λa) − cosh (λa)
in Eq. (830). The vibration frequency of each mode is given by Eq. (834).

Exercise 59 Obtain the natural frequencies and mode shapes of a clamped-free beam
neglecting Nx and m2z .
Solution: We leave this exercise to the reader and simply quote the result

cos(λa) cosh(λa) + 1 = 0 ⇒ λn a = 1.875, 4.694, 7.855, 10.996, ... (847)

with
1 1
U1 = −U3 = ; − U2 = U4 = (848)
sin (λa) + sinh (λa) cos (λa) + cosh (λa)
in Eq. (830). The vibration frequency of each mode is given by Eq. (834).

Exercise 60 Obtain the natural frequencies and mode shapes of a plate in cylindri-
cal bending with vanishing in-plane loads, neglecting Nx and m2z .
Solution: We refer to Exercise 43, page 218, to show that the only change to con-
3 = Ē I¯ ≡ D̄. Then the natural frequencies are
sider in Eq. (821) is 1/D11 x y
s
2 D̄
ω=λ (849)
m0z

where λ is determined from the boundary conditions.

4.6.2 Vibration of simply supported specially orthotropic plates


The field equation for the deflection of a specially orthotropic thin plate, neglecting
the effect of in-plane loads and rotary inertia effects, is

∂ 4 uz ∂ 4 uz ∂ 4 uz
D11 + 2 (D12 + 2D44 ) + D22 + m0z ü0z = 0 (850)
∂x4 ∂x2 ∂y 2 ∂y 4
Then, as usual, we propose a solution which fulfills the simply supported boundary
conditions, namely
mπx nπy mπx nπy iωt
u0z (x, y, t) = u0zmn (t) sin sin = Uzmn sin sin e = W (x, y) T (t) (851)
a b a b

- 239-
4 Mechanical analysis of composite laminates

or alternatively
T (t) = T1 cos ωt + T2 sin ωt (852)
Substituting this proposal in the equilibrium equation we obtain the following nat-
ural frequency
 
2 1  π 4 m4 m2 n 2 4
ωmn = 0 D11 4 + 2 (D12 + 2D44 ) 2 + D22 n (853)
mz b r r

for the vibrating mode


mπx nπy
Wmn (x, y) = Uzmn sin sin (854)
a b
The reader can check that if rotary inertia is to be considered, one has to simply
substitute the mass per unit plate surface m0z in Equation (853) by the equivalent
mass m# z  
# 0 2 mπ 2  nπ 2
mz = mz + mz + (855)
a b
Natural frequencies are frequently adimensionalized as in the case of buckling loads.
For example, we can take
 2 s 0
ω b mz
ωad =   r =ω (856)
π 2 D22 π D22
b m0z
so
2 D11 m4 2 (D12 + 2D44 ) m2 n2
ωad (m, n) = + + n4 (857)
D22 r 4 D22 r2
Finally, we want to note that the shape of the lowest vibration modes strongly
depend on the laminate properties and on the aspect ratio of the plate. The following
examples show this issue.

Exercise 61 Write a Matlab program to compute the lowest modes of a specially


orthotropic simply-supported plate.
Solution: The code follows

function [ws,ms,ns] = SOplateModes(D,r,nws)


%
% [ws,ms,ns] = SOplateModes(D,r,nws)
% Function to compute the lowest modes of a specially
% orthotropic plate
%
% Input:
% D = moment-curvature plate constitutive matrix
% r = aspect ratio
% nws = number of frequencies to compute
%

- 240-
4.6 Vibration analysis

% Output:
% ws = ordered adimensional angular frequencies
% ms = corresponding m value for the mode
% ns = corresponding n value for the mode
%

nm = nws; nn = nws; % number of modes to compute


w2mn = zeros(nm*nn); % initialize modal frequencies
i = 0;
for m = 1:nm % compute frequencies for each m
for n = 1:nn % idem for each n
i = i + 1;
w2mn(i)= n^4 + D(1,1)/D(2,2)*(m/r)^4 + ...
2*(D(1,2)+2*D(3,3))/D(2,2)*(m*n/r)^2;
mw(i) = m; % auxiliar vectors
nw(i) = n; % auxiliar vectors
end
end
%
% order the modes and return only the required ones
%
[w2o,io] = sort(w2mn); % order squared frequencies
ws = w2o(1:nws).^0.5; % frequencies are sq. root
for j=1:nws % order indices
ms(j) = mw(io(j));
ns(j) = nw(io(j));
end
%
return

Exercise 62 Use the code of Exercise 61 to compute the lowest six frequencies of
an isotropic square plate and then plot the respective modes, given by Eq. (854).
Solution: The frequencies are obtained using a laminate with D22 = D11 , D12 =
νD11 and 2D33 = (1 − ν) D11 . The mode shapes are shown in Figure 89. Note
that the second and third modes, as well as the fifth and sixth modes have the same
frequencies and equivalent mode shapes, respectively, because x and y coordinates are
interchangeable in an isotropic plate.

Exercise 63 Plot the six lowest vibration modes of a Graphite-Epoxy square plate
with stacking sequence [0, 0, 0̄]s . Repeat the plots for stacking sequence [0, 90, 0̄]s .
Comment the results.
Solution: The solutions are shown in Figures 90 and 91. Note that, since in the
first figure the fibers are aligned in the x direction, the x direction is much stiffer
than the y-direction and the modes with m = 1 have much lower frequencies than
the equivalent modes with n = 1 (compare the second and fifth modes in Figure 90).
Aside, Figure 91 shows that when significant reinforcement is introduced in the y

- 241-
4 Mechanical analysis of composite laminates

Figure 89: Figure of Exercise 62. Vibration modes and adimensionalized frequencies
of an isotropic square plate.

- 242-
4.6 Vibration analysis

direction, the previous observation is no longer true. Furthermore, it is instructive to


compare Figure 91 to Figure 89 for the isotropic case, specially the sixth mode. This
comparison shows that it is no obvious which are the lowest modes in a composite
plate because all structural stiffness terms play their different role.

Exercise 64 Plot the six lowest frequency modes for a Graphite-Epoxy rectangular
plate with aspect ratio r = 2.
Solution: The result is shown in Figure 92. Results are to be compared to those of
Figure 91.

Exercise 65 Create a MATLAB code to plot the modal adimensional frequencies


of a Graphite-Epoxy plate as a function of the aspect ratio r.
Solution: The code is left to the reader. The plot is shown in Figure 93.

4.6.3 Vibration of antisymmetric laminates


The vibration analysis of antisymmetric laminates also follows the same lines as the
buckling analysis of such plates. However, in this case we are not only interested
in the lowest frequency, but in a given number of frequencies. The modal solution
for free vibrations may be obtained in general form from Equation (739) without
the consideration of external loads (at the right-hand side). Neglecting m1z and m2z
effects and also second-order geometrical effects, it may be written as
  0   0  0   
K11 K12 K13   uxmn   mz 0 0  üxmn 
    0 
  0  0  0
 K12 K22 K23  uymn +  0 mz 0  üymn = 0 (858)

 0 
 0

 0 
    
K13 K23 K33 uzmn 0 0 mz üzmn 0

Assume now that the coefficients u0mn , which include the temporal dependence, are
of the form 

 u0xmn (t) = Uxmn sin (ωt)


u0ymn (t) = Uymn sin (ωt) (859)

 0
 uzmn (t) = Uzmn sin (ωt)

Then
       
K11 K12 K13 1 0 0  Uxmn
 
  0 
 
  2 0 
 K12 K22 K23  − ω mz  0 1 0   Uymn = 0 (860)

 
    
K13 K23 K33 0 0 1 Uzmn 0

or compactly 
[K] − ω 2 m0z [I] {U mn } = {0} (861)
which is an eigenvalue problem for each {m, n} pair.
It should be already known that an analogous matrix equation can be derived
for the specially orthotropic case just analyzed above, where care must be taken

- 243-
4 Mechanical analysis of composite laminates

Figure 90: First figure of Exercise 63. Vibration modes and adimensionalized fre-
quencies for a square plate of Graphite-Epoxy with stacking sequence [0, 0, 0̄]s .

- 244-
4.6 Vibration analysis

Figure 91: Second figure of Exercise 63. Vibration modes and adimensionalized
frequencies for a square plate of Graphite-Epoxy with stacking sequence [0, 90, 0̄]s .

- 245-
4 Mechanical analysis of composite laminates

Figure 92: Figure of Exercise 64. Vibration modes of a Graphite-Epoxy plate with
stacking sequence [0, 90, 0̄]s and aspect ratio r = 2.

- 246-
4.6 Vibration analysis

Graphite−Epoxy [0,90,0,90,0]
20
m=1,n=1
18 m=2,n=1
m=3,n=1
Adimensional frequency ωad 16
m=1,n=2
14 m=2,n=2
m=3,n=2
12 m=1,n=3
m=2,n=3
10 m=3,n=3
8

0
0 0.5 1 1.5 2 2.5 3
Plate acpect ratio r = a/b

Figure 93: Figure of Exercise 65. Adimensional frequencies of a Graphite-Epoxy


plate with stacking sequence [0, 90, 0̄]s as a function of the aspect ratio.

when proposing the three-dimensional displacement field u0 in terms of Fourier-type


functions because they must satisfy the field equations and the boundary conditions.
It can be shown that for this type of symmetric laminates K13 = K23 = 0 and the
third equation is uncoupled from the rest, yielding
 K33
K33 − ω 2 m0z Uzmn = 0 ⇒ ω2 = (862)
m0z

Then, Equation (853) immediately follows.

Vibration of antisymmetric cross-ply laminates


The governing equations are obtained as usual. Neglecting the effect of dynamic
in-plane loads and inertia effects distinct from m0z , they reduce to

∂ 2 u0x ∂ 2 u0x ∂ 2 u0y ∂ 3 u0z


A11 + A44 + (A12 + A44 ) − B 11 =0 (863)
∂x2 ∂y 2 ∂x∂y ∂x3

∂ 2 u0x ∂ 2 u0y ∂ 2 u0y ∂ 3 u0z


(A12 + A44 ) + A44 + A22 + B 11 =0 (864)
∂x∂y ∂x2 ∂y 2 ∂y 3

∂ 4 u0z ∂ 4 u0z ∂ 4 u0z


D11 + 2 (D12 + D 44 ) + D22
∂x4 ∂x2 ∂y 2 ∂y 4
∂ 3 u0x ∂ 3 u0y 2 0
0 ∂ uz
− B11 + B 11 + m z =0 (865)
∂x3 ∂y 3 ∂t2

- 247-
4 Mechanical analysis of composite laminates

As in the buckling problem, the following solution (including now the temporal
dependence) fulfills the simply supported boundary conditions S2 given in Eq. (621)
  mπx   nπx 
 u0 =U cos sin exp (iωt)
 xmn
 x

 a b

 0  mπx   nπx 
uy = Uymn sin cos exp (iωt) (866)
 a b

  mπx   nπx 
 0
 uz = Uzmn sin sin exp (iωt)


a b

and yields, upon substitution in the equilibrium equations, the following eigenvalue
problem—see Eq. (860)
       
K11 K12 K13 0 0 0 
 Uxmn 
   0 
  2 0 
 K12 K22 K23  − ω mz  0 0 0  Uymn = 0 (867)

 
    
K13 K23 K33 0 0 1 Uzmn 0
Operating in terms of matrix blocks
    
K αβ K α3 U αmn 0
= (868)
K Tα3 K33 − ω 2 m0z Uzmn 0
From the first block equation

K αβ U αmn + K α3 Uzmn = 0 ⇒ U αmn = −K −1


αβ K α3 Uzmn (869)

which substituted in the second equation yields


 
K33 − ω 2 m0z − K Tα3 K −1
αβ K α3 Uzmn = 0 (870)

Hence, for Uzmn 6= 0, the frequency of this mode is given by


 2 − K K2 
2 1 2K12 K23 K13 − K22 K13 11 23
ωmn = 0 K33 + 2 (871)
mz K11 K22 − K12
where Kij are the same as those defined in Section 4.5.3, i.e. Equation (813). These
values are frequently adimensionalized through

2 2 m0z b4
ωad = ωmn (872)
D22 π 4

Vibration of antisymmetric angle-ply laminates


We leave to the reader the task of verifying that Equation (871) can be used to com-
pute the natural frequencies for this type of laminates for S3 boundary conditions
if we simply replace the previous Kij coefficients by the ones given in Eq. (819).
Because angle-ply laminates have a strongly varying value D22 , it is frequent to
adimensionalize the frequencies using the Young modulus of the matrix material Ey
instead, i.e.
0 4
2 2 mz b 2 D22
ωad = ω 3 4
= ωad (873)
2
Ey h π Ey h3

- 248-
4.6 Vibration analysis

Exercise 66 Write a Matlab program to obtain the lowest frequencies of a rectangu-


lar plate of some of the types discussed above: simply supported specially orthotropic
laminate, antisymmetric cross-ply laminate with S2 boundary conditions or anti-
symmetric angle-ply laminate with S3 boundary conditions.
Solution: The code follows. It is left to the reader to verify the specially orthotropic
case using the MATLAB function of Exercise 61.

function [ws,ms,ns] = NavierplateModes(A,B,D,r,nws)


%
% [ws,ms,ns] = NavierplateModes(A,B,D,r,nws)
% Function to compute the lowest modes of a Navier solution
%
% Input:
% A = Extension constitutive matrix of plate
% B = Extension-bending coupling constitutive matrix of plate
% D = Moment-curvature constitutive matrix of plate
% r = aspect ratio
% nws = number of frequencies to compute
%
% Output:
% ws = ordered adimensional angular frequencies
% ms = corresponding m value for the mode
% ns = corresponding n value for the mode
%

nm = nws; nn = nws; % number of modes to compute


w2mn = zeros(nm*nn); % initialize modal frequencies
a = 1; b = a/r; % plate dimensions
i = 0; % frequency counter
for m = 1:nm % compute frequencies for each m
for n = 1:nn % idem for each n
i = i + 1;
am = m * pi / a; % some coefficients, see formula
bn = n * pi / b;
K11= A(1,1)*am^2 + A(3,3)*bn^2;
K22= A(2,2)*bn^2 + A(3,3)*am^2;
K33= D(1,1)*am^4 + 2*(D(1,2)+2*D(3,3))*am^2*bn^2 + D(2,2)*bn^4;
K12= (A(1,2) + A(3,3))*am*bn;
K13= -(B(1,1)*am^3 + 3*B(1,3)*am^2*bn + B(2,3)*bn^3);
K23= -(B(1,3)*am^3 + 3*B(2,3)*am*bn^2 + B(2,2)*bn^3);
K = 2*K12*K23*K13 - K22*K13^2 - K11*K23^2;
K = K33 + K / (K11*K22 - K12^2);
w2mn(i)= b^4/pi^4/D(2,2)*K;
mw(i) = m; % auxiliar vectors
nw(i) = n; % auxiliar vectors
end

- 249-
4 Mechanical analysis of composite laminates

end
%
% order the modes and return only the required ones
%
[w2o,io] = sort(w2mn); % order squared frequencies
ws = w2o(1:nws).^0.5; % frequencies are sq. root
for j=1:nws % order indices
ms(j) = mw(io(j));
ns(j) = nw(io(j));
end
%
return

Exercise 67 Write a Matlab script to plot and visualize in a video the lowest vi-
bration modes of laminated plates computed in the previous exercise.
Solution: The code follows. “Play” with it and enjoy the movies! (Note each mode
vibrates with its own frequency).

lamina_name = ’Graphite_Epoxy’; % lamina material type (example)


laminate = [0,90,0,90,0,90,0]; % stacking sequence (example)
a = 1; r = 1; t = a/100; mz0 = 1; % some plate data
[~,A,B,D,ABBD,~,~,~,~] = laminatestiff(lamina_name,laminate,t);
[ws,ms,ns] = NavierplateModes(A,B,D,r,6);
movf = mplot2Dfourier(r,ws,ms,ns,2,’\omega_{ad} = ’);

function movf = mplot2Dfourier(r,w,m,n,nsc,labt)


%
% Function to plot movie plate 2D fourier terms (i.e. sines)
%
% Input:
% r = plate aspect ratio
% w = frequencies of modes (array)
% m = number of halfwaves in x of modes (array)
% n = number of halfwaves in y of modes (array)
% nsc = number of subplots in width of figure
% labt = generic title for each subplot
%
% Output: enjoy the movie
%

x = [0:0.05:1]; % dimension and divisions of plate


y = [0:0.05/r:1/r]; % idem in y-axis
nd = length(x); % number of divisions in plate
np = length(m); % number of plots
nsr = np/nsc; % number of subplots in row
%

- 250-
4.6 Vibration analysis

for ip = 1:np % loop to create z-coordinates


for i = 1:nd
for j = 1:nd % z-scale maximum to 25%
Z(i,j,ip) = 0.25*sin(m(ip)*pi*x(i))...
*sin(n(ip)*pi*y(j)*r);
end
end
end
%
h = figure; % new figure
set(gcf,’Position’,[624 074 672 804]) % nice frame
set(gcf,’renderer’,’Zbuffer’) % less heavy buffer
nf = 100; % number of frames
aviobj = avifile(’D:\plate_modes_movie.avi’) % file of the movie
%
for ifr = 1:nf % for each frame of the movie
ip = 0; % reset plot counter
for i = 1:nsr % for each subplot (row)
for j = 1:nsc % for each subplot (column)
ip = ip + 1;
subplot(nsr,nsc,ip); % points to subplot
iff = cos(0:2*pi/(nf-1)*w(ip)/w(1)...
:2*pi*w(ip)/w(1)); % frames z-scale
z(:,:) = iff(ifr)*Z(:,:,ip); % scale z-coordinates
surf(x,y,z’); % generate mesh
axis equal; % same scale to all axis
xlabel(’x’); ylabel(’y’);
title([labt,num2str(w(ip))]); % label frequency
set(gca,’xlim’,[0,1],’ylim’,[0,1],’zlim’,[-0.25,0.25]);
axis off; % nicer if axis are not shown
end
end
movf(ifr) = getframe(h); % adds frame to the movie
aviobj = addframe(aviobj,movf(ifr));
end
%
aviobj = close(aviobj); % closes movie
%
end

Exercise 68 Use the Matlab codes of Exercises 66 and 67 to plot the six lowest
modes of a Graphite-Epoxy square antisymmetric [0,90] cross-ply laminate.
Solution: The plot is shown in Figure 94.

