You are on page 1of 64

Analytical Methods of Predicting

Performance of Composite Materials

L. N. McCartney
Materials Division, National Physical Laboratory
Teddington, Middlesex, TW11 0LW, UK

Abstract This paper is a collection of various analytical methods


for predicting some of the properties of laminated fibre reinforced
composites that can be used when designing composite laminates, or
when validating numerical methods of estimating these properties.
To begin, convenient methods are given to estimate the properties
of undamaged single plies and undamaged symmetric laminates.
Methods of predicting fracture in homogenized anisotropic materi-
als are then described, which exploit some very useful properties
of orthogonal polynomials. Example solutions are given which are
compared with known accurate solutions. The problem is then con-
sidered of quantifying, using analytical methods, the dependence of
the effective thermoelastic properties of a damaged laminate on the
density of ply cracks in the 900 ply of a cross-ply laminate. Many
very useful inter-relationships are given showing how most of the ef-
fective properties of damaged laminates depend on a single damage
function. Some example predictions are given for a typical carbon
fibre reinforced laminate. Finally, a model is described for predict-
ing the progressive degradation of a unidirectional fibre reinforced
composite that is degraded by an aggressive environment causing
defect growth in the fibres and eventually the catastrophic failure
of the composite. It is also shown how the time dependence of
residual strength may be estimated. An example is given of a nor-
malised failure/time curve, and some associated residual strength
curves that can be the basis of design methods to avoid the failure
of composites that will be exposed to aggressive environments.

1 Introduction
Engineers responsible for the design and maintenance of composite struc-
tures will usually be involved with some form of finite element analysis
(FEA) so that the stress distributions within and the deformation of the

H. Altenbach, T. Sadowski (Eds.), Failure and Damage Analysis of Advanced Materials,


CISM International Centre for Mechanical Sciences
DOI 10.1007/978-3-7091-1835-1_4 © CISM Udine 2015
192 N. McCartney

structure can be estimated. The objective is usually to identify the hot-


spots where local stresses are raised to levels that are in danger of initiating
damage and structural failure, or to ensure that the deformation during ser-
vice does not exceed the design limits. A key requirement for finite element
analysis is a knowledge of all the relevant materials property data such as
the full range of elastic constants, and the thermal expansion coefficients of
laminated composites. For an orthotropic material there are twelve inde-
pendent thermoelastic constants whose values need to be known. Many of
the properties cannot easily be measured in the laboratory, and their values
are either guessed, or are estimated using a variety of predictive models as-
suming that fibre and matrix properties, or single ply properties are known.
A challenging objective for engineers is being able to design structural
components so that damage and failure can be avoided during service. In
structures the presence of stress hot-spots will generate localised damage
that can grow progressively as a result of stress increase, or because of fa-
tigue loading. The damage growth locally degrades the material properties
leading to load transfer in structures, and to the threat of catastrophic
failure. Dealing with the effects of localised damage is exceedingly diffi-
cult. One pragmatic approach is to try to design composite structures so
that damage formation is avoided. This approach is particularly useful for
structures that undergo cyclic loading as it will extend the fatigue damage
initiation phase of the component life.
The topics to be considered in this paper describe various recommended
methods for assessing the properties of composites, and the resistance of
laminated composites to damage formation, by considering a range of an-
alytical modelling methods that are built up from a knowledge of rein-
forcement and matrix geometry and properties, or from knowledge of the
geometry and properties of individual plies. The topics focus on compact
analytical formulae that can be used to provide quickly and reliably in-
formation on properties needed for FEA, and on damage resistance. In
addition, the focus is also on theoretical developments which have not been
published in the literature, other than through NPL Reports or Conference
Proceedings.

2 Properties of an Undamaged Lamina and Laminates


2.1 Notation for Properties of a Single Lamina
First of all, it is assumed that the fibres of the lamina are aligned exactly
in the axial direction forming what is known as a unidirectional fibre rein-
forced composite. Three different notations will now be introduced, which
describe the properties of unidirectional fibre reinforced composites. The
Analytical Methods of Predicting Performance… 193

Axial
direction (x1 )

Through-thickness
Fibre direction direction (x3 )
Transverse
direction (x2 )

Figure 1. Diagram showing method of defining principal directions and


coordinate system for a lamina where fibres are aligned in the axial direction

first two are common notations based on right-handed Cartesian coordinates


(x1 , x2 , x3 ) or (x, y, z) while the third, to be used in this paper, is a more
compact notation that will enable an immediate physical interpretation of
each property. Consider a single lamina of a composite material, as shown
in Fig. 1, where the fibres are aligned in the x1 -direction (or x-direction).
For this case the fibre direction also corresponds with the axial direction of
the lamina, while the x2 -direction (or y-direction) corresponds with the in-
plane transverse direction, and the x3 -direction (or z-direction) corresponds
with the through-thickness direction.
The thermoelastic constants are best defined with respect to stress-strain
relations (see Eqs. (1)-(3) below), and each material constant is described
as shown in Table 1. The thermal expansion coefficients are associated with
a temperature difference ΔT = T − T0 , where T is the uniform temperature
of the lamina and T0 is the uniform reference temperature of the lamina at
which all stresses and strains are zero, when the lamina is in an unloaded
state.
194 N. McCartney

Table 1. Thermoelastic constants of a single lamina


Common Compact
notation notation
Young’s modulus in fibre direction (axial / longi- E11 Exx EA
tudinal)
Young’s modulus in in-plane transverse direction E22 Eyy ET
Young’s modulus in through-thickness direction E33 Ezz Et
In-plane axial Poisson’s ratio ν12 νxy νA
Out-of-plane axial Poisson’s ratio ν13 νxz νa
Transverse Poisson’s ratio ν23 νyz νt
In-plane axial shear modulus μ12 μxy μA
Out-of-plane axial shear modulus μ13 μxz μa
Transverse shear modulus μ23 μyz μt
Axial thermal expansion coefficient α11 αxx αA
In-plane transverse thermal expansion coefficient α22 αyy αT
Through-thickness thermal expansion coefficient α33 αzz αt

2.2 Lamina Stress-Strain Relations


It is assumed that the loading of the lamina is such that the stress and
strain distributions are uniform everywhere within the lamina. Such a stress
and deformation state occurs when the external surfaces are subject to uni-
form applied tractions or linear displacements. The stress-strain relations
referred to the coordinates (x1 , x2 , x3 ) have the following orthotropic form:
1 ν21 ν31 σ12
ε11 = σ11 − σ22 − σ33 + α11 ΔT, ε12 = ,
E11 E22 E33 2μ12
ν12 1 ν32 σ13
ε22 = − σ11 + σ22 − σ33 + α22 ΔT, ε13 = , (1)
E11 E22 E33 2μ13
ν13 ν23 1 σ23
ε33 = − σ11 − σ22 + σ33 + α33 ΔT, ε23 =
E11 E22 E33 2μ23

When referred to the coordinates (x, y, z), the stress-strain relations are
written in the form
1 νyx νzx σxy
εxx = σxx − σyy − σzz + αxx ΔT, εxy = ,
Exx Eyy Ezz 2μxy
νxy 1 νzy σxz
εyy = − σxx + σyy − σzz + αyy ΔT, εxz = , (2)
Exx Eyy Ezz 2μxz
νxz νyz 1 σyz
εzz = − σxx − σyy + σzz + αzz ΔT, εyz =
Exx Eyy Ezz 2μyz
Analytical Methods of Predicting Performance… 195

Using the compact notation to be used in this paper, involving only inde-
pendent thermoelastic constants, the stress-strain relations are written:
1 νA νa τA
εA = σA − σT − σt + αA ΔT, γA = ,
EA EA EA μA
νA 1 νt τa
εT = − σA + σT − σt + αT ΔT, γa = (3)
EA ET ET μa
νa νt 1 τt
εt = − σA − σT + σt + αt ΔT, γt =
EA ET Et μt
The subscripts A, T and t attached to stresses, strains and properties indi-
cate parameters associated respectively with the axial, in-plane transverse
and through-thickness directions of the lamina. It should be noted that
the upper case subscripts A and T are associated only with in-plane direc-
tions and parameters, while the lower case subscripts are associated with
the through-thickness direction and parameters. The above three sets of
stress-strain relations are equivalent only if:

EA = E11 = Exx , ET = E22 = Eyy , Et = E33 = Ezz ,


μA = μ12 = μxy , μa = μ13 = μxz , μt = μ23 = μyz , (4)
αA = α11 = αxx , αT = α22 = αyy , αt = α33 = αzz

and
νa ν13 ν31 νxz νzx
= = = =
EA E11 E33 Exx Eyy
Et
so that ν13 = νxz = νa , ν31 = νzx = νa ,
EA
νt ν23 ν32 νyz νzy
= = = =
ET E22 E33 Eyy Ezz
(5)
Et
so that ν23 = νyz = νt , ν32 = νzy = νt ,
ET
νA ν21 ν12 νyx νxy
= = = =
EA E22 E11 Eyy Exx
ET
so that ν12 = νxy = νA , ν21 = νyx = νA
EA
It should be noted that
γA = 2ε12 = 2εxy , γa = 2ε13 = 2εxz , γt = 2ε23 = 2εyz ,
(6)
τA = σ12 = σxy , τa = σ13 = σxz , τt = σ23 = σyz

The parameters γA , γa , γt are known as engineering shear strains which are


twice the corresponding shear strains introduced when using tensor nota-
tion. It should be noted that just 9 lamina elastic properties appear in the
196 N. McCartney

compact version (3) of the stress-strain relations whereas 12 elastic prop-


erties appear when using the conventional approach of the relations (1) or
(2). The relations (5) indicate that the 9 lamina elastic properties of the
compact notation are independent, adding a further reason for their use in
this paper. The thermoelastic constants of individual plies in a laminate
are usually assumed to be transverse isotropic so that

Et = ET , νa = νA , μa = μA , αt = αT and ET = 2μt (1 + νt )

The number of independent thermoelastic constants then reduces from 12


to 7.