Exercise 69 Use the Matlab codes of Exercises 66 and 67 to plot the six lowest
modes of a Graphite-Epoxy square antisymmetric [(0, 90)4 ] cross-ply laminate.

- 251-
4 Mechanical analysis of composite laminates

Figure 94: Figure of Exercise 68. Vibration modes of an antisymmetric Graphite-


Epoxy [0, 90] laminate.

- 252-
4.6 Vibration analysis

Graphite−Epoxy [0,90]
6
m=1,n=1
m=2,n=1
5 m=3,n=1
Adimensional frequency ωad m=1,n=2
m=2,n=2
4 m=3,n=2
m=1,n=3
m=2,n=3
3 m=3,n=3

0
0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Plate acpect ratio r = a/b

Figure 95: Figure of Exercise 70. Variation of the adimensional frequencies


ωad (m, n) with the plate aspect ratio for a Graphite-Epoxy square antisymmetric
[0,90] cross-ply laminate.

Solution: This exercise is left to the reader.

Exercise 70 Create a Matlab script to plot the adimensional frequencies for m =


1, 2, 3 and n = 1, 2, 3 for varying plate aspect ratios r. Show the results for a Graphite-
Epoxy square antisymmetric [0,90] cross-ply laminate.
Solution: The resulting plot is shown in Figure 95.

Exercise 71 Show the results of the previous Exercise for a Graphite-Epoxy square
antisymmetric [(0, 90)4 ] cross-ply laminate.
Solution: The resulting plot is shown in Figure 96. It is interesting to note that
as the number of layers increase, the specially orthotropic case is approached, i.e.
compare with Figure 93. It is left to the reader to explain why this happens in terms
of how A, B, D evolve.

Exercise 72 Create a Matlab script to plot the adimensional frequencies ωad2 for
antisymmetric angle-ply laminates with varying ply angles and different number of
layers.
Solution: The result is shown in Figure 97. Two effects can be seen: The first one
is that as the number of plies increases, the solution approaches the corresponding
specially orthotropic solution. The second one is that the third and fourth mode
shapes are different for different angles.

- 253-
4 Mechanical analysis of composite laminates

Graphite−Epoxy 4x[0,90]
6
m=1,n=1
m=2,n=1
5 m=3,n=1
Adimensional frequency ωad

m=1,n=2
m=2,n=2
4 m=3,n=2
m=1,n=3
m=2,n=3
3 m=3,n=3

0
0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Plate acpect ratio r = a/b

Figure 96: Figure of Exercise 71. Variation of the adimensional frequencies


ωad (m, n) with the plate aspect ratio for a Graphite-Epoxy square antisymmetric
[(0, 90)4 ] cross-ply laminate.

Exercise 73 Use the Matlab codes of Exercises 66 and 67 to plot the six lowest
modes of Graphite-Epoxy square antisymmetric [(−α, α)2 ] angle-ply laminates, where
α = 10, 20, 30.
Solution: The plots are shown in Figures 98 to 100. It is seen that, as commented
in Exercise 72, the ordering of modes (i.e. which ones have the lowest frequencies)
changes with the laminae angles. Note that the first and second mode shapes are the
same for the tree figures but the third and fourth ones change.

4.7 Edge interlaminar stresses


Interlaminar stresses computed in the previous sections, for example using Equa-
tion (677) assume Saint-Venant hypothesis that local effects dissipate at a sufficient
distance from the source of such effects. This hypothesis is used, for example, to
consider a concentrated load as a point load instead of the actual load distribution
over a small surface. The stresses and strains are accurate at a distance from the
load, but not nearby it.
When computing interlaminar stresses using Equation (677) we assumed that
we were far enough from point loads and also from boundaries. How much is “far
enough” depends on the problem at hand, but as a rule of thumb we may assume that
we are far enough at 2-3 times the local characteristic dimension. Then, interlaminar
stresses computed as in the previous sections may be considered rather accurate at
distances from the boundaries of two-three times the thickness of the laminate, but
not at the boundaries.

- 254-
4.7 Edge interlaminar stresses

Fundamental frequency 2nd lowest frequency


2.8 6

5.5
2.6 ∞x[−α,α]
5
2.4
4.5
2.2
ωad2

ad2
4

ω
2
3.5
2x[−α,α]
1.8
3
[−α,α]
1.6 2.5

1.4 2
0 10 20 30 40 0 10 20 30 40
Ply angles α Ply angles α

rd th
3 lowest frequency 4 lowest frequency
7 11

6.5 10

6
9 Change of
5.5 modal shapes
Change of 8
modal shape
ωad2

ad2

5
ω

7
4.5
6
4

3.5 5

3 4
0 10 20 30 40 0 10 20 30 40
Ply angles α Ply angles α

Figure 97: Figure of Exercise 72. Lowest frequencies for an antisymmetric angle-ply
laminate as a function of the ply angles.

- 255-
4 Mechanical analysis of composite laminates

Figure 98: Figure of Exercise 73. Vibration modes of a Graphite-Epoxy antisym-


metric [(−10, 10)2 ] angle-ply square laminate.

- 256-
4.7 Edge interlaminar stresses

Figure 99: Figure of Exercise 73. Vibration modes of a Graphite-Epoxy antisym-


metric [(−20, 20)2 ] angle-ply square laminate.

- 257-
4 Mechanical analysis of composite laminates

Figure 100: Figure of Exercise 73. Vibration modes of a Graphite-Epoxy antisym-


metric [(−30, 30)2 ] angle-ply square laminate.

- 258-
4.7 Edge interlaminar stresses

Figure 101: Edge delamination

Furthermore, the Kirchhoff plate kinematic equations may be considered an ad-


equate model far from the borders, but at the boundary edges the problem is truly
three-dimensional and the theory of elasticity must be employed, i.e. the beam
and plate theories are not valid to capture these local effects. Of course this is
also true for isotropic materials, but experience dictates that no special problem is
usually found in isotropic materials due to these stresses since material constants
are the same throughout the structure. However, these stresses produce in compos-
ite laminates the important effect of edge delamination because different layers have
different apparent properties and the presence of additional interlaminar stresses be-
comes necessary in order to guarantee compatibility of deformations. These stresses
are usually of no concern inside the laminate, but at boundary edges they may be
high enough to cause that laminae separate at borders weakening the laminate, see
Figure 101.
Since edge effects are local three-dimensional effects, the solution can only be
found solving the 3D elasticity problem. Analytical solutions are virtually impossi-
ble, so numerical ones are to be obtained in order to study the phenomena. Then,
finite elements are a useful tool to obtain such solutions. However it is important to
note that since we are aiming at local 3D effects, continuum solid elements must be
used instead of plate or shell elements. Furthermore, since strong stress gradients
are to be expected, fine meshes with high order interpolation functions are desirable
at those locations.
As examples to analyze the edge interlaminar stresses consider the laminate with
the symmetric cross-ply stacking sequence shown in Figure 102. For each unidirec-
tional fiber-reinforced ply, the fiber direction is much stiffer than the transversal
direction. Hence, as shown in Figure 102, the intermediate layers would contract
in transverse direction more than the outer layers because the relation between the
Poisson coefficients is (c.f. Eq. (141))

Ex
νxy = νyx ≫ νyx (874)
Ey

- 259-
4 Mechanical analysis of composite laminates

Figure 102: Edge interlaminar effects in cross-ply laminates. Left: Different Poisson
effects for the inner and outer laminae. Right: Stresses inside the laminate and at
the border of the laminate (where σyy must be zero).

In order to guarantee compatibility, a compressive σyy stress must exist inside the
laminate in the outer layers (far enough the edges) and a tensile σyy stress must exist
inside the laminate in the inner layers, see Figure 102. These stresses, being dis-
continuous between adjacent laminae, can be computed with the classical laminated
plate theory introduced in Section 3.2 (the reader can use the Matlab code of Exer-
cise 14 to verify these results). However, since there is no load in the y−direction,
the σyy stresses at the edge outside faces must be zero. Equilibrium of a differential
element in the y-direction is then fulfilled by means of a σyz at the top surface of
the differential element. To further assess this phenomena, we can employ finite
elements. Consider the finite element mesh shown in Figure 103 with corresponds
to the laminate given in Figure 102 with dimensions a × b × h with a = b = 4h and
h = 4t.
The specimen is subjected to a tensile load Nx (we actually have prescribed
displacements in x to that face). This mesh is finer at the edge under study, and
of course solid continuum elements must be used to study the local effect. Accord-
ingly, three-dimensional elastic properties are defined for each lamina. Since strong
gradients are to be expected, 27 node brick elements have been used employing
3 × 3 × 3 = 27 stress integration points per element, which means that the accuracy
is much higher than using the same mesh with trilinear elements. The stress band
plots will be given for the zone highlighted in Figure 103, which corresponds to one
of the inner plies in Figure 102. As can be observed, each ply has been discretized us-
ing 5 elements in thickness, so there are 5 × 3 = 15 stress integration points through
the thickness of each lamina. The deformed shape may be seen in Figure 104 and
it is verified that it responds to the thoughts given in Figure 102. In Figure 105 it
is shown the stress distribution of σyy , and it can be observed that it progressively
goes from a tensile (positive) value to zero when approaching the border y = b/2. It
can be also seen that the edge effect can be neglected at distances of about h/2 to
h from the border. We want to note that stresses in finite elements are computed
inside the elements at the integration points and then, can be extrapolated to the
nodes using a handy rule in order to perform the plots or, as done in these Figures,
simply pass the values at the integration points to the corresponding nodes. Each

- 260-
4.7 Edge interlaminar stresses

Figure 103: Mesh to compute edge interlaminar stresses and zone used to plot the
results. Elements employed are 27-node brick elements with full 3 × 3 × 3 stress
integration. Four plies are considered. The results are shown in one quarter of one
of the intermediate laminae as highlighted in the figure with a different colour.

method has its advantages and drawbacks: the values of the former are not truly
computed stresses, but the values of the latter are plotted not at their true location.
When the computation is accurate, differences are negligible. When gradients are
very strong and the computation looses accuracy, discontinuities or saw-teeth-type
band may appear as it happens for example at one of the corner lines of Figure 105.
In Figure 106 it is seen that the σyz stresses vanish inside the laminate (as the
Laminated Plate Theory predicts) and increase when approaching the border of the
laminate (as indicated in Figure 102). Of course they also have to be zero at the
outside face (y = b/2) to satisfy the stress-free boundary condition, and also at the
bottom of that lamina (z = 0) because of symmetry considerations with respect to
the middle plane of the laminate. In this case the edge effect may also be neglected
at a distance about h/2 to h. Furthermore, as deduced from Figure 102, σyy and
σyz may introduce a moment unbalance that can only be equilibrated through a σzz
stress contribution. Figure 107 shows that σzz stress distribution. It can be seen
that z−compression is found near the border, but tensile stresses also appear at a
distance of about h/2 from the border. We remark that tensile σzz stresses are a
possible cause of delamination. Obviously, greater tensile stresses σzz may appear
near the laminate edges for other loading cases and/or plies layup.
In order to have some insight on the edge interlaminar stresses for angle-ply
laminates, consider the [30, −30]s angle-ply laminate of Figure 108. If the laminate
is stretched in the x−direction, a possible unrestrained-laminae deformation is the
one shown in Figure 108, where laminae undergo their respective shear deforma-
tions (in addition to axial extension and transverse contraction). Since laminae are
bonded, compatibility of deformation must hold between adjacent laminae, which is
now achieved by means of the corresponding shear stresses σxy within the laminae,
far from edges, which can be accurately predicted by the laminated plate theory.

- 261-
4 Mechanical analysis of composite laminates

Figure 104: Deformed mesh for the cross-ply example.

Figure 105: σyy in-plane stress component in the 0 lamina for a cross-ply [90, 0]S
laminate under tensile load in the x−direction.

Figure 106: σyz out-of-plane stress component in the 0 lamina for a cross-ply [90, 0]S
laminate under tensile load in the x−direction.

- 262-
4.7 Edge interlaminar stresses

Figure 107: σzz out-of-plane stress component in the 0 lamina for a cross-ply [90, 0]S
laminate under tensile load in the x−direction.

Figure 108: Edge interlaminar effects in angle-ply laminates. Left: Example where
shear strains (due to apparent coupling coefficients) are present for inner and outer
laminae. Right: Stresses inside the laminate and at the border of the laminate
(where σxy must be zero).

The deformed mesh for this case, including tree-dimensional border effects, is given
in Figures 109 and 110, where the same finite element discretization than for the
previous example has been used.
Thus, (negative) shear σxy stresses develop inside the lamina at −30, as shown
in Figures 108 and 111 for the mesh shown in Figure 103. Note that σxy stresses
vanish at border faces. At the bottom face they vanish more gradually because
the lower adjacent layer is also at −30, but at the top interface they vanish just at
the edge (with a very large gradient). At the x = a/2 outside face, those stresses
also vanish, but inside for x approaching a/2 they become larger at the top layer
just near the edge so that the solid deformation satisfy the displacement condition
being prescribed at that face (see Figures 109 and 110 again). In any case, the
border effects may be neglected for distances greater than h, approximately. Aside,
in Figure 108 it is shown that at the edges, σxy vanish and then equilibrium is
accomplished through σxz (negative) stresses at the top face. Figure 112 shows this
fact. Again, it is seen that whereas inside the laminate the σxz vanish in practice,
when approaching the edge, at distances of about h, they may raise importantly.
Figure 113 shows that σzz tensile stresses appear in this case just at the corners,

- 263-
4 Mechanical analysis of composite laminates

Figure 109: Deformed mesh for an angle-ply laminate [30, −30]S under tensile Nx
load.

Figure 110: Deformed mesh for an angle-ply laminate [30, −30]S under tensile Nx
load. Top view.

Figure 111: σxy in-plane stress component in the −30 lamina for an angle-ply
[30, −30]S laminate under tensile load in the x−direction.

- 264-
4.7 Edge interlaminar stresses

Figure 112: σxz out-of-plane stress component in the −30 lamina for an angle-ply
[30, −30]S laminate under tensile load in the x−direction.

Figure 113: σzz out-of-plane stress component in the −30 lamina for an angle-ply
[30, −30]S laminate under tensile load in the x−direction.

where delaminations may be induced.


The examples given in this section are just two typical examples to illustrate the
presence of edge interlaminar stresses which explain edge delamination. The values,
distributions and signs of these stresses may vary strongly depending on the material
properties, stacking sequence, boundary conditions and type of loading. However, we
can conclude from these simple examples that this is an important issue in composite
laminates which is caused by different adjacent laminae properties and that in order
to properly assess them, continuum solid elements with fine meshes and high order
interpolation must be used in the zone of interest. Hence, in the design of laminates,
stacking sequences which favor similar apparent lamina properties and compatible
deformations between laminae are to be preferred in principle than those which favor
incompatible deformations and different apparent lamina properties. For example a
¯ S laminate is expected to behave better in terms of edge interlaminar
[90, 45, 0, −45]
¯ S laminate because incompatibility between laminae is
stresses than a [90, 0, 45, −45]

- 265-
4 Mechanical analysis of composite laminates

smaller (the difference in fiber angles between laminae is smaller).