2.3 Inverted Form of Lamina Stress-Strain Relations


The inverted form of the stress-strain relations (3) may be written

σA = ĒA εA + ν̄A ĒT εT + ν̄a Ēt εt − ĒA ᾱA ΔT, τA = μA γ A ,


σT = ν̄A ĒT εA + ĒT εT + ν̄t Ēt εt − ĒT ᾱT ΔT, τa = μa γ a , (7)
σt = ν̄a Ēt εA + ν̄t Ēt εT + Ēt εt − Ēt ᾱt ΔT, τt = μt γ t ,

where


EA E
ĒA = 1 − νt2 t ,
Λ ET


ET E
ĒT = 1 − νa2 t ,
Λ EA


Et 2 ET
Ēt = 1 − νA ,
Λ EA


ET E
ν̄A ĒT = νA + νa νt t , (8)
Λ ET
Et
ν̄a Ēt = (νa + νt νA ) ,
Λ


Et ET
ν̄t Ēt = ν t + νa νA ,
Λ EA
Et Et 2 ET Et
Λ = 1 − νa2 − νt2 − νA − 2νa νt νA
EA ET EA EA

ĒA ᾱA = ĒA αA + ν̄A ĒT αT + ν̄a Ēt αt ,


ĒT ᾱT = ν̄A ĒT αA + ĒT αT + ν̄t Ēt αt , (9)
Ēt ᾱt = ν̄a Ēt αA + ν̄t Ēt αT + Ēt αt
Analytical Methods of Predicting Performance… 197

It should be noted that the elastic coefficients of the three stress-strain


relations (7) are symmetric as required.

2.4 Using the Contracted Notation for Tensors


The components of the stress and strain tensors are now assembled in
column vectors so that
⎡ ⎤ ⎡ ⎤ ⎡ ⎤ ⎡ ⎤
σA σ1 σ11 σxx
⎢ σT ⎥ ⎢ σ2 ⎥ ⎢ σ22 ⎥ ⎢ σyy ⎥
⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥
⎢ σt ⎥ ⎢ σ3 ⎥ ⎢ σ33 ⎥ ⎢ σzz ⎥
⎢ ⎥≡⎢ ⎥≡⎢ ⎥≡⎢ ⎥
⎢ τt ⎥ ⎢ σ4 ⎥ ⎢ σ23 ⎥ ⎢ σyz ⎥ ,
⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥
⎣ τa ⎦ ⎣ σ5 ⎦ ⎣ σ13 ⎦ ⎣ σxz ⎦
τA σ6 σ12 σxy
⎡ ⎤ ⎡ ⎤ ⎡ ⎤ ⎡ ⎤ (10)
εA ε1 ε11 εxx
⎢ εT ⎥ ⎢ ε2 ⎥ ⎢ ε22 ⎥ ⎢ εyy ⎥
⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥
⎢ εt ⎥ ⎢ ε3 ⎥ ⎢ ε33 ⎥ ⎢ εzz ⎥
⎢ ⎥≡⎢ ⎥≡⎢ ⎥≡⎢ ⎥
⎢ γt ⎥ ⎢ ε4 ⎥ ⎢ 2ε23 ⎥ ⎢ 2εyz ⎥
⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥
⎣ γa ⎦ ⎣ ε5 ⎦ ⎣ 2ε13 ⎦ ⎣ 2εxz ⎦
γA ε6 2ε12 2εxy
General linear elastic stress-strain relations, including thermal expansion
terms, have the contracted matrix form
⎡ ⎤ ⎡ ⎤⎡ ⎤ ⎡ ⎤
σA C11 C12 C13 C14 C15 C16 εA U1
⎢ σT ⎥ ⎢ C21 C22 C23 C24 C25 C26 ⎥ ⎢ εT ⎥ ⎢ U2 ⎥
⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥
⎢ σt ⎥ ⎢ C31 C32 C33 C34 C35 C36 ⎥ ⎢ εt ⎥ ⎢ U3 ⎥
⎢ ⎥=⎢ ⎥⎢ ⎥−⎢ ⎥
⎢ τt ⎥ ⎢ C41 C42 C43 C44 C45 C46 ⎥ ⎢ γt ⎥ ⎢ U4 ⎥ ΔT,
⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥
⎣ τa ⎦ ⎣ C51 C52 C53 C54 C55 C56 ⎦ ⎣ γa ⎦ ⎣ U5 ⎦
τA C61 C62 C63 C64 C65 C66 γA U6
(11)
where CIJ are elastic constants which are components of the second or-
der matrix C and where UI are thermal expansion constants which are
components of the vector U , the indexes I and J ranging from 1 to 6.
For orthotropic materials the stress-strain relations have the simpler matrix
form
⎡ ⎤ ⎡ ⎤⎡ ⎤ ⎡ ⎤
σA C11 C12 C13 0 0 0 εA U1
⎢ σT ⎥ ⎢ C21 C22 C23 0 0 0 ⎥ ⎢ ⎥ ⎢ ⎥
⎢ ⎥ ⎢ ⎥⎢ εT ⎥ ⎢ U2 ⎥
⎢ σt ⎥ ⎢ C31 C32 C33 0 0 0 ⎥⎢ εt ⎥ ⎢ U3 ⎥
⎢ ⎥ ⎢ ⎥⎢ ⎥−⎢ ⎥ ΔT (12)
⎢ τt ⎥=⎢ 0 0 0 C44 0 0 ⎥ ⎢ ⎥ ⎢ ⎥
⎢ ⎥ ⎢ ⎥⎢ γt ⎥ ⎢ 0 ⎥
⎣ τa ⎦ ⎣ 0 0 0 0 C55 0 ⎦⎣ γa ⎦ ⎣ 0 ⎦
τA 0 0 0 0 0 C66 γA 0
198 N. McCartney

By comparing (12) with (7) using (8) and (9), it follows that the non-zero
components of the C matrix are related to the elastic constants defined by
the stress-strain relations (3) as follows



EA Et ET Et
C11 = 1 − νt2 = ĒA , C12 = νA + νa νt = ν̄A ĒT ,
Λ ET Λ ET


Et ET Et
C13 = (νa + νt νA ) = ν̄a Ēt , C21 = νA + νa νt = ν̄A ĒT ,
Λ Λ ET



ET Et Et ET
C22 = 1 − νa2 = ĒT , C23 = νt + νa νA = ν̄t Ēt ,
Λ EA Λ EA


Et Et ET
C31 = (νa + νt νA ) = ν̄a Ēt , C32 = νt + νa νA = ν̄t Ēt ,
Λ Λ EA


Et 2 ET
C33 = 1 − νA = Ēt , C44 = μt , C55 = μa , C66 = μA
Λ EA
(13)
with
Et Et 2 ET Et
Λ = 1 − νa2 − νt2 − νA − 2νa νt νA
EA ET EA EA
The non-zero components of the U vector are related to the thermoelastic
constants defined by the stress-strain relations (3) as follows
U1 = ĒA ᾱA , U2 = ĒT ᾱT , U3 = Ēt ᾱt (14)
The inverse matrix form of (12) is of the form
⎡ ⎤ ⎡ ⎤⎡ ⎤ ⎡ ⎤
εA S11 S12 S13 0 0 0 σA αA
⎢ εT ⎥ ⎢ S21 S22 S23 0 0 0 ⎥ ⎢ σT ⎥ ⎢ αT ⎥
⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥
⎢ εt ⎥ ⎢ S31 S32 S33 0 0 ⎥⎢ σt ⎥ ⎢ αt ⎥
⎢ ⎥ ⎢ ⎥⎢ ⎥+⎢ ⎥ ΔT, (15)
⎢ γt ⎥=⎢ 0 0 0 S 0 0 ⎥ ⎢ τt ⎥ ⎢ 0 ⎥
⎢ ⎥ ⎢ 44 ⎥⎢ ⎥ ⎢ ⎥
⎣ γa ⎦ ⎣ 0 0 0 0 S55 0 ⎦⎣ τa ⎦ ⎣ 0 ⎦
γA 0 0 0 0 0 S66 τA 0
and a comparison with (3) shows that
1 νA νa
S11 = , S12 = − , S13 = − ,
EA EA EA
νA 1 νt
S21 = − , S22 = , S23 = − ,
EA ET ET
(16)
νa νt 1
S31 = − , S32 = − , S33 = ,
EA ET Et
1 1 1
S44 = , S55 = , S66 =
μt μa μA
Analytical Methods of Predicting Performance… 199

It is often useful to be able to estimate lamina properties from those of


the reinforcements and matrix that have been used to make the composite.
For multi-phase composites, where the matrix is reinforced with N different
types of fibre such that the fibre volume fraction of the i th fibre is denoted
by Vfi and that of the matrix by Vm , effective properties denoted with the
superscript eff may be estimated using the following formulae for transverse
isotropic materials derived using Maxwell’s methodology (Maxwell, 1873;
McCartney and Kelly, 2008; McCartney, 2010) (see also reference Hashin
(1983))
1 N
Vfi Vm
eff m
= f(i)
+ m m, (17)
κT + κT i=1 κT + κT
m κ T + κT

1  N
Vfi Vm
= + m , (18)
kTeff m
+ μT i
k + μT
i=1 T
m kT + μm
T

N
Vfi kTi νA
i
Vm k m ν m
+ m T A
i
k + μT
i=1 T
m kT + μm T
eff
νA = , (19)

N
Vfi kTi Vm kTm
+ m
i=1
kTi + μm
T kT + μm T
⎛  2 ⎞
 eff 2 f(i) f(i)
4kTeff νA μm 
N
⎜ f(i) 4kT νA μm ⎟
eff
EA + eff
= Vfi ⎝EA + f(i) ⎠
kT + μm i=1 kT + μm (20)


4k m ν 2 μm
+ Vm Em + mT m ,
kT + μm

1  N
Vfi Vm
= + m , (21)
μeff
A
m
+ μA i
μ + μA
i=1 A
m μ A + μm
A

1 N
Vfi Vm kTm μm
= + m , where μ∗m = T
, (22)
μeff
T

+ μm i
μ + μm
i=1 T
∗ μT + μ∗m kTm + 2μmT

eff
eff eff 4νA (αeff eff eff
T + νA αA ) = V E f αf + V E α
EA αA + 1 1 f A A m m m
eff
+
kT μm
f (23)
4νA (αfT + νAf f
αA ) 4ν (α + νm αm )
+Vf 1 1 + Vm m 1 m 1 ,
+ m +
kTf μm kT μm
200 N. McCartney
 
N
Vfi kTi αiT + νAi i
αA Vm kTm (αm m m
T + νA αA )
+
i=1
kTi + μm
T kTm + μm
T
αeff eff eff
T + νA αA = , (24)
N
Vfi kTi Vm k m
i m + m Tm
k + μT
i=1 T
kT + μT

where kTm = km + 13 μm . Since

eff 2 2
1 1 1 (νA ) νTeff 1 1 eff
4(νA )
= + + , = − − (25)
ETeff 4μeff
T 4kT
eff E eff
A ETeff 4μeff
T 4kT
eff EA
eff

It is possible to use the relations (17)-(24) to estimate values of ETeff and


νTeff .

2.5 Thermoelastic Constants for Angled Laminae


It is required now to consider how these properties are used to derive
those of angled plies. One very important parameter that must be clearly
defined is the angle φ defining the fibre directions in an angled lamina.
Figure 2 illustrates the definition that is used, and it should be noted that
this angle differs in sign from that used in some previous publications by
the author. The change of sign affects only the signs of shear coupling
parameters.
Figure 2 also defines the three orthogonal principal directions of the lam-
inate, and associates these directions with a right handed system (x1 , x2 , x3 )
of Cartesian coordinates defining a set of global axes. The angle φ defining
the fibre direction of the lamina is measured from the global x1 -axis in a
clockwise direction when viewing from a point situated on the negative part
of the global x3 -axis.
Following rotation of the lamina the stress-strain equations, including
thermal expansion terms, relating global strain components to global stress
Analytical Methods of Predicting Performance… 201

Fibre direction Axial


direction (x1 )


x1 Through-thickness

direction (x3 )
x3
Transverse
φ  direction (x2 )
x2 Global axes
Local axes

Figure 2. Diagram showing method of defining principal directions and


coordinate system for an angled lamina, and the angle φ specifying the
fibre direction

components are given, using matrix notation, by

⎡ ⎤ ⎡ ⎤
εA ε11
⎢ εT ⎥ ⎢ ε22 ⎥
⎢ ⎥ ⎢ ⎥
⎢ εt ⎥ ⎢ ε33 ⎥
⎢ ⎥≡ ⎢ ⎥
⎢ γt ⎥ ⎢ 2ε23 ⎥
⎢ ⎥ ⎢ ⎥
⎣ γa ⎦ ⎣ 2ε13 ⎦
γA 2ε12
⎡ ⎤⎡ ⎤ ⎡ ⎤
S11 S12 S13 0 0 S16 σA V1
⎢ S12 S22 S23 0 0 S26 ⎥⎢ σT ⎥ ⎢ V2 ⎥
⎢ ⎥⎢ ⎥ ⎢ ⎥
⎢ S13 S23 S33 0 0 S36 ⎥⎢ σt ⎥ ⎢ V3 ⎥
=⎢

⎥⎢
⎥⎢
⎥+⎢
⎥ ⎢
⎥ ΔT,

⎢ 0 0 0 S44 S45 0 ⎥⎢ τt ⎥ ⎢ 0 ⎥
⎣ 0 0 0 S45 S55 0 ⎦⎣ τa ⎦ ⎣ 0 ⎦
S16 S26 S36 0 0 S66 τA V6
(26)
where on setting m = cos φ and n = sin φ it can be shown that the co-
efficients in these stress-strain relations are related to the thermoelastic
202 N. McCartney

constants of the lamina as follows:




1 2νA m4 n4
S11 = m2 n2 − + + , (27)
μA EA EA ET


1 1 1   νA
2 2
S12 = m n + − − m4 + n 4 , (28)
EA ET μA EA
ν ν
S13 = − m2 a − n2 t , (29)
EA ET


  1 2ν 2m2 2n2
S16 = − mn m2 − n2 − A − + , (30)
μA EA EA ET


1 2ν m4 n4
S22 = m2 n2 − A + + , (31)
μA EA ET EA
ν ν
S23 = −m2 t − n2 a , (32)
ET EA


 2  1 2νA 2m2 2n2
S26 = mn m − n 2
− − + , (33)
μA EA ET EA
1
S33 = , (34)
Et


νt ν
S36 = 2mn − a , (35)
ET EA
m2 n2
S44 = + , (36)
μt μa


1 1
S45 = mn − , (37)
μa μt
m2 n2
S55 = + , (38)
μa μt

 2 2
1 1 2ν m − n2
S66 = 4m2 n2 + + A + , (39)
EA ET EA μA
V1 = m2 αA + n2 αT ,
V2 = m2 αT + n2 αA ,
(40)
V3 = αt , V4 = 0, V5 = 0
V6 = 2mn (αA − αT )
It should be noted that the expressions (26)-(40) reduce to (15) and (16)
when φ = 0 so that m = 1 and n = 0.
Analytical Methods of Predicting Performance… 203

2.6 Inverse Approach


The inverted form for the stress-strain relations of an angled ply is given
by
⎡ ⎤ ⎡ ⎤
σA σ11
⎢ σT ⎥ ⎢ σ22 ⎥
⎢ ⎥ ⎢ ⎥
⎢ σt ⎥ ⎢ σ33 ⎥
⎢ ⎥ ≡⎢ ⎥
⎢ τt ⎥ ⎢ σ23 ⎥
⎢ ⎥ ⎢ ⎥
⎣ τa ⎦ ⎣ σ13 ⎦
τA σ12
⎡ ⎤⎡ ⎤ ⎡ ⎤
C11 C12 C13 0 0 C16 εA U1
⎢ C12 C22 C23 0 0 C26 ⎥⎢ εT ⎥ ⎢ U2 ⎥
⎢ ⎥⎢ ⎥ ⎢ ⎥
⎢ C13 C23 C33 0 0 C36 ⎥⎢ εt ⎥ ⎢ U3 ⎥
=⎢

⎥⎢
⎥⎢
⎥−⎢
⎥ ⎢
⎥ ΔT,

⎢ 0 0 0 C44 C45 0 ⎥⎢ γt ⎥ ⎢ 0 ⎥
⎣ 0 0 0 C45 C55 0 ⎦⎣ γa ⎦ ⎣ 0 ⎦
C16 C26 C36 0 0 C66 γA U6
(41)
where
4 4
 