4.8 Considerations on fatigue and fracture in anisotropic laminates.


In this section we briefly comment on the more advanced and the still unmature
topics of fatigue and fracture in composite materials. A more detailed discussion on
this topic is fairly complex and, hence, is out of the scope of this notes.

4.8.1 Fracture mechanics


One of the most important applications of composite materials are structural com-
ponents for the aerospace industry. Fatigue and fracture have long been related
to this industry since the de Havilland Comet airplane crashes from 1952 to 1954.
Previously the sink of the Liberty ships during WWII and the explosion of pressure
tanks (as in steam engines) were the first accidents due to fracture. Since then,
Fracture Mechanics has been an important issue, specially in the aircraft industry.
Fracture concepts have been developed for isotropic materials and basic ideas are
somehow applicable to composite materials, but since they are much more involved
for the latter case, we will focus most of our discussion to isotropic materials and
then highlight the difficulties for composites.
The initial known results on Fracture Mechanics are due to Griffith during WWI.
His interest was to understand why glass needed as low as 100 MPa to brake down
if the theoretical stress from atomic bonds should be one hundred times bigger.
He made some experiments (Ref. [1]) and realized that the actual fracture stress
depends on the size of the specimen; that is the reason fibers and ropes have a larger
resistance than bulk materials and why fibers in fiber composites have larger strength
than their bulk counterparts (recall Section 1.2.1 and the tables presented therein).
The reason is that the probability of having large material flaws (cracks) in a fine
wire is substantially less than in a bulk material. Then he √
tested the influence of the
flaw length 2a in the stress at fracture, obtaining that σf 2a was fairly constant.

Stresses near a crack


In Figure 114 the problem of a small crack of length 2a inside a large (“infinite”)
isotropic elastic material under a tensile uniaxial stress σ∞ far away the borders and
perpendicular to the crack is analyzed (this is the problem originally used by Irwin).
It is obtained that the in-plane stress and displacement fields close to the crack (i.e.
the asymptotic solution, inside the so-called singularity dominated zone, SDZ ) are

σ ∞ πa I
σαβ (r, θ) = √ σ̃αβ (θ) (875)
2πr

σ ∞ πa r I
uα (r, θ; ι) = ūα + √ ũα (θ, ι) (876)
2πr 2G
where r is the distance to the crack tip, θ is the angle with the crack, ι is an equiv-
alence constant which value is ι = 3 − 4ν for plane strain and ι = (3 − ν) / (1 + ν)
for plane stress, ν is the Possion ratio, G is the shear modulus, ūα is a constant for

- 266-
4.8 Considerations on fatigue and fracture in anisotropic laminates.

Figure 114: Crack in a specimen under tensile load and approximate stress distri-
bution for W ≫ a and θ = 0.

I (θ) and ũI (θ, ι) are dimensionless functions of the angle θ for a
direction α and σ̃αβ α
given problem and which can be found in most texts on fracture. For this particular
case, they are found to be
 I
σ̃ (θ) = cos (θ/2) [1 − sin (θ/2) sin (3θ/2)]
 xx


I (θ) = cos (θ/2) [1 + sin (θ/2) sin (3θ/2)]
σ̃yy (877)


 I
σ̃xy (θ) = cos (θ/2) sin (θ/2) cos (3θ/2)

for both plane stress and plane strain, and


( I  
ũx (θ, ι) = cos (θ/2) ι − 1 + 2 sin2 (θ/2)
Plane strain   (878)
ũIy (θ, ι) = sin (θ/2) ι + 1 − 2 cos2 (θ/2)
(  
ũIx (θ, ι) = cos (θ/2) 2 − (1 + ν) cos2 (θ/2)
Plane stress   (879)
ũIy (θ, ι) = sin (θ/2) 2 − (1 + ν) cos2 (θ/2)
I (θ = 0) =
The form of these functions are not relevant here, except for the fact that σ̃xx
I I I
σ̃yy (θ = 0) = 1, σ̃xy (θ = 0) = 0 and that the crack opening function ũy (θ = π, ι) = 2
for plane stress and approximately 2 for plane strain (with ν ≃ 0.5). From Eq. (875)
it is seen that when r → 0 then σαβ → ∞, which is not a real solution because we
know that stresses are not infinity. As we will see later, the inconsistency found
comes from the erroneous assumption of linear elastic behavior. At the crack tip, at
least some local (small scale) plasticity is developed, so stresses are not infinity at
the crack tip, but σY (the yield stress).

- 267-
4 Mechanical analysis of composite laminates

Figure 115: Compact beam experiment (test).

Energy release rate


Griffith found the necessity to approach the problem from a different viewpoint, and
he did it using energetic arguments. He considered that if an edge crack (as in Figure
115) grows by ∆a, then two surfaces of dimension ∆a (times depth) are created
during de process. Surfaces within the material have a larger energy than bulk
material, otherwise surfaces would be created spontaneously because they would
be energetically more favourable. The difference between the energy of a surface
and the same material in bulk is named surface energy S. Conceptually the energy
release rate G (with units of J / m2 ) is twice the surface energy S (two surfaces are
created in the advance of a crack ∆a), i.e. per unit depth:

G ∆a = 2S ∆a (880)

Hence, G represents the energy dissipated per unit surface area being created during
fracture.
Consider the compact beam specimen of Figure 115. Then, the energy release
rate G may be defined as the negative of the variation of the potential energy respect
to the variation of the crack length when the external load P is kept constant

∂Π
G=− (881)
∂a P

For a linear elastic material

Π = W + V = 21 P u − P u = − 21 P u (882)

so
P ∂u
G= (883)
2 ∂a P
The compliance (inverse of the stiffness) of the body is defined as
u
S= (884)
P

- 268-
4.8 Considerations on fatigue and fracture in anisotropic laminates.

(note the difference with S) so



P ∂u 1 2 ∂S
G= = P (885)
2 ∂a P 2 ∂a P

The quantity (∂S/∂a)P can be estimated using an experiment of the type of Figure
115 without the need of considering the compliance of the testing machine. The
quantity ∂S/∂a can be estimated from the compliances S of two specimens with a
small difference ∆a in crack length and the energy release rate computed as

1 2 ∆S 1 u2 ∆S
G= P = (886)
2 ∆a P 2 S 2 ∆a P

Exercise 74 Show that the energy release rate is independent of the compliance of
the testing machine of Figure 115.
Solution: Consider that the machine has a stiffness Km and a compliance Sm =
1/Km . The displacement of the load application point at the specimen is u and the
deformation of the machine is um = ut − u, where ut is the total displacement of the
machine. Then, the potential energy is

Π = W + V = 21 P ut − P ut = − 12 P (u + um ) (887)

Then
   
P ∂ut P ∂ (u + um ) 1 2 ∂S 1 2 ∂Sm
G= = = P + P (888)
2 ∂a P 2 ∂a
P 2 ∂a P 2 ∂a P

However, obviously Sm only depends on the structure of the testing machine and
does not depend on the crack length a, so Eq. (885) is recovered.

Stress intensity factor


The form of Equations (875) and (876) is found in multiple problems involving
cracks. Then it is typical to write these equations as
KI I
σαβ (r, θ) = √ σ̃αβ (θ) (889)
2πr
KI r I
uα (r, θ; ι) = ūα + √ ũα (θ, ι) (890)
2πr 2G
where KI is a constant called stress intensity factor which depends on the problem
at hand and has SI units of N / m3/2 . Then in fracture mechanics the dependence

with distance of stresses is of the order of 1/ r and for the case of displacements

is of r. The constant for symmetric loads respect to the crack tip is referred with
subindex I, and the fracture mode is named mode I, see Figure 116. For the case
given in Eq. (875) of an inner crack of length 2a under a tensile stress σ ∞ far

away the crack is KI = σ ∞ πa. If the crack is an edge crack and has length a,

∞ πa. Infinite aligned cracks of length 2a separated 2b have
then KI = 1.1215σ
∞ √ p
KI = σ πa ρ tan (1/ρ) with ρ = (2b) / (πa). For the case of a specimen with

- 269-
4 Mechanical analysis of composite laminates

Figure 116: Fracture modes. Mode I: axial/normal mode. Mode II: In-plane shear
mode. Mode III: Antiplane shear mode.

finite (versus “infinite”) width Eq. (889) is written using a corrected stress intensity
factor by function Y (a/W ), √
KI = Y σ ∞ πa (891)
where Y depends on the type of crack. For example for the case at hand (inner
crack of length 2a)
a  a  a 2 r  
πa
Y = 1 + 0.256 − 1.152 + ... ≃ sec (892)
W W W W
and for an edge crack of length a
a 0.41  a  18.7  a 2
Y = 1.1215 − √ + √ − ... (893)
W π W π W

For in-plane antisymmetric loads with respect to the crack tip for the cases of
plane stress and plane strain, the fracture mode is mode II (see Fig. 116) and the
expressions have the same form
KII II
σαβ (r, θ) = √ σ̃αβ (θ) (894)
2πr
KII r II
uα = ūα + √ ũα (θ, ι) (895)
2πr 2G
II (θ = 0) = 1 and σ̃ II (θ = 0) =
but KII is the stress intensity factor of mode II and σ̃xy xx
II (θ = 0) = 0. In the case the stresses far away from the inner crack of length 2a are
σ̃yy

only a shear stress τ ∞ in plane stress/strain, then we similarly have KII = τ ∞ πa.
There is another mode for antisymmetric fields in which the loads are parallel to
the planes of fracture, but acting in the z direction (out-of-plane), which is known
as antiplane mode or mode III of fracture. In this case
 
σxz (r, θ) KIII − sin (θ/2)
=√ (896)
σyz (r, θ) 2πr cos (θ/2)

- 270-
4.8 Considerations on fatigue and fracture in anisotropic laminates.

KIII r
uα = √ [4 sin (θ/2)] (897)
2πr 2G

where again KIII = τ ∞ πa for the case of a crack of length 2a in an “infinite” body

with shear τ ∞ in antiplane mode. KIII = τ ∞ πa also in the case of an edge crack
of length a with a shear τ ∞ in antiplane mode in an “infinite” body.
Stress intensity factors and correction functions Y for a large variety of typical
crack configurations are available in several fracture mechanics books and hand-
books, specially for isotropic materials. For a general anisotropic material, we usu-
ally are interested on being able of characterizing fracture mechanics numerically.
We note that in linear elasticity, stress intensity factors and energy release rates are
equivalent.

Computation of the general release rate in an elastic system


If the material is elastic and the external conditions remain fixed (no change in
external loads and displacements), then the decrease in potential energy (elastic
energy) due to a crack advance ∆a is spent in surface energy, creating the new
surfaces 2∆a of the crack; i.e. in this case, the principle of conservation of energy
implies for mode I and for a unit depth
Z ∆a Z 0
GI ∆a + σ̄yy (x, uy (x)) duy (x) dx = 0 (898)
0 σyy

where duy (x) is the differential of crack opening displacement at x, σ̄yy dx is the load
normal to the crack per unit crack depth during the releasing process and σyy is the
value of the stress prior to crack advance. The x-axis is assumed aligned with the
crack, with its origin at the crack tip. In a linear elastic problem the inner integral
is one half of the limit values
Z ∆a Z 0 Z
1 ∆a
σ̄yy (x, uy (x)) duy (x) dx = − σyy Uy (x) dx (899)
0 σyy 2 0

where Uy (x) = u+ −
y (x) − uy (x) = uy (r, θ = π) − uy (r, θ = −π) is the final crack
opening displacement at x and σyy is the stress before crack advance. Then
Z ∆a
1
GI ∆a = σyy (x) Uy (x) dx (900)
2 0

This expression is valid for any linear elastic material, isotropic or anisotropic, and is
often used to compute the energy release rate in numerical methods. The procedure
is as follows:

1. Generate a mesh for the desired problem with a small crack of length a.

2. Solve the elastic problem and compute the stresses σy (x) for that problem.

3. Generate an identical mesh but with a crack length of a + ∆a, with ∆a very
small.

- 271-
4 Mechanical analysis of composite laminates

4. Compute the crack opening displacements Uy (x) for x ∈ [0, ∆a].

5. Compute the energy release rate in mode I as


1
R ∆a
2 0 σy (x) Uy (x) dx
GI = (901)
∆a

In finite elements, stresses are transferred between elements by forces at nodes.


Then, forces should be computed at nodes and crack openings are the difference in
displacements at those nodes after adjacent elements separate, so

1 Fyn Uyn
GI = (902)
2 ∆a
where Fyn is the nodal force at node n just before the separation, Uyn is the nodal
separation and ∆a is the element side. The drawback of this procedure is that
elements have to be very small so ∆a can be considered as da. Moreover, loads at
nodes are inaccurate because stresses are computed at integration points; at nodes
are simply an extrapolation. Linear finite elements are also incapable of capturing
strong stress gradients. Then, obtaining an accurate stress intensity factor using
this procedure is very expensive in computational terms .

Relation between stress intensity factor and energy release rate


The energy release rate may be related to the stress intensity factor using Equations
(889), (890) and (900) using ũIy (θ = π, ι) = (1 + ι) and ũIy (θ = −π, ι) = − (1 + ι):

KI+ √
Uy (x) = 2uy (r, θ = π) = (1 + ι) √ ∆a − x ; x ≤ ∆a (903)
G 2π
KI
σyy (x) = σyy (r, θ = 0) = √ ; x>0 (904)
2πx
where KI+ is the stress intensity factor after crack advance; we can assume KI+ ≃ KI
for ∆a ≪ a. Then
R
1 ∆a
2 0 σyy (x) Uy (x) dx
GI =
∆a r
2 Z ∆a
1 + ι KI ∆a − x
= dx
∆a 4Gπ 0 x
K2
= (1 + ι) I (905)
8G
Then for the cases of plane stress and plane strain, considering that G = E/(2 (1 + ν))

 KI2
 Plane stress:

E
GI = (906)
 2
 Plane strain: 1 − ν 2 KI


E

- 272-
4.8 Considerations on fatigue and fracture in anisotropic laminates.

The energy release rates may be computed for each fracture mode separately and
then added to compute the total energy release rate. Mode II follows the previous
form, and for mode III
2
KIII
GIII = (907)
2G
For anisotropic materials Eq. (889) is no longer valid and the definition of the stress
intensity factor is more arguable, in contrast to the energy release rate. In these
cases one can define

KI = lim σyy (r, θ) 2πr (908)
r,θ→0

This expression would be valid also for numerical computations using finite elements.
However, because of the singularity, it is difficult to obtain an accurate solution using
standard elements in the singularity dominated zone. Special finite elements and
techniques have been developed with different degree of success.

Limit of linear elastic fracture mechanics: small scale yielding

In all derivations above we have considered linear elasticity. Equation (875) shows
that the stresses at the crack tip (r → 0) would be infinite, which we know it is not
possible. The reason they are not infinite is that even for very brittle materials, a
small zone of plastic behavior is developed at the crack tip, so the stresses there take
the value σY , where σY is the yield stress (elastic limit).
The Griffith theory is then modified by Irwin adding to the surface energy the
energy rate Gp dissipated at this plastic zone

G = 2S + Gp (909)

In brittle materials the addend 2S is dominant over the addend Gp (for example
in glass G ≃ 2S = 2 J / m2 and Gp ≪ 2S), whereas in ductile materials as steel
G ≃ Gp = 1000 J / m2 and 2S ≪ G p ).
The shape and the characteristic dimension rp of the plastic zone for mode I
can be obtained by simple substitution of the plastification criterion on the stress
distribution and using the assumption that equilibrium is preserved (so the integral
below the linear elastic stress curve is the same as that below the elastoplastic one
in the plastic zone, see Figure 117). The dimension of the plastic zone depends
on the hardening characteristics of the material. However, a simple substitution of
Equations (889) in the von Mises criterion (plane stress) yields a dimension of the
order of
   
1 KI 2 KI 2
rp ≃ = 0.16 (910)
2π σY σY

It is assumed that the small scale plastic zone treatment is valid if rp is much smaller
that the typical dimensions of the problem, i.e. rp ≪ a and rp ≪ h (thickness) in
plane strain.

- 273-
4 Mechanical analysis of composite laminates

Figure 117: Linear elastic and elastoplastic stress distribution under the assumption
of small scale yielding.