C11 = m Ē A+ n  ĒT + 2m2 n2 ν̄A ĒT + 2μA , 
m2 + n ν̄2A ĒT + m n ĒA + ĒT − 4μA ,
4 4 2 2
C12 =
C13 = m ν̄ a + n ν̄t Ēt ,  2  
C16 = mn m2 ĒA − n2 ĒT − m  − n2 ν̄A ĒT + 2μA ,
4 4 2 2
C22 = m Ē T + n ĒA + 2m n ν̄A ĒT + 2μA ,
2 2
C23 = m ν̄ t + n ν̄a Ēt ,   
C26 = mn n2 ĒA − m2 ĒT + m2 − n2 ν̄A ĒT + 2μA , (42)
C33 = Ēt
C36 = mn (ν̄a − ν̄t ) Ēt ,
C44 = m 2 μt + n 2 μa ,
C45 = mn (μa − μt ) ,
C55 = m 2 μa + n 2 μt ,
   2
C66 = m2 n2 ĒA + ĒT − 2ν̄A ĒT + m2 − n2 μA ,

and
U1 = m2 ĒA ᾱA + n2 ĒT ᾱT ,
U2 = n2 ĒA ᾱA + m2 ĒT ᾱT ,
(43)
U3 = Ēt ᾱt, 
U6 = mn ĒA ᾱA − ĒT ᾱT

It should be noted that (42) and (43) reduce to (13)-(14) when φ = 0 so


that m = 1 and n = 0.
204 N. McCartney

2.7 Shear Coupling Parameters and Reduced Stress-Strain Re-


lations
The final task of Sect. 2 is to introduce the concept of shear coupling and
to define the thermoelastic constants that characterise the phenomenon. It
can be shown that
(P ) (P ) (P )
1 νA νa λA (P )
εA = σ −
(P ) A
σ −
(P ) T
σ −
(P ) t
τ
(P ) A
+ αA ΔT,
EA EA EA μA
(P ) (P ) (P )
νA 1 νt λT (P )
εT = − σ +
(P ) A
σ −
(P ) T
σ −
(P ) t
τ
(P ) A
+ αT ΔT,
EA ET ET μA
(P ) (P ) (P )
(44)
νa νt 1 λt (P )
εt = − (P )
σA − (P )
σT + (P )
σt − τ
(P ) A
+ αt ΔT,
EA ET Et μA
(P ) (P ) (P )
λA λT λt 1 (P )
γA = − σ −
(P ) A
σ −
(P ) T (P )
σt + τ
(P ) A
+ αS ΔT,
μA μA μA μA

where the superscript P denotes that the thermoelastic constant refers to


an angled ply. By comparing (26) and (44) it is clear that
(P ) (P ) (P )
1 νA νa λA
S11 = (P )
, S12 = − (P )
, S13 = − (P )
, S16 = − (P )
,
EA EA EA μA
(P ) (P )
1 νt λT 1
S22 = (P )
, S23 = − (P )
, S26 = − (P )
, S33 = (P )
,
ET ET μA Et (45)
(P )
λt 1
S36 = − (P )
, S66 = (P )
,
μA μA
(P ) (P ) (P ) (P )
V1 = αA , V2 = αT , V3 = αt , V6 = αS

The parameters λA , λT , λt are dimensionless shear coupling properties as


they characterise the coupling of the shear stress τA to the non-shear strains
εA , εT and εt . The parameter αS characterises a shear deformation response
to temperature changes. When the fibres are aligned in the axial and in-
plane transverse directions all four parameters have zero values. It should
be noted that the signs of the shear coupling parameters λA , λT , λt and the
expansion coefficient αS depend upon the sign of the orientation angle φ,
indicating why it is essential to define exactly how this angle is defined.
From (44)4
(P ) (P ) (P ) (P ) (P ) (P )
τA = λA σA + λT σT + λt σt + μA γA − μA αS ΔT, (46)
Analytical Methods of Predicting Performance… 205

and on substituting in the remaining relations of (44)

(P ) (P )
(P ) 1 ν̃A ν̃a (P )
ε̃A ≡ εA + λA γA = σ −
(P ) A
σ −
(P ) T (P )
σt + α̃A ΔT,
ẼA ẼA ẼA
(P ) (P )
(P ) ν̃A 1 ν̃t (P )
ε̃T ≡ εT + λT γA = − σ +
(P ) A
σ −
(P ) T (P )
σt + α̃T ΔT, (47)
ẼA ẼT ẼT
(P ) (P )
(P ) ν̃a ν̃t 1 (P )
ε̃t ≡ εt + λt γA = − σ −
(P ) A (P )
σT + (P )
σt + α̃t ΔT,
ẼA ẼT Ẽt

where

1 1  2 1 1 1  2 1
(P ) (P )
(P )
= (P )
− λA (P )
, (P )
= (P )
− λT (P )
,
ẼA EA μA ẼT ET μA
1 1  2 1 (P )
ν̃t
(P )
νt
(P ) (P )
λ t λT
(P )
(P )
= (P )
− λt (P )
, (P )
= (P )
+ (P )
,
Ẽt Et μA ẼT ET μA
(P ) (P ) (P ) (P ) (P ) (P ) (P ) (P ) (48)
ν̃A νA λA λT ν̃a νa λt λA
(P )
= (P )
+ (P )
, (P )
= (P )
+ (P )
,
ẼA EA μA ẼA EA μA
(P ) (P ) (P ) (P ) (P ) (P ) (P ) (P )
α̃A = αA + λA αS , α̃T = αT + λT αS ,
(P ) (P ) (P ) (P )
α̃t = αt + λt αS

The relations (47) are known as the reduced stress-strain relations for the
angled lamina as they have exactly the same form as three of the stress-
strain relations (3) which apply when φ = 0.

2.8 Mixed Form of Stress-Strain Relations


The final steps of this Sect. 2 are to manipulate the stress-strain equa-
tions (46) and (47) so that they are in a form that will be useful when
considering the effective properties of laminates. The objective is to express
stresses and strains in terms of the parameters εA , εT , γA , σT and ΔT which
will have the same values in all plies of a laminate. It can be shown that

(P ) (P ) (P ) (P ) (P )
σA = Ω11 εA + Ω12 εT + Ω13 σt + Ω16 γA − ω1 ΔT,
(P ) (P ) (P ) (P ) (P )
σT = Ω12 εA + Ω22 εT + Ω23 σt + Ω26 γA − ω2 ΔT,
(P ) (P ) (P ) (P ) (P ) (49)
εt = −Ω13 εA − Ω23 εT + Ω33 σt − Ω36 γA + ω3 ΔT,
(P ) (P ) (P ) (P ) (P )
τA = Ω16 εA + Ω26 εT + Ω36 σt + Ω66 γA − ω6 ΔT,
206 N. McCartney

where
(P ) (P ) (P ) (P )
(P ) ẼA (P ) ν̃ Ẽ (P ) ν̂a
Ω11 = (P )
, Ω12 = A (P T) , Ω13 = (P ) ,
Ψ Ψ Ψ
(P ) (P ) (P ) (P )
(P ) ẼA (P ) ν̃A ẼT (P ) (P ) ẼT
Ω16 = (P ) λA + λ T , Ω 22 = ,
Ψ Ψ(P ) Ψ(P )
(P ) (P ) (P ) (P )
(P ) ν̂ (P ) ν̃ Ẽ (P ) Ẽ (P )
Ω23 = t(P ) , Ω26 = A (P T) λA + T(P ) λT , (50)
Ψ Ψ Ψ
(P ) (P )
(P ) Λ (P ) (P ) ν̂a (P ) ν̂t (P ) (P )
Ω33 = , Ω 36 = λ + λ + λt ,
Ψ(P ) Ẽt
(P ) Ψ(P ) A Ψ(P ) T
(P ) (P )
(P ) ẼA (P ) (P ) ẼT (P ) (P ) (P )
Ω66 = λ λ̂ + (P ) λT λ̂T + μA ,
Ψ(P ) A A Ψ
(P ) (P ) (P )
(P ) ẼA (P ) ν̃ Ẽ (P )
ω1 = (P )
α̃A + A (P T) α̃T ,
Ψ Ψ
ẼT  (P ) 
(P )
(P ) (P ) (P )
ω2 = α̃T + ν̃ A α̃A ,
Ψ(P )
(51)
(P ) (P )
(P ) ν̂a (P ) ν̂ (P ) (P )
ω3 = α̃ + t(P ) α̃T + α̃t ,
Ψ(P ) A Ψ
(P ) (P )
(P ) ẼA (P ) (P ) ẼT (P ) (P ) (P ) (P )
ω6 = λ̂ α̃ + (P ) λ̂T α̃T + μA αS ,
Ψ(P ) A A Ψ
and where
 2 Ẽ (P )
(P )
Ψ(P ) = 1 − ν̃A T
(P )
,
ẼA
(P ) (P ) (P )
ν̂a = ν̃a(P ) + ν̃t ν̃A , (52)
(P )
(P ) (P ) (P ) ẼT
ν̂t = ν̃t + ν̃a(P ) ν̃A (P )
ẼA
(P ) (P )
Ẽt (P ) Ẽt (P )
Λ(P ) = Ψ(P ) − ν̃a(P ) ν̂ (P ) − ν̃t
(P ) a
ν̂
(P ) t
ẼA ẼT
 2 Ẽ (P )  2 Ẽ (P )  2 (P )
ẼT
(P ) (P )
= 1 − ν̃a(P ) t
(P )
− ν̃ t
t
(P )
− ν̃ A (P )
(53)
ẼA ẼT ẼA
(P )
(P ) (P ) Ẽt
−2ν̃a(P ) ν̃t ν̃A (P )
,
ẼA
(P )
(P ) (P ) (P ) (P ) ẼT (P ) (P ) (P ) (P )
λ̂A = λA + λT ν̃A (P )
, λ̂T = λT + λA ν̃A (54)
ẼA
Analytical Methods of Predicting Performance… 207

x1
h
x1 = L

Plane External
symmetry surface

h1 h2 hi hn

Axial
direction
Through-thickness
direction

x3
(1) (i−1) (i) (n−1) (n)
0 x3 x(2)
3 x3 x3 x3 x3
Figure 3. Schematic diagram of geometry for one half of a general sym-
metric laminate

2.9 Effective Thermoelastic Properties of Undamaged Symmet-


ric Laminates
Consider now a laminate that is made by perfectly bonding together
various laminae (i.e. plies) having different orientations so that there are no
defects (i.e. the laminate is undamaged). The situation under consideration
concerns the deformation of a symmetric multi-layered laminate of total
thickness 2h constructed of 2n perfectly bonded plies that can have any
combination of orientations, such that symmetry about the mid-plane of the
laminate is preserved. The plies in each half of the laminate can be made
of different materials and each can have a different thickness as illustrated
in Fig. 3. As laminate symmetry is assumed, it is necessary to consider
only the right hand set of n layers (see Fig. 3). A global right handed
set of Cartesian coordinates is chosen having the origin at the centre of
the mid-plane of the laminate. The x1 -direction defines the longitudinal
or axial direction, the x2 -direction defines the in-plane transverse direction
and the x3 -direction defines the through-thickness direction. The locations
of the n − 1 interfaces in one half of the laminate (x3 > 0) are specified by
208 N. McCartney
(i)
x3 = x3 , i = 1, . . . , n − 1. The mid-plane of the laminate is specified by
(0) (n)
x3 = x3 = 0 and the external surface by x3 = x3 = h where h is the
half-thickness of the laminate. The thickness of the ith layer is denoted by
(i) (i−1)
hi = x3 − x3 such that


n
h= hi . (55)
i=1

Stress, strain and displacement components, and material properties asso-


ciated with the ith layer are denoted by a superscript (i). The orientation of
the ith layer is specified by the angle φi (measured clockwise when looking
in the direction of positive values of x3 ) between the x1 -axis and the fibre
direction of this layer (see Fig. 2). The representative volume element of
the laminate to be considered is specified by |x1 | ≤ L, |x2 | ≤ W, |x3 | ≤ h.
It is now required to determine the effective laminate properties S and
V in terms of the Young’s and in-plane shear moduli, Poisson’s ratios and
thermal expansion coefficients of the laminate. This is achieved by modify-
ing the relations (26) derived for a single angled lamina so that they apply
to a laminate rather than to an individual angled ply. When the laminate is
considered as a homogenised plate, the effective stress-strain relations must
be of the following form analogous to the corresponding single ply relations

⎡ ⎤ ⎡ ⎤
εA ε11
⎢ εT ⎥ ⎢ ε22 ⎥
⎢ ⎥ ⎢ ⎥
⎢ εt ⎥ ⎢ ε33 ⎥
⎢ ⎥ ≡⎢ ⎥
⎢ γt ⎥ ⎢ 2ε23 ⎥
⎢ ⎥ ⎢ ⎥
⎣ γa ⎦ ⎣ 2ε13 ⎦
γA 2ε12
⎡ ⎤
(L)
S11 S12
(L) (L)
S13 0 0 S16 ⎡
(L) ⎤ ⎡ (L) ⎤
σA V1
⎢ (L) (L) ⎥
⎢ S12 (L)
S22
(L)
S23 0 0 S26 ⎥⎢ ⎥ ⎢ (L) ⎥
⎢ (L) (L) ⎥⎢
σT ⎥ ⎢ V2 ⎥
⎢S (L) (L)
S36 ⎥ ⎢ ⎥ ⎢ (L) ⎥
=⎢
S23 S33 0 0 ⎥⎢ σt ⎥+⎢ V3 ⎥ ⎥ΔT,
⎥ ⎢
13
⎢ 0 0 ⎥⎢⎥ ⎥
⎥ ⎢
(L) (L) τt
⎢ 0 0 S44 S45 ⎢ 0 ⎥
⎢ (L) (L) ⎥⎣ τa ⎦ ⎢
⎣ 0 ⎦
⎣ 0 0 0 S45 S55 0 ⎦
(L) (L) (L) (L) τA V6
(L)
S16 S26 S36 0 0 S66
(56)
where the superscript (L) is used to denote effective thermoelastic constants,
Analytical Methods of Predicting Performance… 209
and where
(L) (L) (L)
(L) 1 (L) νA (L) νa (L) λA
S11 = (L)
, S12 = − (L)
, S13 = − (L)
, S16 = − (L)
,
EA EA EA μA
(L) (L)
(L) 1 (L) νt (L) λT (L) 1
S22 = (L)
, S23 = − (L)
, S26 = − (L)
, S33 = (L)
,
ET ET μA Et
(L)
(L) λt (L) 1 (L) (L) 1 (L) 1
S36 = − (L)
, S44 = (L) , S45 = Φ(L) , S55 = (L) , S66 = (L)
,
μA μt μa μA
(L) (L) (L) (L) (L) (L) (L) (L)
V1 = αA , V2 = αT , V3 = αt , V6 = αS
(57)
It can be shown that the laminate properties appearing in (57) are calculated
using the following relations
 2  2
(L) (L)
1 1 λA 1 1 λT
(L)
= (L)
+ (L)
, (L)
= (L)
+ (L)
,
EA ẼA μA ET ẼT μA
 2  
(L)
1 1 λt (L)
ν̃t λt λT
(L) (L)
(L) (L)
(L)
= (L)
+ (L)
, νt = ET (L)
− (L)
,
Et Ẽt μA ẼT μA
   
(L) (L) (L) (L) (L) (L)
(L) (L) ν̃A λA λ T (L) ν̃a λ t λA
νA = EA (L)
− (L)
, νa(L) = EA (L)
− (L)
,
ẼA μA ẼA μA
(L) (L) (L) (L) (L) (L) (L) (L)
αA = α̃A − λA αS , αT = α̃T − λT αS ,
(L) (L) (L) (L)
αt = α̃t − λ t αS ,
(58)
where
 2  2
(L) (L)
Ω12 (L)
Ω12 Ω12
(L) (L) (L) (L) (L)
ẼA = Ω11 − (L)
, ν̃A = (L)
, ẼT = Ω22 − (L)
,
Ω22 Ω22 Ω11
(L)
(L) (L) (L) (L) (L) (L) ẼT (L)
ν̃a(L) = Ω13 − ν̃A Ω23 , ν̃t = Ω23 − ν̃A Ω ,
(L) 13
ẼA
 2  2
(L) (L) (L) (L)
1 ν̃a (L) ν̃a ν̃t (L) ν̃t (L) (L)
(L)
= (L)
Ω11 + 2 (L) (L) Ω12 + (L)
Ω22 + Ω33 ,
Ẽt ẼA ẼA ẼT ẼT
(59)
(L) (L)
(L) 1 (L) ν̃A (L) (L) 1 (L) ν̃A (L)
λA = Ω −
(L) 16
Ω , λT
(L) 26
= Ω −
(L) 26 (L)
Ω16 , (60)
ẼA ẼA ẼT ẼA
210 N. McCartney
(L) (L) (L) (L) (L) (L)
λt = Ω36 − Ω13 λA − Ω23 λT , (61)
 2  2
(L) (L) (L) (L) (L) (L) (L) (L) (L)
μA = Ω66 − Ω11 λA − 2Ω12 λA λT − Ω22 λT (62)

(L) (L)
(L) 1 (L) ν̃A (L) (L) 1 (L) ν̃A (L)
α̃A = ω
(L) 1
− ω , α̃T
(L) 2
= ω
(L) 2
− (L)
ω1 , (63)
ẼA ẼA ẼT ẼA

(L) (L) (L) (L) (L) (L)


α̃t = ω3 − Ω13 α̃A − Ω23 α̃T ,
(L) 1  (L) (L) (L) (L) (L)
 (64)
αS = (L) ω6 − Ω16 α̃A − Ω26 α̃T
μA
(L) (L)
The quantities ΩIJ and ωI are calculated using the following summations