Initiation of fracture
We have seen that we can characterize the maximum stress state near the crack tip
by either the stress intensity factor or either the energy release rate. In plasticity,
the stress state is characterized by a function (for example the von Mises criterion)
and then compared to the yield stress in order to know if plasticity takes place. In
fracture, the stress intensity factor K (or equivalently the energy release rate G) is
compared to the critical stress intensity factor Kc or to the critical energy release
rate Gc , so during fracture

K ≥ Kc and G ≥ Gc (911)

The critical stress intensity factor Kc is considered a material parameter that must
be measured in equivalent loading conditions and geometry. For brittle materials,
the critical stress intensity factor Kc is fairly independent of the crack length a (i.e.
constant with ∆a) because 2S is constant, but for ductile materials it follows a sort
of “hardening” law, mainly because plasticity is involved and dissipation rate Gp
is not constant, but depends on the hardening of the material. The critical stress
intensity factor is also lower for plane strain cases than for plane stress cases —c.f.
Eq. (906), so in the interior of an specimen plane strain conditions prevail, see
Figure 118. Then, the most relevant critical stress is that of plane strain and this
material property is called fracture toughness (with units of N m3/2 ), designated
by KIC , and characterizes the initiation of fracture for a sharp crack for a plane
strain specimen under mode I. This material property is measured in expensive
standard tests and values for many materials are widely available. We can of course
relate KIC to its associated
 critical energy release rate by means of Eq. (906), i.e.
GIC = KIC 2 1 − ν 2 /E. From Eq. (909) it is seen that this property for ductile

- 274-
4.8 Considerations on fatigue and fracture in anisotropic laminates.

Figure 118: Variation of the critical stress intensity factor with the thickness. The
value KIC is the toughness.

Figure 119: Variation of the toughness with the temperature.

materials is closely related to the plastic dissipation, which in a uniaxial test is the
area below a stress-strain curve. Hence from a uniaxial test we can infer how tough
or brittle a material is.
The fracture toughness in metals may have a strong dependence on temperature
for a given range of temperatures, see Figure 119. There is a threshold after which the
toughness may be considered fairly constant and high, but below a given temperature
metals behave in a brittle manner, i.e. there is a ductile-to-brittle transition. This
was one of the reasons behind the accidents of the Liberty ships during WWII.

Transition flaw size

We are used to compute failure of materials using the yield stress σY so if σ ∞ ≥ σY


plastic flow will take place. However we have seen that fracture may be the dominant

failure mode if KI ≥ KIC or σ ∞ ≥ KIC / πa. Then the following crack length marks
the transition between plasticity and fracture dominant behavior
 2  2
KIC 1 KIC KIC
√ = σY ⇒ at = = 0.32 (912)
πat π σY σY

- 275-
4 Mechanical analysis of composite laminates

If a < at , then plastic behavior is dominant. If a > at fracture failure is dominant.


Metals have at of the order of 10 mm to 100 mm, whereas glass and ceramics of the
order of 0.001 mm. Comparing Equations (910) and (912)

rp ∼ at /2 (913)

For a linear elastic fracture dominated problem we have that a ≫ at and a ≫ rp ,


for example a ∼ 10at . Indeed, the requirement for ASTM standard tests is a ≥
2.5 (KIC /σY )2 and to ensure plane strain also h ≥ 2.5 (KIC /σY )2 .
The form of Eq. (912) is also used to compute the crack size that induces fracture
failure for a given stress level. For example if σ ∞ = σY /2, then the critical length
of the inner crack is given by
 2  2
2 KIC 8 KIC
KI = KIC ⇒ 2ac = = (914)
π σ∞ π σY

It is apparent that for a given stress level related to the yield stress the quotient
(KIC /σY ) governs the critical crack size for fracture failure. Hence, if we want “easy
to check” materials we need high KIC and low σY .

Stable and unstable crack growth

Once there is a crack of initial length a0 , it is relevant to know whether its growth
will be stable or unstable once its associated critical stress σ ∞ is reached. Stable
growth means that an small increment in the loading conditions produces an small
growth of the crack size, or in other words crack advance will stop itself. Unstable
crack growth means that, after an small increment in the loading conditions, fracture
will proceed until total rupture. Then, it is convenient to make distinction among
the critical stress intensity factor for initiation of crack advance Ki and the critical
stress intensity factor for unstable crack advance Ku . For very brittle materials,
Kc = Ki = Ku , so there is little discussion here. For more ductile materials Ku is
greater than Ki , see Figure 120, and it is usual to consider Kc = Ku , i.e. the critical
value is associated to the unstable crack growth initiation11 .
The critical stress intensity factors curves associated to the crack given in Figure
120 are often referred to as the resistance curve of the stress intensity factor and
denoted by KR (∆a). The curve KR (∆a) strongly depends on the plastic behavior
of the material. If curves KR (∆a) and K (a) intersect, then the possible crack
growth is stable, see Figure 120, because a small increment of σ ∞ provokes a small
increment of K and a subsequent controlled increment of a. There is a value σu∞
such that K (a) is tangent to KR (∆a), which determines the critical value Ku = Kc .
It is apparent that a further variation on the load will produce an unstable crack
growth and an imminent failure if load is not released. That point can be evaluated
for the given problem in order to know Ku , σu∞ and the crack length au .

11
Some authors use Kc = Ki , so Kc is associated to the stable crack advance initiation.

- 276-
4.8 Considerations on fatigue and fracture in anisotropic laminates.

K
unstable
K growth
unstable initiation K R (Δa)
growth
initiation
Ku=Kc
K R (Δa) stable growth
Ki=Ku=Kc
Ki initiation

no growth no growth

Figure 120: Stress intensity resistance curve for brittle (left) and ductile (right)
materials for a given crack of initial length a0 .

The J-integral
When plasticity or nonlinear elastic behavior is important, the stress intensity factor
is often not a convenient parameter to characterize the stresses and fracture possibil-
ities. In this case, other material parameters that may account for linear behavior
and also nonlinear one are more appropriate. The J−integral developed by Rice
and independently by Cherepanov in 1968 is one of the possibilities widely extended
in the USA. The CTOD (Crack Tip Opening Displacement) from Wells in 1961
and the CTOA (Crack Tip Opening Angle, a more stable measure) are alternatives
developed and used mainly in Europe and can be related to the J−integral.
The J-integral definition departs from the fact that the variation of the potential
energy in an isolated system is, per unit thickness
Z Z
Π̇ = ẆdA − t · u̇ dΓ = 0 (915)
A Γ

where Γ is an arbitrary path that surrounds the crack tip (the result is path inde-
pendent). If that variation takes place because a crack is propagating in x direction
(note that an increase of a means Π decreases) then the following integral, denoted
as JΓ , vanishes Z Z
∂Π ∂W ∂u
JΓ = − = dA − t · dΓ = 0 (916)
∂a A ∂a Γ ∂a
i.e. moving the domain instead of the crack
Z Z Z Z
∂Π ∂W ′ ∂u ∂W ∂u
JΓ = − = dA − t· dΓ′ = dA − t · dΓ = 0 (917)
∂x A′ ∂x Γ′ ∂x A ∂x Γ ∂x

where A′ and Γ′ are simply moved da along x respect to A and Γ. We note that
   
∂W ∂W ∂εij 1 ∂ ∂ui ∂uj ∂ ∂ui
= = σij + = σij (918)
∂x ∂εij ∂x 2 ∂x ∂xj ∂xi ∂xj ∂x

- 277-
4 Mechanical analysis of composite laminates

where we used the symmetry of the stress tensor so σij = σji . Since by equilibrium,
in the absence of body loads (we neglect them) we have ∇ · σ = 0 or σij,j = 0, we
can write  
∂W ∂ ∂ui
= σij
∂x ∂xj ∂x

and using Gauss generalized theorem, where Γ is the boundary of A and n̂ is the
vector normal to A at the boundary
Z Z   Z Z
∂W ∂ ∂ui ∂ui ∂u
dA = σij dA = n̂j σij dΓ = t· dΓ (919)
A ∂x A ∂xj ∂x Γ ∂x Γ ∂x

which proves Eq. (917). However, using again Gauss generalized theorem (which in
this case is nothing else than the fundamental theorem of calculus applied in x)
Z Z
∂W
dxdy = Wdy (920)
A ∂x Γ

so Z Z
∂u
JΓ = Wdy − t· dΓ = 0 (921)
Γ Γ ∂x
for any closed curve Γ, which is the usual form found in the literature for the
J−integral of Rice (and the one given by himself in his paper). The utility is, of
course, to apply the expression for open curves. Evaluating the integral in a portion
Γ1 of a cleverly selected path yields the value also in the other part of the path Γ2
which will include the crack tip, i.e.
Z Z
∂u
JΓ = JΓ1 + JΓ2 = J − J = 0 ⇒ J = Wdy − t· dΓ (922)
Γ1 Γ1 ∂x

By Equation (916)1 we note that in linear fracture (small scale yielding) J = G and,
for example
1 − ν2 2
J= KI (923)
E
in plane strain cracks in mode I (and similarly in other cases). However, we note
that the use of the J integral is valid also in the nonlinear fracture mechanics case
and the stress intensity factors in this case may be written in terms of the nonlinear
behavior of the material, for example, a Ramberg-Osgood law of the form
 n
ε σ
=α (924)
εo σ0

where ε0 and σ0 are reference yield strain and stress and α and n are material
parameters. If n = 1 we obtain a linear elastic behavior and if n → ∞ we obtain a
perfectly plastic behavior. The solution is out of the scope of these notes but can
be found in most nonlinear fracture mechanics books.

- 278-
4.8 Considerations on fatigue and fracture in anisotropic laminates.

The finite element computation of stress intensity factors


An important fact that any engineer must recognize using finite elements is that the
usual standard formulation is unable to adequately capture the strong discontinuity
at a crack tip, no matter how fine the mesh is. There are many procedures in
the literature to circumvent this difficulty. One of them is to use a zone wider
than the Singularity Dominated Zone, i.e. the so-called Three Parameter Zone,
which includes an additional term in the solution but also a larger, more tractable
zone. Other option is to use a special placement of mid-side notes in serendipity

elements so the strains take naturally the 1/ r distribution as a consequence of
the modified interpolation. Another option is to create special finite elements that
naturally include the singularity terms in the interpolation functions. Whereas the
first two options may be used in any finite element code with standard elements,
the last option can be used only if the special elements are available in the code.
New XF EM and meshless techniques also present interesting procedures to model
fracture.

Three-parameter-zone If we consider a further term in Equation (889) we have

KI √
σy (θ = 0) = √ + A1 x (925)
2πx
and Eq. (890) for θ = π (crack opening zone) is
q
2 p 4
E uy (θ = π) = KI 2π |x| − A1 |x|3

(926)
π 3

where E ∗ = E for plane stress and E ∗ = E/ 1 − ν 2 for plane strain. Then, we can
define for this zone an apparent stress intensity factor KIap when compared to the
singularity dominated zone such that
n √ o √
σy (θ = 0) 2πx =: KIap = KI + 2πA1 x (927)

and for θ = π (crack opening zone)


( ) √
π ∗ 1 ap 2π
E uy (θ = π) p =: KI = KI − A1 |x| (928)
2 2π |x| 3

The point here is that in both cases, the quantities in curly brackets will yield
apparent stress intensity factors in the three parameter zone which vary linearly
with the distance to the crack tip. Then, when plotted those linear curves may be
extrapolated to obtain the value KI at the crack tip (x = 0). With this procedure,
the stress intensity factor may be obtained from nodes and stress integration points
just in the three parameter zone, which is significantly larger (several times) than
the singularity dominated zone. Note that displacements are the ones obtained at
the nodes (stresses at integration points), so displacements usually yield a better
accuracy. However plotting both apparent values against x will facilitate a more
accurate determination of KI , see Figure 121.

- 279-
4 Mechanical analysis of composite laminates

Figure 121: Computation of the stress intensity factor from the results of displace-
ments and stresses in the three-parameter zone.

Finite element interpolation for singularities Other procedure that can be


used is to modify the finite element interpolation to capture the singularity at the
cracks. This can be done modifying the middle node placement in isoparametric
elements or by directly modifying the interpolation functions and leaving the nodes
at their natural position.
For the case of isoparametric serendipity elements, the eight-node element may
be collapsed to a triangle assigning the same location to nodes 1, 4 and 8 (see Figure
122–top), being this location that of the crack vertex. If nodes 5 and 7 are placed
at 1/4 from the crack vertex nodes and at 3/4 from the opposite side, then the
resulting strain field in the element as a result of the isoparametric interpolation
transformation takes the form

Ax Bx
 εx = √r + r + Cx


(929)
A B
 εy = √y + y + Cy


r r

The interesting point is that if linear elasticity is considered, then the 1/ r term
dominates the solution and represents the asymptotic solution in the singularity
dominated zone, see Eq. (889) and note that stresses and strains are simply related
to each other by elastic constants. Furthermore, if nodes 1, 4 and 8 have initially
the same location but are left untied (i.e. upon deformation they may have different
locations in space) it can be shown that a crack blunting is obtained and that the
1/r term dominates the solution. This is the solution near the crack tip obtained
for perfectly-plastic materials. We note that a similar solution is obtained for the
isoparametric 12 node serendipity element if nodes are located as shown in Figure
122–bottom.
These modified elements allow for a simple computation of stress intensity factors

- 280-
4.8 Considerations on fatigue and fracture in anisotropic laminates.

Figure 122: Special nodal placement in serendipity finite elements for capturing the
singularity in fracture.

- 281-
4 Mechanical analysis of composite laminates

using Eqs. (890) and (895)


 √
 2πG u+ y − uy


 K = √
 I
 r (ι + 1)
√ (930)


 2πG |u+ −
x − ux |
 KII =
 √
r (ι + 1)

where u+ and u− refer to the displacements of two points in opposite faces of the
crack at a distance of r from the tip. It is assumed that these points are near the
tip in the singularity dominated zone (i.e. the closest ones to the tip with r > 0)
and that the crack is parallel to the x− axis. Note that these special finite elements
interpolate accurately the near tip behaviour at the cost of a poor representation of
smooth distributions. Hence they should be used only in the singularity dominated
zone.
One of the drawbacks of this procedure is that mid-side nodes must be accurately

positioned because otherwise the 1/ r term is not accurately represented and large
errors may be obtained when computing the stress intensity factors. Furthermore,
this procedure presents the inconvenience that nodes must be located “by hand”.

Finite elements that explicitly include the 1/ r term in the interpolation functions
may be developed and can be found in some codes. The procedure to include
these functions preserving the interpolation property is rather straightforward, and
common to the procedure for transition elements.
Assume that we have a standard element with Ni (x) (i = 1, ..., n) shape func-
tions which fulfill the interpolation property in the i = 1, .., n nodes, i.e. Ni (xi ) = 1
and Ni (xj ) = 0 for j 6= i. Then we wish to modify these functions inserting a new
node n + 1 with an associated shape function Nn+1 (x) which does not fulfill the
interpolation property; note that inserting a new node also brakes the interpolation
property in the previous functions. Then, we can restore the interpolation property
in all n + 1 nodes and shape functions using the following procedure

1. Correct new function to fulfill the interpolation condition in i = 1, ..., n nodes


n
X
Nn+1 (x) − Nn+1 (xi ) Ni (x)
i=1
Nn+1 (x) ← n (931)
X
Nn+1 (xn+1 ) − Nn+1 (xi ) Ni (xn+1 )
i=1

2. For i = 1 to n, correct the i shape functions to fulfill the interpolation condition

Ni (x) ← Ni (x) − Ni (xn+1 ) Nn+1 (x) (932)

If more nodes/shape functions are to be introduced, the procedure is repeated


for those new nodes/shape functions.

Exercise 75 Use the procedure given in Eqs. (931) and (932) to generate a uni-
dimensional three-node element with an associated shape function which models a

- 282-
4.8 Considerations on fatigue and fracture in anisotropic laminates.

singularity of the form r α and also captures the linear behavior.


Solution: Consider an element with initial nodes at locations r1 = 0 and r2 = 1
and an additional node placed at r3 = 1/2. The following (initial) shape functions
are the standard linear ones

N1 (r) = 1 − r, N2 (r) = r (933)

such that N1 (r1 = 0) = 1, N1 (r2 = 1) = 0 and N2 (r1 = 0) = 0, N2 (r2 = 1) = 1,


so the interpolation condition is fulfilled in the first two nodes and shape functions.
The third required shape function is of the form

N3 = r α (934)

The reader can verify that the interpolation condition is not satisfied upon the con-
sideration of this additional node. To achieve this we use Eqs. (931) and (932)

N3 (r) − N3 (r1 ) N1 (r) − N3 (r2 ) N2 (r)


N3 (r) ←
N3 (r3 ) − N3 (r1 ) N1 (r3 ) − N3 (r2 ) N2 (r3 )
rα − 0 − 1 × r α+1 r − r
α
= 1 1 = 2 (935)
2α − 0 − 1 × 2
2α − 2

1 r − rα
N1 (r) = (1 − r) − 2α+1 α
2 2 −2
1 α+1 r − r α
= (1 − 2r) + r − 2 (936)
2 2α − 2
α α
2 r − 2r
= (1 − 2r) + (937)
2α − 2

1 r − rα 2α r α − 2r
N2 (r) = r − 2α+1 α = (938)
2 2 −2 2α − 2
It can be shown that

1

N1 (0) = 1, N1 2 = 0, N1 (1) = 0 (939)

1

N2 (0) = 0, N2 2 = 0, N2 (1) = 1 (940)
1

N3 (0) = 0, N3 2 = 1, N3 (1) = 0 (941)

It is left to the reader demonstrate that if u (r) = Ar α + Br + C, then

3
X
ũ (r) = u (ri ) Ni (r) ≡ u (r) (942)
i=1

so u (r) is exactly interpolated over r ∈ [0, 1], as required.