1 1
n n
(L) (i) (L) (i)
Ω11 = hi Ω11 , Ω12 = hi Ω12 ,
h i=1 h i=1
1 1
n n
(L) (i) (L) (i)
Ω22 = hi Ω22 , Ω33 = hi Ω33 ,
h i=1 h i=1
(65)
1 1
n n
(L) (i) (L) (i)
Ω16 = hi Ω16 , Ω26 = hi Ω26 ,
h i=1 h i=1
1
n
(L) (i)
Ω66 = hi Ω66 ,
h i=1

1 1 1
n n n
(L) (i) (L) (i) (L) (i)
Ω13 = hi Ω13 , Ω23 = hi Ω23 , Ω36 = hi Ω36 , (66)
h i=1 h i=1 h i=1

1 1
n n
(L) (i) (L) (i)
ω1 = hi ω 1 , ω 2 = hi ω 2 ,
h i=1 h i=1
(67)
1 1
n n
(L) (i) (L) (i)
ω3 = hi ω 3 , ω 6 = hi ω 6
h i=1 h i=1

These relations are derived by applying the following expressions for effective
stresses and strains to the relations (50) that apply to individual plies

1 1 1 1
n n n n
(i) (i) (i) (i)
σA = hi σA , σT = hi σT , τA = hi τA , εt = hi ε t
h i=1 h i=1 h i=1 h i=1
(68)
Analytical Methods of Predicting Performance… 211
The effective properties characterising the out-of-plane shear of the laminate
are given by
 
1
n
(L) m2i n2i 1
S44 = hi (i)
+ (i) = (L) ,
h i=1 μt  μa μt

1
n
(L) 1 1
S45 = h i mi n i (i)
− (i) = Φ(L) , (69)
h i=1 μ
 a μt
(L) 1 n
mi 2
ni2
1
S55 = hi (i)
+ (i) = (L) ,
h i=1 μa μt μa

where

1 1 1
n n n
(L) (i) (L) (i) (L) (i)
S44 = hi S44 , S45 = hi S45 , S55 = hi S55 (70)
h i=1 h i=1 h i=1

3 Fracture in Homogenised Anisotropic Materials


This Section will introduce a very convenient and elegant method of solving
crack problems in plates based on orthogonal polynomials. The methodol-
ogy will then be applied to bridged cracks, especially those arising in lami-
nates where ply cracks are confined to a limited number of plies (usually the
90o plies). The use of orthogonal functions enables very simple expressions
for stress intensity factors and energy release rates. The first objective is to
consider a very useful method of analysing isolated crack problems for com-
posite materials that have been homogenised as an anisotropic continuum.

3.1 Stress-Strain Relations


It is assumed that there is no through-thickness loading and that the
axial loading direction is parallel to the y-axis. The plate is assumed to be
orthotropic having the following stress-strain relations

εxx = a11 σxx + a12 σyy + a16 σxy ,


εyy = a12 σxx + a22 σyy + a26 σxy , (71)
2εxy = a16 σxx + a26 σyy + a66 σxy ,

where
1 1 νA 1
a11 = , a22 = , a12 = − , a66 = , a = a26 = 0 (72)
ET EA EA μA 16
212 N. McCartney

For conditions of plane strain such that εzz ≡ 0,




1 − νT2 1 2 ET νA (1 + νT )
a11 = , a22 = 1 − νA , a12 = − ,
ET EA EA EA (73)
1
a66 = , a = a26 = 0
μA 16
For a 0o -90o -0o cross-ply laminate the values of EA and ET (the axial and
transverse Young’s moduli), νA (the axial Poisson’s ratio) and μA (the axial
shear modulus) are calculated from single ply properties. It should be noted
that thermal expansion effects have been neglected in (71).

3.2 A Representation for Stress and Displacement Fields


Consider a straight through-crack of length 2c embedded within an in-
finite transverse isotropic plate so that the crack plane is normal to the
surfaces of the plate. For conditions of generalised plane stress and plane
strain, crack problems are two dimensional. Introduce a set of rectangular
Cartesian coordinates (x, y) such that the crack is parallel to the x -axis and
occupies the region |x − a| ≤ c, y = b. The complex variable z is introduced
such that
z = x + iy (74)
The locations of the two crack tips are specified by
z = t1 = a − c + ib, (75)
z = t2 = a + c + ib (76)
The stress and displacement representation developed by Sih and Liebowitz
(1968) is used, which is valid for plane strain deformations in rectilinearly
anisotropic bodies containing a crack. In the absence of body forces, as is
assumed in the present model, this representation may also be applied to
the conditions of generalised plane stress that are being considered.
The representation involves two analytic functions φ (z1 ) and ψ (z2 ) of
the complex variables
z1 = x − a + s1 (y − b), z2 = x − a + s2 (y − b), (77)
where s 1 and s 2 are the two roots having positive imaginary parts (see
below) of the following quadratic equation (since a 16 = a 26 = 0)
a11 s4 + (2a12 + a66 )s2 + a22 = 0 (78)
The representation for the stress components is given by
 
σxx =  s21 φ (z1 ) + s22 ψ  (z2 ) , (79)
Analytical Methods of Predicting Performance… 213

σyy =  [φ (z1 ) + ψ  (z2 )] , (80)


σxy =  [s1 φ (z1 ) + s2 ψ  (z2 )] , (81)
and that for the displacement components is

u =  [p1 φ(z1 ) + p2 ψ(z2 )] , (82)

v =  [q1 φ(z1 ) + q2 ψ(z2 )] , (83)


where (since a 16 = a 26 = 0)
0
pk = a11 s2k + a12
for k = 1, 2 (84)
qk = a12 sk + a22 /sk

Putting
a22 2a12 + a66
γ= , δ= , (85)
a11 2a11
it follows from (78) that

s2 = −δ ± δ2 − γ (86)

It can be shown that for both GRP and CFRP, the values of s 2 are real
and negative. The following distinct pure imaginary roots are, therefore,
obtained  
 
s1 = i δ − δ 2 − γ, s2 = i δ + δ 2 − γ (87)
The other two roots of (78) are given by s̄1 = −s1 , s̄2 = −s2 . The stress and
displacement representation automatically satisfies the equilibrium equa-
tions and the stress-strain relations (71) for any analytic functions φ(z) and
ψ(z) of the complex variable z. They are now assumed to take the following
form:
)t2
1 1
φ(z) ≡ w(t)ρ̂(t) ln dt, (88)
2πi z−t
t1

)t2
1 1
ψ(z) ≡ w(t)σ̂(t) ln dt (89)
2πi z−t
t1

The density functions ρ̂(t) and σ̂(t) are assumed to be polynomials and
t 2 − t1
w(t) ≡ 1/2
(90)
[ (t − t1 )(t − t2 ) ]
214 N. McCartney

In order that the displacement components u and v are single-valued func-


tions at all points in the complex plane lying outside the crack, the following
crack tip closure conditions must be satisfied:

)t2 )t2
w(t)ρ̂(t)dt = 0, w(t)σ̂(t)dt = 0 (91)
t1 t1

Differentiating (88) and (89) leads to

)t2
 1 w(t)ρ̂(t)
φ (z) ≡ dt, (92)
2πi t−z
t1

)t2
1 w(t)σ̂(t)
ψ  (z) ≡ dt (93)
2πi t−z
t1

Provided the conditions (91) are satisfied

φ (z) = O(z −2 ), ψ  (z) = O(z −2 ) as |z| → ∞, (94)

indicating that the stress field arising from the representation (79)-(81) has
zero net force applied at infinity.
The algebra may be simplified by changing variables to

ζ = ξ + iη, (95)

where
2z − (t1 + t2 )
ζ = ζ(z) = , (96)
t 2 − t1
x−a y−b
ξ= , η= (97)
c c
The crack is then described by −1 < ξ < 1, η = 0 and it follows from (92)
and (93) that
)1
 1 ρ(s) ds
φ (z) ≡ √ , (98)
π 1−s 2 ζ(z) −s
−1

)1
 1 σ(s) ds
ψ (z) ≡ √ , (99)
π 1 − s ζ(z) − s
2
−1
Analytical Methods of Predicting Performance… 215

where ρ(ξ) ≡ ρ̂(t), σ(ξ) ≡ σ̂(t). From (91) the crack closure conditions are
written
)1 )1
ρ(s)ds σ(s)ds
√ = 0, √ =0 (100)
1−s 2 1 − s2
−1 −1

3.3 Chebyshev Polynomial Expansion


Chebyshev polynomials of the first kind are defined over the interval
[−1, 1] by
Tn (cos α) ≡ cos nα, 0 ≤ α ≤ π, n ≥ 0 (101)
Chebyshev polynomials of the second kind are defined by
sin(n + 1)α
Un (cos α) = , 0 ≤ α ≤ π, n ≥ 0 (102)
sin α
The functions Tn (z ) and Un (z ) can both be analytically continued to the
entire complex plane, simply by considering the usual analytic continuation
of the functions sin(z) and cos(z). Use will be made of the following two
identities:
Tn ((λ + 1/λ)/2) ≡ (λn + 1/λn )/2, n ≥ 0, (103)
Un−1 ((λ + 1/λ)/2) ≡ (λn − 1/λn ) / (λ − 1/λ) , n ≥ 1 (104)
The density functions ρ(ξ) and σ(ξ) are now assumed to be of the form

N 
N
ρ(ξ) ≡ An Tn (ξ), σ(ξ) ≡ Bn Tn (ξ), (105)
n=1 n=1

where An and Bn are complex coefficients. The crack closure conditions


(100) are automatically satisfied and substitution of (105) in (98) and (99)
leads to

N N
φ (z) ≡ An Hn (ζ), ψ  (z) ≡ Bn Hn (ζ), (106)
n=1 n=1
where, using a result established by Gladwell and England (1977),
 n
)1 ζ − (ζ 2 −1 )
1/2
1 Tn (s) ds
Hn (ζ) ≡ √ ≡  2 1/2 ,n ≥ 0 (107)
π 1 − s2 ζ − s ζ −1
−1

By selecting the branch of (ζ 2 −1 )1/2 which is asymptotic to ζ as |ζ| → ∞,


it can be shown that for n ≥ 1,
 n
ζ − (ζ 2 −1 )1/2 Tn (ζ)
Hn (ζ) ≡  2 1/2 ≡  1/2 − Un−1 (ζ) (108)
ζ −1 ζ −1
2
216 N. McCartney

3.4 Traction Distribution on the Crack


Let Sn+ , St+ , Sn− , St− be the normal and transverse tractions on the upper
and lower surfaces of the crack. Since
⎧  2

⎪ ξ −1, ξ > 1, η = 0,

⎪ 

 2 1/2 ⎨ i 1 − ξ 2 , |ξ| < 1, η = 0+,
ζ −1 =  (109)

⎪ −i 1 − ξ 2 , |ξ| < 1, η = 0−,



⎩  2
− ξ −1, ξ < −1, η = 0,
it follows from (108) that
Tn (ξ)
Hn± (ξ) = ±i  − Un−1 (ξ), |ξ| < 1, n ≥ 1, (110)
1 − ξ2
where Hn+ and Hn− are the limiting values of Hn (ζ) on the positive and
negative sides of the crack respectively. Substituting (106) into (80) and
(81) using (110) leads to the following relations valid only for |ξ| < 1:
 
N
Tn (ξ)
±
Sn (ξ) = ± [An + Bn ]  −  [An + Bn ] Un−1 (ξ) , (111)
n=1 1 − ξ2
 
N
Tn (ξ)
±
St (ξ) = ± [s1 An + s2 Bn ]  +  [s1 An + s2 Bn ] Un−1 (ξ)
n=1 1 − ξ2
(112)
The tractions Sn+ , St+ , Sn− , St− must be bounded as ξ → ±1, for values of
ξ in the range |ξ| < 1 leading to the following conditions for the complex
coefficients An , Bn :
 [An + Bn ] =  [s1 An + s2 Bn ] = 0 (113)
These conditions may easily be satisfied by choosing real αn , βn such that
αn = An + Bn , βn = −(s1 An + s2 Bn ), (114)
and
αn s2 + βn αn s1 + βn
An = − , Bn = , (115)
s1 − s2 s1 − s2
which, furthermore, yields the following simple expression for the tractions
on the crack surfaces:
 ±  
N
±
Sn + iSt (ξ) = − (αn + i βn ) Un−1 (ξ), |ξ| < 1 (116)
n=1
Analytical Methods of Predicting Performance… 217

3.5 Stress and Displacement Fields Around the Crack


Define  n
Gn (ζ) ≡ ζ − (ζ 2 − 1)1/2 , n ≥ 1, (117)

noting that numerical calculation is convenient by making use of the fact


that
Gn (ζ) ≡ exp [−n(α + iβ)] , when ζ = cosh(α + iβ) (118)
It is easily shown that

d
Gn (ζ) = −nHn (ζ), n ≥ 1 (119)

and it then follows from (106) that

N
An
φ(z) = −c Gn (ζ), (120)
n=1
n

N
Bn
ψ(z) = −c Gn (ζ) (121)
n=1
n

Let
ζ1 = ξ + s1 η and ζ2 = ξ + s2 η (122)
On substituting (106), (115), (120) and (121) into the representation (79)-
(83), the stresses and displacement components may be expressed


N
 
σxx = −  (αn s1 s2 + βn (s1 + s2 )) Hn (ζ1 ) + s22 (αn s1 + βn )ΔHn (ζ1 , ζ2 ) ,
n=1
(123)

N
σyy =  [αn Hn (ζ1 ) − (αn s1 + βn )ΔHn (ζ1 , ζ2 )] , (124)
n=1


N
σxy =  [βn Hn (ζ1 ) + s2 (αn s1 + βn )ΔHn (ζ1 , ζ2 )] , (125)
n=1

N
1
u = c  [(αn (a11 s1 s2 − a12 ) + βn a11 (s1 + s2 )) Gn (ζ1 )
n=1
n (126)

+ (a11 s22 + a12 )(αn s1 + βn )ΔGn (ζ1 , ζ2 ) ,
218 N. McCartney

N 



1 s1 + s2 a22
v = c  −αn a22 + βn a12 − Gn (ζ1 )
n=1
n s1 s2 s1 s2
  
+ a12 s2 + as22
2
(αn s 1 + β n ) ΔGn (ζ ,
1 2ζ ) ,
(127)
where
⎧  

⎪ Hn (ζ1 ) − Hn (ζ2 )

⎪ , for s1 = s2 ,

⎪ s1 − s2 

⎨ Hn (ζ1 ) − Hn (ζ2 )
ΔHn (ζ1 , ζ2 ) ≡ lim
⎪ s1 →s2
 s1 − s2 



⎪ n ζ

⎩ = −η
1
⎪ + 2 H(ζ1 ), for s1 = s2
(ζ12 −1 )
1/2
ζ1 −1
(128)
and
⎧  

⎪ Gn (ζ1 ) − Gn (ζ2 )
⎨ , for s1 = s2 ,
ΔGn (ζ1 , ζ2 ) ≡ s1 − s2 

⎪ Gn (ζ1 ) − Gn (ζ2 )
⎩ lim = −nηH(ζ1 ), for s1 = s2 ,
s1 →s2 s1 − s2
(129)
One limiting situation, s1 = s2 = i, occurs when the material is isotropic,
and in this case the expressions (123)-(127) coincide with those of McCart-
ney and Gorley (1987) for the case of parallel cracks. It should be noted
that
ΔGn (ζ1 , ζ2 ) → 0 , ΔHn (ζ1 , ζ2 ) → 0 as y → 0

3.6 Displacement Discontinuity Across the Crack


From (108) and (117)
 1/2  1/2
Gn (ζ) ≡ Hn (ζ) ζ 2 − 1 ≡ Tn (ξ) − ζ 2 − 1 Un−1 (ζ) (130)

It is deduced from (109) that the limiting values of Gn (ζ) on the crack faces
are given by

G±n (ξ) = Tn (ξ) ± i 1 − ξ Un−1 (ξ)
2 (131)

By considering the limiting distributions for the normal and tangential dis-
placements along the upper and lower surfaces of the crack, which are de-
noted by Vn+ , Vn− , Ut+ , Ut− , use can be made of (126) and (127) together with
(131) to obtain an expression for the displacement discontinuities across the
Analytical Methods of Predicting Performance… 219

crack:

√   αn
N
Δv(ξ) ≡ (Vn+ − Vn− )(ξ) = 4c a22 g 1 − ξ 2 Un−1 (ξ), (132)
n=1
n

 N
√ βn
Δu(ξ) ≡ (Ut+ − Ut− )(ξ) = 4c a11 g 1 − ξ2 Un−1 (ξ), (133)
n=1
n
where
1 √
g= (2 a11 a22 + 2a12 + a66 ) (134)
4
In the isotropic limit, this result agrees with the corresponding result of
McCartney and Gorley (1987).