- 283-
4 Mechanical analysis of composite laminates

log(A)

Figure 123: Crack growth rate in fatigue for different values of the cyclic stress
intensity factor (Paris Law).

4.8.2 Fatigue

Fatigue is a phenomenon by which materials subjected to small alternating or cyclic


stresses become progressively damaged until failure. The level of the alternating
loads leading to failure may be much smaller that those under static loading. Fatigue
in materials is prompted or increased by several circumstances. Material defects
(either on origin or originated by use), surface rugosity, environment and, of course,
stress concentrations and irregularities may strongly affect the fatigue life of an
specimen. Fatigue is initiated at small imperfections where cracks nucleate. Under
small (but highly enough) cyclic stresses an imperfection grows into a relevant crack
until the critical size is attained. Then failure is rapidly obtained.
Composite materials behave in fatigue somehow differently than bulk materials.
In composite materials imperfections are larger and more frequent than in bulk ma-
terials. Hence, fatigue initiation is faster. However, crack propagation in composite
materials is much slower than in metals and the critical size is usually also larger.
The reason is that the different materials and discontinuities act as crack stoppers
and stress distributors. Deformations under damage are also larger and more no-
ticeable, so an important point in composite materials is that fatigue effects may be
noticed on time to avoid catastrophic failure.
Fatigue life in terms of cycles N strongly depends on the amplitude of the cycles
of the alternating stress (typically denoted by S = σmax − σmin ). Then, fatigue life
is typically characterized by Wöhler S − N curves which relate the amplitude of the
stress cycles S to the expected life in terms of cycles N . The curves are much less
sensitive to the mean stress level (in a reasonable interval), but this level may also
be included as a third variable in the S − N curves.
If fatigue is studied within the framework of linear fracture mechanics, we note
that S = σmax − σmin may be related to a stress intensity factor range ∆K =
Kmax − Kmin . It has been shown experimentally that below a given threshold ∆K0 ,

- 284-
4.9 Introduction to the design of composite structures

fatigue effects may be neglected because life expectancy is extremely large (cracks
do not nucleate or if so, do not propagate). For larger ∆K values the defect growth
is stable and the number of cycles of life may be related to the crack length using
the Paris Law
da
= A (∆K)m (943)
dN
where A and m depend on the material, the environment and the crack resistance.
This law may be used to compute the number of cycles in which the critical crack
size will be obtained, and may also be used to compute the time span (in cycles)
available between detectable cracks and critical cracks. The parameter m is typically
between 2 and 4, and is measured as the slope in log (da/dN ) vs log (∆K) plots,
see Figure 123. If ∆K is large enough, then the crack/imperfection growth becomes
unstable and leads to sudden failure in very few cycles. At this stage, the Forman
law is suitable
da A
= (∆K)m (944)
dN (1 − ρ) Kc − ∆K
where Kc is the critical stress intensity factor and ρ = σmin /σmax is the stress ratio.
An estimation of the life of a specimen under fatigue loading may be obtained
by integration of, for example, Paris’ law
Z ac
1
Nf ail = [∆K (a)]−m da (945)
a0 A

where a0 and ac are the initial and critical crack lengths and ∆K (a) is the stress
intensity factor range as a function of the alternating stress amplitude S = σmax −
σmin and the crack length a.
For the case of cycles of different amplitude, it is customary to normalize the
number of cycles n(i) at a given cyclic stress by the expected life in cycles Nf ail(i) at
that stress level.Then, failure occurs when the addition of all ratios reach the value
of one, i.e.
X n(i)
≤1 (946)
Nf ail(i)

4.9 Introduction to the design of composite structures


The purpose of this section is to simply make the reader aware of the big difference
between the analysis and the design of a structure, and the difficulty in achieving a
standard or so formal training in design: major design training is frequently obtained
with practice.
In analysis of structures, a problem is usually given. This means that we have a
structure with given dimensions, with a given material and geometry, to withstand
given forces and with a defined manufacturing process. Cost is not typically a
variable. The solution is usually unique.
In contrast, in design very few parameters are defined beforehand. Usually we
wish a structure to meet a given purpose at the minimum possible cost; and by
minimum cost we understand for example cost of materials, cost of manufacturing,
cost of maintenance, incurred operational costs and even ulterior cost of disposal.

- 285-
4 Mechanical analysis of composite laminates

In order to meet the purpose, some loads must be supported by the structure, but
these loads may depend on the structure itself and even on variables like cost or
design life. The geometry is also usually a variable, the material(s) must meet some
requirements, but typically there are several possibilities. Manufacturing process is
also extremely important because it strongly influences the cost and even the possi-
bility of the structure to be built. The manufacturing process of course also depends
on the selected material. Metals and composite materials have very different manu-
facturing procedures. Finally, we note that the solution in design is not unique and
not optimum. Engineering cost is also a variable and reaching an optimum solution
is not really “optimal” because the cost of the increased engineering manpower to
reach such solution may also overcome the cost benefits of that solution over others.
In summary, whereas analysis of structures is close to mathematics in philosophy
because things are rather deterministic, design is closer to medicine in philosophy
because it is more heuristic, with solutions based on experience and more problem-
oriented. Design can also be considered as an art, because personal talent and
creativity for the search of untypical solutions also plays an important role.

4.9.1 The approach

A typical approach in design is to first define the problem and the minimum require-
ments to be met. For example whatever is built must usually be able to withstand
its own weight and some given loads. A set of possible materials is usually also
defined (for example types of metals in corrosive areas, types of ceramics working
at high temperatures, ductility requirements, and so on). Experience usually gives
some guiding dimensions and shapes as a point of departure (think for example in
a wing or in a bicycle frame).
Once the point of departure is defined, the approach is always successive. Do not
attempt at first to solve “the” structure, but just to get a rough idea of the order of
magnitude of the solution, just approximate numbers. For example, a wing may be
approximated by a beam with a section similar to that of the wing, without details.
Handbooks, hand calculations and solutions to similar problems are typically used
at this point. These numbers allow for a more detailed definition of the problem
and also to gain confidence in the solutions obtained.
After a rough idea of the problem is obtained, a more sophisticated analysis
can be performed. The geometry now is better approximated, and the loads and
requirements better defined. In composite materials, “equivalent” isotropic material
properties may be used at this point in order to study and understand the behavior
in an approximate manner. From this understanding, desirable orthotropic material
properties may be defined. Using tables of materials and tools (Matlab scripts) as
the ones given in these notes, a laminate structure that approaches those properties
may be designed.
At this point, a loop of successive refinement is entered. The structure is gaining
shape and the material properties and loads are better defined. The engineer is
also gaining confidence in the results obtained. The analysis may bring conclusions
which recommend some changes in the proposal. Manufacturing possibilities and
procedure do also enter here as variables. Details of the structure such as config-

- 286-
4.9 Introduction to the design of composite structures

uration, stiffeners, elements, etc., are also defined during this procedure. Finally a
tentative design (or a very few number of possible variants) is obtained.
The analysis and study of variants is the following step. Here a more detailed
analysis is performed, and usually includes static, dynamic and buckling analysis.
Nonlinearities may also be included depending on the problem. At this stage it is
typical the “what if” type of analyses to refine the solution. Of course the possible
manufacturing procedure and the procedures to reduce cost are also to be taken into
account.

4.9.2 The details

Once the geometry, loads and possible manufacturing procedures are being defined,
some details need to be considered. For example, if the structure is composed of
different parts, the procedure to join those parts and make them work together as
desired must be defined. Different parts may be joined together by bonding, by
bolts, etc. The geometry of the joint is also relevant; it depends on the way the
joint is working and, of course, on other parameters like manufacturing possibilities.
Design of joints is extremely important both in metal structures and in composite
ones. The execution procedure of the joints is also very important and strongly
affects the carrying capacity of the joint and of the structure as a whole. However,
the analysis of these joints is out of the scope of these introductory notes. The
reader may refer to the given References.
Another detail that affects the durability of the structure is the edge and surface
termination. We have seen that in the particular case of composite structures,
edge interlaminar stresses may be significant and give way to delamination. These
interlaminar stresses depend on the laminae layout in the laminate but also on the
edge termination. Progressive ply drops, stiffeners and stress releasers are typical to
avoid stress concentrations, delaminations and undesirable effects at discontinuities.
A detailed analysis of these effects must be taken into account in the design process.
Of course, during the study and definition of the details, many aspects of the
tentative design may need a relevant modification, so the design loop may go again
to previous design stages.

4.9.3 Responsibility of the structure

One of the main issues when designing a structure in general is the responsibility
of the structure. By this we mean the consequences of a failure of that structure.
For example, the consequences of the failure of the wing, the horizontal stabilizer
or the tail of an airplane are much more catastrophic than the consequences of the
failure of a bicycle frame. Moreover, the consequences of a catastrophic failure of
a bike frame are larger than those of just a large deformation. The consequences
of buckling of a component of a wing which does not lead to catastrophic failure
may be smaller than the buckling of a component of the landing gear. Frequently
by “consequences” we talk about “lives” but also about economic ones. Failure of a
structure may provoke an important loss on the image of a company even if direct
losses are not important. To rebuild that image, if ever possible, may be very costly.

- 287-
4 Mechanical analysis of composite laminates

Depending on the responsibility of the structure and the consequences of failure,


the loads and safety factors are different. Nonlinear analysis with large and per-
manent deformations may be acceptable in some cases, but unacceptable in others.
For example, during a car accident, the important issue is to save the life of the
occupants. For this task is even necessary to have plastic deformations to absorb
the energy during the impact. However we do not want to have those permanent
deformations for small impacts during parking and, even during important impacts,
some parts of the car must remain under very low deformations to avoid injuring the
occupants. Then, for each load case and structural component, the possible level of
deformation (and hence linear or nonlinear behavior) must be considered.

4.9.4 Mathematical tools for optimal design


As we mentioned previously, obtaining a truly “optimal” design is a virtually impos-
sible task because the number of variables involved in the process of design can be
very large. A usual approach to obtain a good solution is just “try-and-error” (i.e.
by hand). However, there are well-developed mathematical tools that may facilitate
the search for an optimal solution if the number of possible variables are restricted
to a reasonable number. For example, given a load state in a structural member,
it is possible to search for optimal laminates (in terms of number and orientation
of plies) that at the minimum cost optimize the strength or the stiffness. Here the
design variables would be the number of plies, their orientations and possible layups.
Symmetry considerations on the stacking sequences should also be taken into ac-
count. Discrete values and bounds are usually given to restrict the possibilities to
reasonable ones. Of course the dimension and shape of the structure may also be
included as design variables.
These mathematical procedures are part of the mathematical optimization dis-
cipline, which is common to many engineering and economic fields. For example,
Matlab has a toolbox on mathematical optimization. There are also sets of subrou-
tines written in high level languages like Fortran, as for example the IMSL library.
Many commercial finite element programs have optimization tools, or at least, the
input and output files may be linked by the user to the commented subroutines.
In summary, the mathematical optimization easies the search for optimal solu-
tions. There are many algorithms available. The treatment of such algorithms is
far beyond the scope of these notes because they constitute a discipline themselves.
However, we just mention that common to those algorithms is the definition of both
the design variables and the cost or objective function. Design variables are those
variables we allow to change within some given intervals; the objective function is
the function to be minimized, which may include the weight of the structure, its
cost, its strength, etc., or any combination of all of them.

- 288-
A Cauchy’s Tetrahedron
From the infinitesimal volume shown in Figure 124, defined at a given point P and
known as the Cauchy tetrahedron, the balance of the four forces acting on the four
surfaces yield the relation12

F int int int int


n + F (−x) + F (−y) + F (−z) = 0 (947)
tn An + t(−x) A(−x) + t(−y) A(−y) + t(−z) A(−z) = 0 (948)
tn An − tx Ax − ty Ay − tz Az = 0 (949)

where the identities A(−i) = Ai and t(−i) = −ti (directly deduced from Eq. (13) and
known as the Cauchy’s fundamental lemma) have been used.
z
int
F(-x)
int
int
F(-y) Fn
y
x int
F(-z)
Figure 124: Cauchy Tetrahedron

From geometrical considerations, we observe that, for example, the area Ax is


the projection of An onto the plane Oyz. Accordingly we can write Ax = An · ex =
An n̂ · ex . Therefore, the traction vector tn may be expressed in terms of tx , ty and
tz through
Ax Ay Ay
tn = tx + ty + tz (950)
An An An
tn = tx (n̂ · ex ) + ty (n̂ · ey ) + tz (n̂ · ez ) (951)
tn = tx n̂x + ty n̂y + tz n̂z (952)

where n̂x , n̂y and n̂z are the components of the normal unit vector n̂ in the reference
frame Xref = Oxyz. Since the orientation of n̂ in Eq. (952) is completely arbitrary
(for Xref fixed and any direction n̂, a different tetrahedron with associated vector
tn can be defined), we can conclude from Eq. (952) that any traction vector at the
point P can be calculated in terms of the six independent stress components σxx ,
σyy , σzz , σxy , σyz and σzx contained in the vectors tx , ty and tz , which is an essential
12
Volume forces acting on the tetrahedron volume ∆V are neglected in the balance equation since
they are negligible with respect to surface forces when the limit ∆V → 0 is considered, since then
∆V /∆A → 0.

- 289-
A Cauchy’s Tetrahedron

result. That is, we only need the considered six stress components at a given point
(or other six components in other arbitrary reference frame) in order to completely
define the stress state at that point. Finally, note that Eq. (952) may be expressed
using the second-order symmetric stress tensor σ
 
σxx σxy σzx
[σ]Xref =  σxy σyy σyz  (953)
σzx σyz σzz
as  
h i  n̂x 
{tn }Xref = {tx }Xref {ty }Xref {tz }Xref n̂ (954)
 y 
n̂z
or
tn = σ · n̂ (955)
For a given stress tensor σ, Eq. (955) represents a fundamental linear relation
(Cauchy’s postulate) between the traction vector tn and the normal vector n̂.

- 290-
B Transformation Rules and Rotations of Vectors and
Second-Order Tensors
B.1 Transformation Rules for Change of Coordinates
Consider two different systems of representation in the three-dimensional space cen-
tered at a common origin O, X ′ = O1′ 2′ 3′ and X = O123. A vector v is expressed
in terms of the basis vectors associated to each reference frame through
3
X 3
X
v= vj ′ ej ′ = vj ej (956)
j=1 j=1

The dot product of v and a X ′ -basis vector ei′ (with i = 1, 2, or 3) yields

3
X 3
X
v · ei′ = vj ′ (ej ′ · ei′ ) = vj (ej · ei′ ) (957)
j=1 j=1

Since ej ′ · ei′ = 1 if j = i and ej ′ · ei′ = 0 if j 6= i, then

3
X
vi ′ = (ei′ · ej ) vj i = 1, 2, 3 (958)
j=1

In matrix notation, we write


    
 v1′  e1′ · e1 e1′ · e2 e1′ · e3  v1 
v′ =  e2′ · e1 e2′ · e2 e2′ · e3  v (959)
 2   2 
v3′ e3′ · e1 e3′ · e2 e3′ · e3 v3

or, compactly
{v}X ′ = [r]X→X ′ {v}X (960)

with
 
e1′ · e1 e1′ · e2 e1′ · e3
[r]X→X ′ =  e2′ · e1 e2′ · e2 e2′ · e3  (961)
e3′ · e1 e3′ · e2 e3′ · e3

The matrix [r]X→X ′ defines the transformation rule for the change of coordinates
of v from X to X ′ , which is uniquely determined for a given pair of reference frames
X and X ′ . Since a · b (a and b being unit vectors) represents both the projection of
a onto b and the projection of b onto a, note the jth column of [r]X→X ′ gives the
representation of the vector ej in the coordinate axes X ′ and the ith row of [r]X→X ′
gives the representation of the vector ei′ in the coordinate axes X. That is
 
T
   {e 1 X
′ }
[r]X→X ′ = {e1 }X ′ {e2 }X ′ {e3 }X ′ =  {e2′ }TX 

(962)
{e3′ }TX

- 291-
B Transformation Rules and Rotations of Vectors and
Second-Order Tensors

This last interpretation is somehow mnemonic since, for example, if v = e1 , then


{v}X = {e1 }X = { 1 0 0 }T and
 
  1 
{v}X ′ = [r]X→X ′ {v}X = {e1 }X ′ {e2 }X ′ {e3 }X ′ 0 = {e1 }X ′ (963)
 
0

Aside, it follows from Eq. (960) that

{v}X ′ = [r]X→X ′ {v}X ⇒ {v}X = [r]−1


X→X ′ {v}X ′ (964)

We also have the transformation rule

{v}X = [r]X ′ →X {v}X ′ (965)

with  
e1 · e1′ e1 · e2′ e1 · e3′
[r]X ′ →X = e2 · e1′
 e2 · e2′ e2 · e3′  (966)
e3 · e1′ e3 · e2′ e3 · e3′
whereupon the matrix for the inverse transformation [r]X ′ →X results the transpose
matrix of [r]X→X ′ (compare with Eq. (961)). Hence

[r]X ′ →X = [r]−1 T
X→X ′ = [r]X→X ′ (967)

and the matrix for the change of coordinates [r]X→X ′ becomes orthogonal.
Using the foregoing vector transformation rules it is straightforward to obtain
the transformation rule for second-order tensors. Consider a second-order tensor A
providing a linear mapping between the vectors u and v, that is

u = Av (968)

When this transformation is represented in the reference frame X, it reads

{u}X = [A]X {v}X (969)

If the coordinates of both vectors are changed from X to X ′ , then

[r]X ′ →X {u}X ′ = [A]X [r]X ′ →X {v}X ′ (970)


{u}X ′ = [r]TX ′ →X [A]X [r]X ′ →X {v}X ′ (971)
{u}X ′ = [A]X ′ {v}X ′ (972)

Therefore, the components of A in the system of representation X ′ are given in


terms of the components of A in the system of representation X and the matrix of
the change of basis [r]X→X ′ through

[A]X ′ = [r]X→X ′ [A]X [r]TX→X ′ (973)

- 292-
B.2 Rotations

B.2 Rotations
Consider now two different vectors v and v ′ . Both vectors have identical norm
kvk = kv ′ k but are oriented about different directions in the three-dimensional
space. That is, they are related by means of a solid-rigid rotation of angle θ around
a certain axis Oξ. Consider also two systems of representation, X = O123 and
X ′ = O1′ 2′ 3′ , which are related through the same rotation that applies between v
and v ′ . Then, the coordinates of v in X coincide with the coordinates of v ′ in X ′ ,
i.e.
3
X
v= vi ei (974)
i=1
X3 3
X

v = v e i′ =
i′ vi ei′ (975)
i=1 i=1

We say that v ′ is obtained from v (or the basis X ′ is obtained from X) through a
solid-rigid rotation mapping defined by a rotation tensor Qθξ ≡ Q such that

v ′ = Qv (976)

Equation (976) is qualitatively different to Eq. (960). Equation (960) relates the
corresponding expressions of the same vector in different systems of representation
through a coordinates transformation matrix and Eq. (976) relates two different
(rotated) vectors through a rotation mapping tensor.
In order to obtain the corresponding expression of Q, let us previously introduce
the so-called dyadic product r ⊗ s of two generic vectors r and s. The product r ⊗ s
is a second-order tensor, which when applied to a third vector t gives as a result a
vector in the direction of r with modulus the projection of t onto s. That is

(r ⊗ s) t = (s · t) r (977)

The matrix representation of the tensor r ⊗ s in a given reference frame X = O123


is
   
 r1  r1 s1 r1 s2 r1 s3
T
[r ⊗ s]X = {r}X {s}X = r s1 s2 s3 =  r2 s1 r2 s2 r2 s3  (978)
 2 
r3 r3 s1 r3 s2 r3 s3

thereby (check it)


  
r1 s1 r1 s2 r1 s3  t1 
[(r ⊗ s) t]X =  r2 s1 r2 s2 r2 s3  t (979)
 2 
r3 s1 r3 s2 r3 s3 t3
 
 r1 
= (s1 t1 + s2 t2 + s3 t3 ) r = [(s · t) r]X (980)
 2 
r3

- 293-
B Transformation Rules and Rotations of Vectors and
Second-Order Tensors

Then, Eq. (975) yields


3
X 3
X 3
X

v = vi e i ′ = (ei · v) ei′ = (ei′ ⊗ ei ) v (981)
i=1 i=1 i=1
3
!
X
= ei′ ⊗ ei v = Qv (982)
i=1

where we readily identify


3
X
Q= ei′ ⊗ ei (983)
i=1

The expression of Q in a generic system of representation X ′′ is obtained by means


of
X3
[Q]X ′′ = {ei′ }X ′′ {ei }TX ′′ (984)
i=1

It is remarkable that the matrix representation of the rotation tensor Q depends on


the axes in which it is calculated, whereas the matrix for the change of coordinates
[r]X→X ′ has components being invariants (scalar products) that do not depend on
the reference frame in which they are calculated. Aside, note that with the definition
given in Eq. (983) the tensor Q, in effect, reorientate the X-components of any
vector v onto the corresponding directions of the X ′ -basis vectors, that is

3
! 3 
3 X 3
X X X
Qv = ei′ ⊗ ei  vj ej  = vj (ei′ ⊗ ei ) ej (985)
i=1 j=1 i=1 j=1

X 3
3 X 3
X
= vj (ei · ej ) ei′ = vi e i ′ = v ′ (986)
i=1 j=1 i=1

It follows from Eq. (976) that

v′ = Qv ⇒ v = Q−1 v ′ (987)

where Q−1 is the inverse tensor of Q and stands for the inverse rotation mapping
3
X
−1
Q = ei ⊗ ei′ (988)
i=1

defined by a rotation angle −θ about the axis Oξ. Moreover, from Eq. (978) it is
easily seen that (r ⊗ s)T = s ⊗ r, so

3
X 3
X
Q−1 = ei ⊗ ei′ = (ei′ ⊗ ei )T = QT (989)
i=1 i=1

and the rotation tensor Q is orthogonal.

- 294-
B.3 Relation between Change of Basis Rules and Rotations

Finally, the rotated tensor A′ relative to a second-order tensor A by means of


the rotation tensor Q can be derived operating over the linear mapping

u = Av (990)

Using the fact that QT Q = I (I representing the second-order identity tensor) we


obtain

Qu = QAv (991)
T

= QAQ Qv (992)

so
u′ = A′ v ′ (993)
and the tensorial expression for the rotated tensor A′ (which relates the rotated
vectors u′ and v ′ ) in terms of A and of the corresponding rotation tensor Q is
readily identified
A′ = QAQT (994)
which again is a relationship between two different (rotated) tensors.

B.3 Relation between Change of Basis Rules and Rotations


The matrix expression taken by Q, Eq. (983), when it is represented in the Cartesian
axes defining the rotation itself, i.e. X or X ′ , reads
3
X 3
X
[Q]X = [ei′ ⊗ ei ]X = {ei′ }X {ei }TX (995)
i=1 i=1
 
  e1′ · e1 e2′ · e1 e3′ · e1
= {e1′ }X {e2′ }X {e3′ }X =  e1′ · e2 e2′ · e2 e3′ · e2  = [r]X ′ →X
e1′ · e3 e2′ · e3 e3′ · e3
(996)

and
3
X 3
X
[Q]X ′ = [ei′ ⊗ ei ]X ′ = {ei′ }X ′ {ei }TX ′ = (997)
i=1 i=1
   
{e1 }TX ′ e1′ · e1 e2′ · e1 e3′ · e1
=  {e2 }TX ′  =  e1′ · e2 e2′ · e2 e3′ · e2  = [r]X ′ →X
 
(998)
{e3 }TX ′ e1′ · e3 e2′ · e3 e3′ · e3

Thus, we deduce that the matrix of the change of basis from X ′ to X, i.e. [r]X ′ →X ,
is equivalent
P to the matrix representation of the rotation tensor that maps X to X ′ ,
i.e. Q = ei′ ⊗ ei , when this last tensor is expressed either in X or X ′

[r]X ′ →X = [QX→X ′ ]X = [QX→X ′ ]X ′ (999)

- 295-
C Thermodynamics of solids

C Thermodynamics of solids
C.1 Temperature effects
In order to take into account adequately the temperature effects, we must consider
the first and second principles (or laws) of thermodynamics for a solid

First principle of thermodynamics


This principle states that energy (power) is conserved in any process. The quantities
involved in lagrangian description typically in solids the for a given volume V are

• Internal energy: Z
U= ρudV (1000)
V
where u is the internal energy per unit mass and ρ is the density. This energy,
of difficult characterization, contains all the energy (bonds, internal potentials,
internal kinetic energy, etc) except the kinetic energy as a continuous medium.
This energy is micromechanical in nature. In practice, as we will see later,
we are not interested in knowing its value, but its change under different
processes. Furthermore, we can consider R that the change of this energy is due
1
to deformation Ẇ (for example W = V 2 σ : ε in elasticity) and heating Q
(defined below), i.e.
U̇ = Ẇ + Q (1001)

• Kinetic energy: Z
1
K= ρv · v dV (1002)
2 V
where ρ is the density and v is the velocity field. This energy is excluded from
the “internal energy” by definition (or convention), but it has similar nature.

• Total energy of a system:

E =U +K (1003)

• Mechanical power:
Z Z
P= b · vdV + t · vdS (1004)
V S

where b are the body loads per unit volume (which can include inertia terms),
t are the loads per unit of bounding surface S and v are the velocities.

• Heat power Z Z
Q= ρrdV − q · n̂dS (1005)
V S

where r is the heat generated per unit mass due to volumetric (internal may
be a misleading word here) heat sources, n̂ is the unit vector normal to the

- 296-
C.1 Temperature effects

surface S (i.e. the direction of dS), and q is the local heat flux vector. The
minus sign is a usual convention so q is the heat flowing from hot to cold, and
then the Fourier’s heat conduction law is

q = −K∇T (1006)

where K is the thermal conductivities tensor (which may be isotropic or


anisotropic) and ∇T is the gradient of temperatures.

The first principle states that the variation of the energy of a body (i.e. internal
and kinetic energies) equals the power (either heat, mechanical or of other natures
not considered hereafter) introduced in the system
d
(E) = P + Q+ other possible powers (1007)
dt
d
(U + K) = P + Q+ other possible powers (1008)
dt
or
E˙ = U̇ + K̇ = Ẇ + K̇} + Q = P + Q
| {z (1009)
P
Then if there is no change in mass (lagrangian formulation) so d/dt(ρdV ) = 0
Z Z Z Z Z Z
ρu̇dV + ρv̇ · v dV = ρrdV − q · n̂dS + b · vdV + t · vdS (1010)
V V V S V S

Using Gauss theorem we have


Z Z
q · n̂dS = ∇ · qdV (1011)
S V

and Z Z Z
t · vdS = n̂ · σ · vdS = ∇ · (σ · v) dV (1012)
S S V
We note that

∇ · (σ · v) = (σij vj ),i = σij,i vj + σij vj,i = (∇ · σ) · v + σ : ∇v (1013)

so Z Z Z
∇ · (σ · v) dV = (∇ · σ) · vdV + σ : ∇vdV (1014)
V V V
Then using Equation (1010)
Z Z Z Z
ρu̇dV = (ρr − ∇ · q) dV + (b−ρv̇ + ∇ · σ) · vdV + σ : ∇vdV (1015)
V V V V

We also note that the first principle of thermodynamics must hold for any arbitrary
velocity field v, and for any arbitrary volume V , so the following identities must
hold simultaneously 
 ∇ · σ + b−ρv̇ = 0
(1016)

ρu̇ = ρr − ∇ · q + σ : ∇v

- 297-
C Thermodynamics of solids

The reader should recognize the first equation to be the dynamic equilibrium equa-
tion. The last term of the second identity may be recognized if we take into account
that because of the symmetry of σ we have13

d d
σ : ∇v = σ : sym (∇v) = σ : [sym (∇u)] = σ : [ε] = σ : ε̇
dt dt
where sym (·) indicates the symmetric part of (·) and ε = sym (∇u) is the well
known small strain engineering tensor. Then, Eq. (1016)2 is

ρu̇ = ρr − ∇ · q + σ : ε̇ (1017)

which is often referred to as the intensive form of the first principle of thermody-
namics in solids14 .
We here note that in the absence of mechanical work the first principle gives
U̇ = Q and the specific heat at constant volume is defined as the increment of internal
energy per increment of temperature for a unit mass, i.e. (assuming homogeneous
values)
1 ∂Q 1 ∂U ∂u
cv = = = (1018)
ρV ∂T ρV ∂T ∂T
and
1 ∂Q ∂u
c= = ρcv = ρ (1019)
V ∂T ∂T
is the heat capacity per unit volume, i.e. if strains are considered constant we obtain
a typical expression
u̇ = cṪ = ρcv Ṫ (1020)

Second principle of thermodynamics


The second principle establishes the existence of an internal variable, called entropy
that measures the irreversibility of a thermodynamical process. By observation, we
know that, for example, if two solids (or systems in general) with different tem-
peratures are in contact, they equal their temperatures through a heat flux. The
new variable entropy accounts for how much the equalization of the temperatures
has been going. But we know that the reversed process never takes place without
external intervention, i.e., if two solids (or systems) are in contact at the same tem-
perature, those temperatures do not become different spontaneously. Hence there
is some preference in the direction in which thermodynamical processes take place.
The second principle of thermodynamics accounts for this fact. And after temper-
atures are equal, no further process takes place; i.e. the entropy has reached its
13
The equation in index notation is σij vi,j and because of the symmetry of σ we have σij = σji
and σij vi,j = σji vi,j = σij vj,i (the last identity is just a renaming of indices i ↔ j). Then

σij vi,j = σij 12 (vi,j + vj,i ) = σ : sym (∇v)

14
In the case of fluids σ = −pI, where p is the fluid pressure (positive in compression as the
convention in fluid mechanics). Then σ : ε̇ = −p tr (ε̇) = −pV̇ (1/V ) and the first principle is
easily recognized in a typical form.

- 298-
C.1 Temperature effects

maximum value. The entropy is additive in the sense that there is an specific en-
tropy s (entropy per unit mass), so the total entropy S of a system is the extensive
variable Z
S= ρsdV (1021)
V
The entropy is assumed to be a function of the number of particles N of the system
(say mass in a continuum), of some extensive kinematic variables Ξ (for example
volume in a gas, infinitesimal deformations or deformation gradient in a solid), and
also an monotonically increasing function of the internal energy U for N, Ξ constant,
i.e.
∂S
S = S (N, Ξ, U ) with >0 (1022)
∂U
so it can be inverted to yield the following dependence

U = U (N, Ξ, S)

The second principle of thermodynamics states that among all possible states con-
sistent with the kinematic constraints, an isolated system acquires the state in which
the entropy is maximum. In an equivalent statement, we can say that the entropy
of an isolated system can never decrease in any process, it must increase or at most
remain the same; i.e.
Ṡ ≥ 0 for an isolated system (1023)
If we recall the two solids (label them A and B) such that TA > TB , then if N, Ξ
remain constant

∂SA ∂SB
Ṡ = U̇A + U̇B (1024)
∂U N,Ξ=const ∂U N,Ξ=const

Since the system is isolated such that by the first principle there is no change of the
total energy
U̇A + U̇B = 0 (1025)
we obtain in Eq. (1024)
!
∂SA ∂SB
Ṡ = − U̇A > 0 (1026)
∂U N,Ξ=const ∂U N,Ξ=const

Since TA > TB we expect A to get colder and then to decrease its internal energy,
i.e. U̇A < 0, so
∂SA ∂SB
< (1027)
∂U N,Ξ=const ∂U N,Ξ=const
and vice-versa, it we had TA < TB we would obtain the > sign in the previous
equation. Then we note that there is a clear dependence on the temperature but
with reversed tendency. Then we can propose

∂S 1
= (1028)
∂U N,Ξ=const T

- 299-
C Thermodynamics of solids

which fulfills the requirements15 . Conversely



∂U
=T (1029)
∂S N,Ξ=const

Then
∂U ∂U ∂U ∂U ∂U
U = U (N, Ξ, S) ⇒ U̇ = Ṅ + Ξ̇ + Ṡ = U̇ = Ṅ + Ξ̇ + T Ṡ (1030)
∂N ∂Ξ ∂S ∂N ∂Ξ
i.e. if N is constant (the number of particles is held constant as usually assumed in
solids)  
1 ∂U
Ṡ = U̇ − Ξ̇ (1031)
T ∂Ξ
But since Ξ contains the internal kinematic variables
∂U ∂U
P= Ξ̇ + K̇ ⇒ Ξ̇ = P − K̇ (1032)
∂Ξ ∂Ξ
and using the first principle
∂U
U̇ − Ξ̇ = U̇ + K̇ − P = Q (1033)
∂Ξ
Hence Eq. (1031) can be written as
Z Z
RS QRS 1 q · n̂
Ṡ = RS = ρrdV − dS (1034)
T V T S T

This is the external entropy input given by Q from a heat source at a reversible
source temperature T RS . However, we consider that the heat source has the same
temperature at the contacting points than the solid, hence the substitution of T RS
by T . Of course, if the process is not reversible, in general the entropy generated
Ṡ will be larger for example because some heat is lost and the QIR given by the
irreversible source may be larger than that taken by the solid QRS
Z Z
QIS RS QRS 1 q · n̂
Ṡ = ≥ Ṡ = RS = ρrdV − dS (1035)
T T V T S T

This equation is known as the Clausius-Planck inequality. Then we can define the
internal entropy production rate (note that the external source is reversible) as
Z Z
RS 1 q · n̂
Υ̇ = Ṡ − Ṡ = Ṡ− ρrdV + dS ≥ 0 (1036)
V T S T

Equation (1035) may be written in intensive form using Gauss theorem as


r 1 q
ṡ ≥ − ∇· (1037)
T ρ T
15
Of course any monotonically decreasing function of T would be a solution. But then we could
redefine T to obtain this same expression, which can be taken as a definition for the S, T couple.

- 300-
C.1 Temperature effects

Note that
q q qi,i T,i 1 1
∇· = = − qi 2 = ∇ · q − 2 q · ∇T (1038)
T T i,i T T T T
so
1
q · ∇T ≥ 0
ρT ṡ − (ρr − ∇ · q) − (1039)
T
Using Eq. (1017) to substitute the part in parenthesis
1
ρT ṡ − (ρu̇ − σ : ε̇) − q · ∇T ≥ 0 (1040)
T
and since in an intensive variable we do not consider N and we take for small strains
in solids Ξ = ε
∂u ∂u
u̇ (s, ε) = ε̇ + ṡ (1041)
∂ε ∂s
we have    
∂u ∂u 1
ρ T− ṡ + σ−ρ : ε̇ − q · ∇T ≥ 0 (1042)
∂s ∂ε T
This expression is often called the Coleman-Noll equation. Since the inequality must
hold for any arbitrary process, we can imagine processes where two of the addends
are zero, so we can conclude that the inequality must hold for each term separately.
Furthermore, in a reversible process the identity must hold for arbitrary variations
of s, ε, so we must have
∂u (s, ε)
T = (1043)
∂s
∂u (s, ε)
σ=ρ (1044)
∂ε
1
− q · ∇T ≥ 0 (1045)
T
The last identity implies a dependence of q on ∇T , being the simplest one Fourier’s
law
q = −K∇T (1046)
where K is the tensor of thermal conductivities. Equation (1043) yields again the
definition of the absolute temperature, Equation (1044) yields the typical definition
of the stress tensor in an adiabatic process (i.e. no entropy change). Then, the
tensor of elastic constants in an isentropic process is

∂ 2 u (s, ε)
C|s = ρ (1047)
∂ε∂ε
so
∂ 2 u (s, ε)
σ = σ 0 + C|s : (ε − ε0 ) + ρ (s − s0 ) (1048)
∂ε∂s
The tensor C|s contains the adiabatic elastic constants16 .
16
These are constants to be used in fast dynamics (impacts) and in low heat transfer conditions
(space). These constants differ substantially from the isothermal ones used for example in statics
or low speed dynamics with heat transfer.

- 301-
C Thermodynamics of solids

Helmholtz free energy


In order to change the internal variable from entropy to temperature, we define the
Helmholtz free energy by the Laurent-Legendre transformation17

ψ (ε, T ) = u − sT ⇒ ψ̇ = u̇ − sṪ − ṡT (1049)

Substituting u̇ in Eq. (1040)

1
−ρψ̇ − ρsṪ + σ : ε̇ − q · ∇T ≥ 0 (1050)
T
and using
∂ψ ∂ψ
ψ̇ = : ε̇ + Ṫ (1051)
∂ε ∂T
we obtain    
∂ψ ∂ψ 1
σ−ρ : ε̇ + ρ −s − Ṫ − q · ∇T ≥ 0 (1052)
∂ε ∂T T
which yields the isothermal definition of the stresses and a new definition of the
entropy18
∂ψ
σ=ρ (1053)
∂ε
∂ψ
s=− (1054)
∂T
The stress tensor given in Equation (1053) can be written as

∂ψ (ε, T ) ∂ 2 ψ (ε, T ) ∂ 2 ψ (ε, T )


σ=ρ = σ0 + ρ : (ε − ε0 ) + ρ (T − T0 ) (1055)
∂ε ∂ε∂ε ∂ε∂T
= σ0 + C : (ε − ε0 ) − β (T − T0 ) (1056)

where
∂ 2 ψ (ε, T ) ∂σ
C=ρ = (1057)
∂ε∂ε ∂ε
is the tensor of elastic constants (in this case isothermal elastic constants) and

∂ 2 ψ (ε, T ) ∂s
β = −ρ =ρ
∂ε∂T ∂ε
are the thermal stiffnesses. This expression is the Duhamel-Neumann law for a
thermoelastic solid, and may be inverted to give

ε = ε0 + S : (σ − σ 0 ) + α (T − T0 ) (1058)

where C = S−1 and β = Cα. The tensor S is the tensor of elastic compliances and
the tensor α is the tensor of thermal expansion coefficients.
17
The term free energy implies the energy that may be usable as work. Note that d (sT ) is heat.
18
Strains can be obtained in a similar way for isentropic and isothermal processes defining the
enthalpy (for the internal energy) and the Gibbs free energy (for the Helmholtz free energy).

- 302-
C.1 Temperature effects

We finally note that the quadratic coupled form of the free energy increment,
typically used in coupled linear finite elements is

∂ψ ∂ψ
ρψ (ε, T ) = ρψ0 + ρ : (ε − ε0 ) + ρ (T − T0 ) + ...
∂ε 0 ∂T 0

1 ∂ 2 ψ
+ (ε − ε0 ) : ρ : (ε − ε0 ) + ...
2 ∂ε∂ε 0

1 ∂ 2 ψ
+ (ε − ε0 ) : ρ (T − T0 ) + ...
2 ∂ε∂T 0

1 ∂ 2 ψ
+ (T − T0 ) ρ (T − T0 ) (1059)
2 ∂T ∂T 0
or

ρψ = ρψ0 + σ0 : (ε − ε0 ) − ρs0 (T − T0 ) + ...


1 1 c
+ (ε − ε0 ) : C : (ε − ε0 ) − (ε − ε0 ) : β (T − T0 ) − (T − T0 )2
2 2 T0
(1060)

where we have used the definition of heat capacity Eq. (1019) and Eq. (1028) so

∂ 2 ψ ∂s ∂s ∂u 1 ∂u 1
ρ =−ρ = ρ = ρ = c (1061)
∂T ∂T 0 ∂T 0 ∂u 0 ∂T 0
T0 ∂T 0 T0

In Eq. (1060) only the second line is regarded because usually only relative values
to state 0 are relevant.

Other thermodynamic potentials


Equation (1058) can be formally obtained defining the enthalpy and the Gibbs free
energy through the Legendre following transformations that gives complementary
stored energies
h (s, σ) = u (s, ε) − σ : ε (1062)
g (T, σ) = u (s, ε) − T s − σ : ε (1063)
so the basic state variables are the stresses and the dependent variables are the
strains. These functions are known as complementary energy functions. The fol-
lowing table summarizes dependencies in solids. It is left to the reader the task of
obtaining the work-conjugate variables following the previous Coleman-Noll proce-
dure.
Energy Variables Legendre transf. Work conjugate variables
∂u ∂u
Internal (s,ε) u T = σ=
∂s ∂s
∂ψ ∂ψ
Helmholtz free (T,ε) ψ = u − Ts s=− σ=
∂T ∂ε
∂h ∂h
Enthalpy (s,σ) h=u−σ :ε T = ε=−
∂s ∂σ
∂g ∂g
Gibbs free (T,σ) g =ψ−σ :ε s=− ε=−
∂T ∂σ

- 303-
C Thermodynamics of solids

Then we can write the typical expressions


Energy Variation
∂u ∂u
Internal u̇ = ṡ + : ε̇ = T ṡ + σ : ε̇
∂s ∂ε
∂ψ ∂ψ
Helmholtz free ψ̇ = Ṫ + : ε̇ = −sṪ + σ : ε̇
∂T ∂ε
∂h ∂h
Enthalpy ḣ = ṡ + : σ̇ = T ṡ − ε : σ̇
∂s ∂σ
∂g ∂g
Gibbs free ġ = Ṫ + : σ̇ = −sṪ − ε : σ̇
∂T ∂σ

The heat equation


We note that the previous equations –for example Equation (1060)– show that me-
chanical and thermal fields are fully coupled. However, it is frequent to decouple
both problems assuming that the mechanical power does not affect the tempera-
ture distribution. This is frequently the case in quasistatic problems (no relevant
dynamic effects) and in elastic problems (there is no plastic or viscous heating). In
such cases, using the first principle Eq. (10162 )
∂u ∂u
ρu̇ = ρr − ∇ · q + σ : ∇v ⇒ ρ : ε̇ + ρ Ṫ = ρr − ∇ · q + σ : ε̇ (1064)
∂ε ∂T
and using Eq. (1044) assuming there is no internal heat
∂u
ρ
Ṫ = −∇ · q (1065)
∂T
We note that if we consider K constant
∇ · q = qi,i = ∇ · (−K∇T ) = − (Kij T,j ),i = −Kij T,ij
Taking Fourier’s law Eq. (1046) with assumed constant coefficients and the definition
of heat capacity Eq. (1019), we arrive at
∂T
c = K : ∇∇T (1066)
∂t
This equation is known as the heat equation and, in essence, it is nothing else than
the second Fick law applied to the temperature, whereas Fourier’s law Eq. (1046) is
nothing else than the first Fick law applied to the temperature. Of course in thermal
equilibrium (t → ∞) Ṫ = 0 and K : ∇∇T = 0 is the equation to be solved19 .
The boundary value problem to be solved in order to obtain the temperature
field is

Find T in volume V surrounded  ∂T
 c ∂t − K : ∇∇T = 0 in V

by boundaries ST and Sq
(1067)
(ST ∪ Sq = S and ST ∩ Sq = ∅)   T = T̄ , prescribed in ST

such that q · n̂ = q̄n , prescr. in Sq
19
Equation K∇∇T = 0 can be frequently solved as ∇2 T = 0 using a transformation of coordi-
nates.

- 304-
C.2 Hygrometric effects

where q̄n is the prescribed heat flux through the boundary Sq and contains both
radiation and convection terms

q̄n = q̄nr + q̄nc = q̄nr − hc (T − Tc ) (1068)

where Tc is the reference ambient (sink) outside temperature and hc is the convective
heat transfer coefficient (similar to the thermal conductivity and that depends on the
surrounding fluid, boundary rugosity, etc). Note that the prescribed conductivity
depends on the temperature of the solid (this is a similar case to the follower forces
in mechanics), so it is also common to place it on the left-hand-side of the equation.
For this term, we have also applied the usual convention that the heat flux is positive
from hot to cold and negative if the solid loses heat.
We finally remark that in the case of isotropy, K = KI, where K is the isotropic
thermal conductivity and I is the identity tensor. Then Fourier’s law reduces to

q = −K∇T (1069)

and the heat equation to


∂T
c = K∇2 T (1070)
∂t
which is the best known expression. In the case of thermal equilibrium ∂T /∂t = 0
and we arrive at ∇2 T = 0.

C.2 Hygrometric effects


Moisture effects have different nature but similar consequences in solids. In this case
we are dealing with a loss of humidity (water contents) in the solid and, hence, a
decrease in volume (or vice-versa). Most materials can be considered insensitive ti
humidity because the gain/loss of humidity is negligible and/or humidity does not
suffer dramatic changes. However, other materials (think for example in clay-like
soils or concrete) are strongly influenced by moisture contents, not only in terms of
dilatation and weight but also in terms of material behavior (think again of soils or
of concrete). A full treatment of moisture effects is out of the scope of these notes,
so in the following paragraphs we only consider the dilatation effects of moisture
contents.

Fick’s laws
Fick’s diffusivity law is a consequence of the second principle of thermodynamics
when applied to any substance concentration. The thermodynamic treatment follows
similar steps as for the case of temperature, but in this case entropy is considered also
a function of moisture (humidity) concentration H, usually through the number of
particles N . Then, there is a moisture flux vector qH which depends on the moisture
gradient ∇H, going from higher to lower values, such that the first term of Taylor’s
expansion series is known as Fick’s first law

qH = −K H ∇H (1071)

- 305-
C Thermodynamics of solids

where K H is the tensor of moisture (mass) diffusion coefficients. Equation (1071)


has a clear parallelism with Fourier’s law, Equation (1046).
Any substance must also fulfill the equation of conservation of the substance (i.e.
conservation of mass). That is also the case for moisture. This equation reads
Z Z

H dV = − q H · n̂ dS (1072)
∂t V S

The left-hand-side of the equation is the change in the moisture (water) content in
the volume, whereas the right-hand-side is the moisture entering the system through
the boundary. Again the minus sign implies the usual convention that a gain of
humidity is obtained for q H · n̂ negative (moisture entering the system). Gauss
theorem and the localization theorem may be used to arrive at the intensive form,
known as Fick’s second law
∂H ∂H
= ∇ (K H ∇H) ⇒ = K H : ∇∇H (1073)
∂t ∂t
where we again assumed that the diffusion coefficients are constant.
The humidity field H (x, t) in a solid may be found solving the following bound-
ary value problem, mathematically identical to the heat problem

Find H in volume V surrounded  ∂H
 ∂t − K H : ∇∇H = 0 in V

by boundaries SH and SqH
(1074)
(SH ∪ SqH = S and SH ∩ SqH = ∅)   H = H̄, prescribed in SH

such that q H · n̂ = q̄Hn , prescr. in SqH

The equilibrium problem (t → ∞) is again obtained solving K H : ∇∇H = 0. How-


ever, note that in this case the moisture diffusion coefficients are frequently several
orders of magnitude smaller than thermal ones, so equilibrium will be obtained after
a much longer period (for example months instead of hours).

Volumetric hygroscopic dilatation


Once the humidity distribution H (x, t) in the solid is known, the volumetric increase
due to moisture contents can be computed. Full thermodynamical treatment in this
case is beyond the scope of these notes. However, following similar lines as for the
case of temperature, the following expression can be obtained —c.f. Eq. (1058):

ε = ε0 + S : (σ − σ 0 ) + α (T − T0 ) + αH (H − H0 ) (1075)

where αH are the coefficients of hygroscopic expansion and ε are the hygrothermo-
mechanical strains. This expression can be inverted to obtain a similar expression
to that of Eq. (1056)

σ = σ 0 + C : (ε − ε0 ) − β (T − T0 ) − β H (H − H0 ) (1076)

where β H = C : αH are the hygroscopic stiffnesses and σ are the hygrothermome-


chanical stresses. In incremental form:

∆σ = C : (∆ε − α∆T − αH ∆H) (1077)

- 306-
C.3 Piezoelectric and pyroelectric effects

Figure 125: Polarization of a PZT material. Left: tetragonal unit cell of lead titanate
above the Curie temperature; no dipole present. Right: same cell below the Curie
temperature showing a dipole because of the displacement of T i/Zr.

C.3 Piezoelectric and pyroelectric effects


From an electrical point of view, there are three types of materials: conductors
(with free charges as electrons in metals), semi-conductors (with charges that can
be freed after a threshold voltage) and dielectrics (no free charges as in ceramics).
The piezoelectric effect is present in some dielectrics. Piezoelectric sensors and
actuators are a special case of composite materials where the piezoelectric effect is
used. The piezoelectric effect is the relevant coupling that exist in some materials
between strains and electric charge. Ultrasonic transducers, quartz watches and
piezoelectric igniters are early applications of this effect. Piezoelectric actuators
are also a common application. In a similar way, the pyroelectric effect relates the
polarization charge to the changes in temperature.
We first introduce the polarization vector, then Gauss law for the electric field
(i.e. the first Maxwell law) and finally the thermodynamics of piezoelectricity.

Polarization vector
The piezoelectric and pyroelectric effects appear because of the existence of dipoles in
the solid. These dipoles are generated in some materials as PZT (P b[Zrx T i1−x ]O3
or Lead Zirconate Titanate) through a strong electric field just below the Curie
temperature (arround 300o C), see Figure 125. This is known as doping. The ar-
rangement in the solid orientates and stretches the cells in the preferred direction
marked by the imposed field. Since dipoles p~ are vectors (with units of C m), so is
the dipole surface density or polarization vector P (with units of C / m2 ) defined as
P
ip~i
P = lim (1078)
dV →0 dV

Over a given volume V , this dipole density or polarization vector yields an equivalent
bound charge 20 qp (units of C) that is computed over the surface S that encloses the
20
Dipoles inside the volume are cancelled-out, so only the bound dipoles contribute to the global
charge.

- 307-
C Thermodynamics of solids

Figure 126: Net polarization effect in the boundaries of a polarized solid.

volume, or using Gauss theorem (see Figure 126)


Z Z
qp = − P · n̂dS = − ∇ · P dV (1079)
S V

where the minus sign accommodates the usual sign convention of charges (note that
the polarization vector goes from negative to positive). Then, the equivalent bound
charge density ρp (units of C / m3 ) is defined as

ρp = −∇ · P (1080)

Gauss law for the electric field


Maxwell equations are a unified presentation of electromagnetism in four equations,
summarizing the works of Gauss, Coulomb, Faraday, Ampere among others. The
first one is Gauss’s equation for the electric field. It relates the free electric flux
φE over a closed surface S to the total charge q in the volume V enclosed by that
surface S (i.e. a charge generates an electric flux)

ǭ0 φE = q (1081)

where ǭ0 = 8.854 × 10−12 F / m (or equivalently C2 /(m2 N)) is the conversion con-
stant named permittivity in vacuum21 or “the” electric constant.
In our case, the charge can be divided in the usual free (body) charge qf and
the new piezoelectric (or boundary) charge qP , so q = qf + qp and we can define the
equivalent intensive quantity of free charge density ρf and electric field (strength)
E (units of V / m) such that
Z I
qf = ρf dV and φE = E · n̂ dS (1082)
V S
21
Recall that Coulomb’s law states that the load in vacuum between two charges q1 and q2
separated by a distance r is given by
1 q1 q2
F =
4πǭ0 r r

- 308-
C.3 Piezoelectric and pyroelectric effects

so
I Z
q qf + qp 1
φE = ⇒ φE = ⇒ E · n̂ dS = (ρf + ρp ) dV (1083)
ǭ0 ǭ0 S ǫ0 V

and the intensive form of Gauss’s law can be obtained using Gauss divergence the-
orem
1
∇·E = (ρf + ρq ) (1084)
ǭ0
Since E is an irrotational vector field, it can be thought of as the gradient of a scalar
field V (measured in Volts)
E = −∇V (1085)
which gives a clear meaning22 for E and which frequently simplifies the solution of
the problem through Poisson’s equation
1
∇2 V = − (ρf + ρq ) (1086)
ǭ0

Electric displacement
The electric charge displacement vector (units of C / m2 ) is defined as23
D = ǭ0 E + P (1087)
The relation between the polarization P and the electric field is given in linear
form
1
P = ǭ0 κE ⇒ E = κ −1 P (1088)
ǭ0
where κ is the tensor of electric susceptibilities (dimensionless) and ǫr = (I + κ) the
tensor of relative permittivities or relative dielectric coefficients. Then the electric
charge displacement vector is obtained as
D = ǭ0 E + ǭ0 κE = ǭ0 (I + κ) E = ǭ0 ǫr E (1089)
D = ǫE (1090)
where I is the identity tensor and ǫ = ǭ0 ǫr = ǭ0 (I + κ) is the second order tensor of
permittivities, or simply dielectric coefficients (units of F / m) of the solid. We note
that E is a easily measurable quantity (voltage divided by distance), but D is not
(charge per square meter). Hence E is preferable as a basic variable in constitutive
equations.
Hence Equation (1084) can be written in a different way. Taking the divergence
of the electric displacement and using Eq. (1087), Eq. (1080) and Eq. (1084)
∇ · D = ǭ0 ∇ · E + ∇ · P
= ǭ0 ∇ · E − ρp
 
1
= ǭ0 (ρf + ρq ) − ρp
ǭ0
22
If we apply a voltage V between two parallel plates separated a distance d, then E = V /d and
the direction is perpendicular to the plates, but from (+) to (−)
23
The physical meaning is to be the charge density in the (+) previous plate.

- 309-
C Thermodynamics of solids

so
∇ · D = ρf (1091)
is the equivalence of Gauss law for a dielectric and gives another physical meaning.
If there are no free charges in the dielectric

∇·D =0 (1092)

Finally, the electric power may be expressed as


Z Z Z
ẆD = E · Ḋ dV = ǭ0 E · Ė dV + E · Ṗ dV (1093)
V V V

and hence the energy


Z Z Z
1 1 1
WD = E · D dV = ǭ0 |E|2 dV + E · P dV = WE + WP (1094)
V 2 V 2 V 2
It is common to define the artificial internal pseudo-energy of the dielectric as

Ũ = U − WE (1095)

Piezoelectric effect
The piezoelectric effect is the appearance of changes in dimension j when a voltage
is applied in dimension i. The relation is given by the piezoelectric (strain) constants
dij
∆lj V
= di(jj) (no sum on j) (1096)
lj li
where li is the dimension of the specimen in direction i. Then since εjj = ∆lj /lj
(no sum on j) and Ei = V /li
εjj = Ei di(jj) (1097)
In a general way, the equation for the inverse (or converse) piezoelectric effect is

ε=E·d or εij = Ek dk(ij) (1098)

The tensor d is known as the tensor of piezoelectric (strain) constants or coefficients


or piezoelectric charge coefficients24 (units of m / V or C / N, see below). We used
parenthesis in indices to easy understanding of index pairs. The induced stresses
are (we use index notation)

σij = Cijkl εkl = Cij(kl) Em dm(kl) = Cij(kl) dm(kl) Em = e(ij)m Em (1099)

where e(ij)k is a third order tensor containing the piezoelectric moduli or piezoelectric
stress constants —units of N / (V m). It is customary to define dTk(ij) = d(ij)k so

σ = C : dT · E = e · E; with e = C : dT (1100)
24
Typical values for PZT polarized in z−axis are d3(11) = d3(22) = −171 × 10−12 C / N, d3(33) =
374 × 10−12 C / N, d3(12) = d3(21) = 584 × 10−12 C / N.

- 310-
C.3 Piezoelectric and pyroelectric effects

Another view of the piezoelectric effect is the direct piezoelectric effect. In this
case, if a pressure (load per unit surface) is applied at a face of a block of the
material, a charge per unit surface (dipole surface density of polarization vector)
appears. It can be shown that they are related by the same constants d

P =d:σ or Pi = di(jk)σjk (1101)

Then, the electric displacement vector is obtained from Eq. (1087) in the absence
of other effects.
D=d:σ (1102)

Pyroelectric effect
For the case of the pyroelectric effect, the change in the polarization vector for a
given temperature change (T − T0 ) is

P = p∆T (1103)

where p is the vector of pyroelectric coefficients with units of C / m2 K .

Thermodynamics and constitutive equations of piezoelectricity


In piezoelectricity we need to consider the energies given in Eq. (1094)

d
(U + K) = P + Q + WD + other possible powers (1104)
dt

The energies in this case can be formulated as shown in this table (note that
we use the symbol ǫ for the intensive internal energy, but the permittivities are a
tensor, hence in boldface):

Energy Variables Legendre transf.


Internal (s,ε, D) u
Helmholtz free (T,ε, E) ψ = u − Ts − E · D
Enthalpy (s,σ, D) h=u−σ :ε
Gibbs free (T,σ, E) g = u − Ts − σ : ε − E · D

The internal energy contains, again, a non-measurable quantity D, so it is cus-


tomary to apply a Legendre transformation to work with the measurable E. Then
in this case we consider the Helmholtz free energy

ψ (T,ε, E) = u − T s − E · D (1105)

Assuming linear behavior, we can apply the previous Coleman-Noll procedure as we


did in Section C.1. We leave the details to the reader. It is straightforward (but

- 311-
C Thermodynamics of solids

lengthy) to arrive at the following quadratic expression (we assume relative values
respect to the reference state, i.e. ε0 = 0, E 0 = 0, ...)

ψ = u − Ts − E · D
1 1 c
= ε : C : ε − ε : β (T − T0 ) − (T − T0 )2 ...
2 2 T0
1
−ε : e · E − E · ǭ E − p̄ · E (T − T0 ) (1106)
2
where the last line of the equation contains the new terms. In Eq. (1106) the
following identities which define work conjugacy are obtained from the procedure
∂ψ ∂ψ ∂ψ
σ= ; s=− ; D=− ; (1107)
∂ε ∂T ∂E
These expressions yield the following constitutive equations for the dependable vari-
ables

σ = C : ε − β (T − T0 ) − e · E (1108)
T
D = e : ε + ǭ E + p̄ (T − T0 ) (1109)
c
s = β : ε + p̄ · E + (T − T0 ) (1110)
T0
These equations may be partially inverted (or alternatively obtained using the Gibbs
free energy) to obtain the strain

ε = C−1 [σ + β (T − T0 ) + e · E]

= C−1 : σ + C−1 : β (T − T0 ) + C−1 : e · E
= S : σ + α (T − T0 ) + dT · E (1111)

which substituted in the other ones yield


 
D = eT : C−1 : σ + α (T − T0 ) + dT · E + ǭ E + p̄ (T − T0 )
 
= d : σ + eT : dT + ǭ · E + eT : α + p̄ (T − T0 )
= d : σ + ǫ̃ · E + p̃ (T − T0 ) (1112)

where ǫ̃ = eT : dT + ǭ is the tensor of dielectric permittivities for constant tem-


peratures and (“mechanical”) stresses 25 . The tensor ǭ is the tensor of dielectric
permittivities for constant
 temperatures and (“mechanical”) strains. The vectors p̄
and p̃ = eT : α + p̄ are similar vectors of pyroelectric coefficients for their respec-
tive cases. Finally the entropy equation may also be obtained in terms of stresses.
This equation is left to the reader.
Another common equation set is obtained expressing (ε, E, s) as a function of
(σ, D, T ). In this case we factor-out E from Eq. (1112)

E = ǫ̃−1 · D − ǫ̃−1 · d : σ − ǫ̃−1 · p̃ (T − T0 ) (1113)


25
Typical values for PZT are ǫ̃αα = 1730ǭ0 F / m, ǫ̃zz = 1700ǭ0 F / m and ǫ̃ij = 0 for i 6= j. These
values of course depend on doping.

- 312-
C.3 Piezoelectric and pyroelectric effects

and substitute into Eq. (1111)


 
ε = S : σ + α (T − T0 ) + dT · ǫ̃−1 · D − ǫ̃−1 · d : σ − ǫ̃−1 · p̃ (T − T0 ) (1114)

so
ε = SD : σ + gT · D + αD (T − T0 ) (1115)
where the following modified constants have been defined

SD = S − dT · ǫ̃−1 · d (1116)
αD = α − dT · ǫ̃−1 · p̃ (1117)
T T −1
g = d · ǫ̃ (1118)

The D superscript means that are quantities obtained if measured at constant elec-
tric displacement. The tensor g is the tensor of piezoelectric stress constants (units
of m2 / C) or piezoelectric voltage coefficients (because it also has units of V m / N)
and ǫ̃−1 is the tensor of impermittivities (m / F). With these definitions Eq. (1113)
can be written as
E = −g : σ + ǫ̃−1 · [D − p̃ (T − T0 )] (1119)
It is instructive to relate free energy changes for changes in uniaxial stresses or
strains and for changes in electric field. This relation gives us the energy conversion
from mechanical to electrical energies and vice-versa. For a given electrical field,
the complementary electric energy (we leave to the reader the task of obtaining this
expression directly from the enthalpy, but it is inconsequential for our purpose) is,
see Equation (1106)

1 1
WE = E · D = D · ǫ̃−1 D
2 2
1  1
= σ : d · ǫ̃−1 d : σ = σ(ij) d(ij)m ǫ̃−1
T
mn dn(kl) σ(kl) (1120)
2 2
1  1
= σ : g T · d : σ = σ(ij) g(ij)n dn(kl) σ(kl) (1121)
2 2
where we have used ǫ̃ for unrestrained specimens (complementary energy values) and
also employed definition. The linear elastic mechanical energy (whether Helmholtz
or enthalpy contribution, this is inconsequential in linear cases) is, see Eq. (1106)

1 1 1
WM = σ : ε = σ : S : σ = σ(ij) S(ij)(kl) σ(kl) (1122)
2 2 2
Hence, we can relate both energies as

2 WE d(ij)m ǫ̃−1
mn dn(kl) g(ij)n dn(kl)
k(ij)(kl) = = = (1123)
WM S(ij)(kl) S(ij)(kl)

If only z is the axis with a normal stress (σz 6= 0 but the rest is σij = 0), then

1 2
WM = S3333 σ33 ⇒ 12 S33 σ32 in Voigt notation (1124)
2

- 313-
C Thermodynamics of solids

The electric energy for the usual case of ǫ̃mn = 0 if m 6= n and d1(33) = d2(33) = 0
and d3(33) 6= 0

1
WE = d2(33)3 ǫ̃−1 2 1 2 −1 2
33 σ33 ⇒ 2 d33 ǫ̃33 σ3 in Voigt notation (1125)
2
so in Voigt notation, the square of the energy conversion factor is

2 d233
k33 = (1126)
S33 ǫ̃33

- 314-
References

References
[1] A.A. Griffith. The Phenomena of Rupture and Flow in Solids. Philosophical
Transactions of the Royal Society, 1920, 221A, pp. 163–198.

[2] B.D. Agarwal, L.L. Broutman. Analysis and Performance of Fiber Composites.
John Wiley & Sons, Willey Interscience, 1990.

[3] W.S. Slaughter. The Linearized Theory of Elasticity. Birkhäuser, 2002.

[4] R.M. Jones. Mechanics of Composite Materials. Taylor & Francis, 1999.

[5] R.M. Christensen. A Two-Property Yield, Failure (Fracture) Criterion for Ho-
mogeneous, Isotropic Materials. Journal of Engineering Materials and Technol-
ogy, 2004, vol. 126, pp. 45.

[6] R. Hill. The Mathematical Theory of Plasticity. Oxford University Press, 1950.

[7] S.W. Tsai. Strength Theories of Filamentary Structures. In Fundamental As-


pects of Fiber Reinforced Plastic Composites. Conference Proceedings. R.T.
Schwartz and H.S. Schwartz (Editors). Dayton, Ohio, May 1966. Willey In-
terscience, 1968, pp. 3–11.

[8] O. Hoffman. The Brittle Strength of Orthotropic Materials. Journal of Compos-


ite Materials, April 1967, pp. 200–206.

[9] S.W. Tsai, E.M. Wu. A General Theory of Strength for Anisotropic Materials.
Journal of Composite Materials, January 1971, pp. 58–80.

[10] R. Narayanaswami, H.M. Adelman. Evaluation of the Tensor Polynomial and


Hoffman Strength Theories for Composite Materials. Journal of Composite Ma-
terials, October 1977, pp. 366–377.

[11] S.W. Tsai. Structural Behavior of Composite Materials. NASA CR-71, July
1964.

[12] D.A. Hopkins, C.C. Chamis. A Unique Set of Micromechanical Equations for
High Temperature Metal Matrix Composites. In Testing Technology of Metal
Matrix Composites. ASTM STP 964. American Society for Testing and Mate-
rials, 1988, pp. 159–176.

[13] J.J. Hermans. The Elastic Properties of Fiber Reinforced Materials when the
Fibers are Aligned. Proceedings of the Koninklijke Nederlandse Akademie van
Wetenschappen. Amsterdam, 1967, Ser. B, Vol. 70, Num. 1, pp. 1–9.

[14] R. Hill. Theory of Mechanical Properties of Fibre-Strengthened Materials – III.


Self-Consistent Model. Journal of the Mechanics and Physics of Solids, August
1965, pp. 189–198.

[15] J.C. Halpin. Primer on Composite Materials: Analysis. Technomic, 1984.

- 315-
References

[16] B. Walter Rosen. Tensile Failure of Fibrous Composites. AIAA Journal, Novem-
ber 1964, pp. 1985–1991.

[17] J.N. Reddy. Mechanics of Laminated Composite Plates and Shells. CRC, 2004.

- 316-

S-ar putea să vă placă și