3.7 Stress Intensity Factors


For the crack tip at t1 , the mode I and mode II stress intensity factors
KI1 , KII
1
are defined by
 
KI1 + iKII
1
= lim 2πc(−ξ − 1) (Sn (ξ) + iSt (ξ)) , (135)
ξ→−1

where Sn (ξ) and St (ξ) are the normal and tangential tractions acting on
ξ < −1, η = 0. It follows from (80), (81), (106), (108) and (115) that on
ξ < −1, η = 0 the tractions are given by
 
N
Tn (ξ)
Sn (ξ) + iSt (ξ) = − (αn + i βn )  2 + Un−1 (ξ) (136)
n=1 ξ −1

Substituting into (135) leads immediately to the simple result

√  N
KI1 + iKII
1
= πc (−1 )n+1 (αn + i βn ) , (137)
n=1

since Tn (−1) = (−1 )n . Similarly for the crack tip at t2 ,


 
KI2 + iKII
2
= lim 2πc(ξ − 1) (Sn (ξ) + iSt (ξ)) , (138)
ξ→1

where Sn (ξ) and St (ξ) are the normal and tangential tractions acting on
ξ > 1, η = 0. For ξ > 1, η = 0,
 
N
Tn (ξ)
Sn (ξ) + iSt (ξ) = (αn + i βn )  2 − Un−1 (ξ) (139)
n=1 ξ −1
220 N. McCartney

Since Tn (1) = 1, substitution into (138) yields the following corresponding


simple result
√  N
KI2 + iKII2
= πc (αn + i βn ) (140)
n=1

3.8 Example Prediction


Consider now a test example of two collinear cracks of equal length em-
bedded in an isotropic material, where the exact solution is known and can
be compared with the predictions obtained using orthogonal polynomials.
Figure 4 shows the geometry where the two cracks have length 2c and the
separation of the inner crack tips is 2s. A uniaxial stress σ is applied in a
direction normal to the crack planes. The crack problem is such that the
deformation at the crack tips is mode I, i.e. the mode II stress intensity
factors are zero.
The exact solution for the mode I stress intensity factors is given by
Rooke and Cartwright (1976)
  
√ s + 2c 1 E(k)
KI (x = s) = σ πc − α2 ,
αc k K(k)
   (141)
√ s + 2c 1 E(k)
KI (x = s + 2c) = σ πc 1−
c k K(k)

with k = 1 − α2 and α = c/(s + 2c). For a unit applied stress and when
c = 0.45 c0 , s = 0.1c0 and on selecting N = 100, for any normalising crack
length c0 , the methodology based on orthogonal polynomials leads to the
results

KI (x = s) = 1.4923379, KI (x = s + 2c) = 1.2916506

The corresponding values obtained from the exact solution (141) is

KI (x = s) = 1.4923379, KI (x = s + 2c) = 1.2916506,

which are identical to the numerical estimates, thus confirming the validity
of the methodology based on orthogonal polynomials.
It should be noted that the relation (139) can be used to investigate
magnitude of the tractions on the crack surfaces. For the example considered
the tractions, which should be zero, have the order of 10−16 indicating the
very high accuracy of the methodology used.
Consider now a test example of three equally spaced vertically stacked
cracks of equal length embedded in an infinite isotropic material. Figure
Analytical Methods of Predicting Performance… 221

2c x
2c

2s

Figure 4. Diagram illustrating geometry and loading of part of an infinite


isotropic plate having two collinear cracks

5 illustrates the geometry where the three cracks have length 2c and the
vertical separation of the cracks is s. A uniaxial stress σ is applied in a
direction normal to the crack planes. The crack problem is such that the
deformation at the tips of the central crack is mode I, and the deformation
at the other tips is mixed mode. The magnitudes of the model I and mode
II stress intensity factors for the upper and lower cracks are expected to be
the same.
For a unit applied stress and when c = 0.5 c0 , s = 0.5c0 and on selecting
N = 100, for any normalising crack length c0 , the methodology based on
222 N. McCartney

x
s 2c

Figure 5. Diagram illustrating geometry and loading of part of an infinite


isotropic plate having three equally spaced vertically stacked cracks

orthogonal polynomials leads to the results

Top crack : KI = 0.92924447, KII = 0.14265165,


Central crack : KI = 0.70031026, KII = 0.0,
Bottom crack : KI = 0.92924447, KII = −0.14265165

It is seen that the mode I stress intensity factor for the central crack is less
than that of the upper and lower cracks. This illustrates the shielding effect
on the central crack because of the presence of the other two cracks. The
model II stress intensity of the lowest crack is seen to be negative because the
local shear stress is negative. For this example the tractions, which should
be zero, have the order of 10−16 again indicating the very high accuracy of
Analytical Methods of Predicting Performance… 223

the methodology used.

4 Generalised Plane Strain Theory for Cross-Ply


Laminates
Stress transfer phenomena in cross-ply laminates have been examined ex-
tensively in the literature using an approximate method of stress analysis,
first developed for UD composites, that is known as ‘shear-lag theory’. An
improved approach was developed by the author many years ago (McCart-
ney, 1992) by removing various approximations that had to be made when
developing shear-lag solutions. The objective of this Section is to show how
the methodology described in reference (McCartney, 1992) may be further
extended.
Using a simple cross-ply laminate as an example, alternative methods
will be described for determining the stress and displacement distributions
when there is a regularly spaced array of ply cracks in the 90o ply. It will
be shown how the effective elastic and thermal constants can be estimated.
The example considered here, given in detail for the first time, is a very
good way of presenting the important physical principles that have already
been applied to the more complex case of general symmetric laminates. The
approach to be described and developed in detail here has been extended in
previous work by the author recently summarised in one of the publications
(McCartney, 2013a) of an International Exercise (Kaddour et al., 2013a).
Consider the model of a simple [0/90]s cross-ply laminate illustrated in
Fig. 6 where two possible representative volume elements (RVEs) are shown.
For the first shown in Fig. 6(a) one ply crack in the inner 90o ply is located
on the plane x1 = 0 and neighbouring ply cracks (not shown) are on the
planes x1 = ±2L. For the second RVE shown in Fig. 6(b), the plane x1 = 0
is mid-way between two neighbouring ply cracks in the inner 90o ply on the
planes x1 = ±L. The inner 90o ply has total thickness 2a and the two outer
0o plies each have thickness denoted by b so that the total thickness of the
laminate is 2h where h = a + b. A set of Cartesian coordinates (x1 , x2 , x3 )
is introduced such that the origin lies on the mid-plane of the laminate at
the mid-point between two neighbouring cracks in the 90o ply. The x1 -axis
is directed along the principal loading direction and the x3 -axis is directed
in the through-thickness direction.
The following equilibrium equations must be satisfied for both the 0o
224 N. McCartney

2h 2h

x1 x1

x3 x3
2L
0 0

b b b b
2a 2a

(a) (b)
Figure 6. Representative volume elements for a cracked cross-ply laminate

and 90o plies

∂ σ11 ∂ σ12 ∂ σ13


+ + = 0,
∂ x1 ∂ x2 ∂ x3
∂ σ12 ∂ σ22 ∂ σ23
+ + = 0, (142)
∂ x1 ∂ x2 ∂ x3
∂ σ13 ∂ σ23 ∂ σ33
+ + = 0,
∂ x1 ∂ x2 ∂ x3

where σij are the stress components. The plies are regarded as transverse
isotropic solids so that the stress-strain-temperature relations involve the
axial and transverse values of the Young’s modulus E, Poisson’s ratio ν,
shear modulus μ and thermal expansion coefficient α. Superscripts ‘0’ or
‘90’ will be used to denote the ply to which a stress, strain and displacement
Analytical Methods of Predicting Performance… 225

component refers. For the 0o plies


∂ u01 1 0 νA0
νa0 0
ε011 = = σ
0 11 − σ
0 22
0
− 0
0 σ33 + αA ΔT,
∂ x1 EA EA EA
∂ u02 ν0 0 1 0 ν0 0
ε022 = = − A0 σ11 + 0 σ22 − t0 σ33 + α0T ΔT,
∂ x2 EA ET ET
∂ u03 ν0 0 ν0 0 1 0
ε033 = = − a0 σ11 − t0 σ22 + 0 σ33 + α0t ΔT,
∂ x3 EA ET Et
(143)
∂ u01 ∂ u02 σ0
2ε012 = + = 12 ,
∂ x2 ∂ x1 μ0A
∂ u01 ∂ u03 σ0
2ε013 = + = 13 ,
∂ x3 ∂ x1 μ0a
∂ u2 ∂ u03 σ0
2ε023 = + = 23 ,
∂ x3 ∂ x2 μ0t
while for the 90o plies
∂ u90 1 90 ν 90 90 νt90 90
ε90
11 =
1
= σ11 − A90 σ22 − σ + α90 T ΔT,
∂ x1 ET90 EA ET90 33
∂ u90 ν 90 90 1 90 νa90 90
ε90
22 =
2
= − A90 σ11 + 90 σ22 − 90
90 σ33 + αA ΔT,
∂ x2 EA EA EA
∂ u90 ν 90 90 ν 90 90 1 90
ε90
33 =
3
= − t90 σ11 − a90 σ22 + σ + α90t ΔT,
∂ x3 ET EA Et90 33
(144)
∂ u90
1 ∂ u90
2 σ 90
2ε90
12 = + = 12 ,
∂ x2 ∂ x1 μ90
A
∂ u90
1 ∂ u90
3 σ 90
2ε90
13 = + = 13 ,
∂ x3 ∂ x1 μ90
t
∂ u90
2 ∂ u90
3 σ 90
2ε90
23 = + = 23 ,
∂ x3 ∂ x2 μ90
a

where the strain and displacement components are denoted by εij and
ui respectively. The subscripts A, T and t are attached to the proper-
ties to associate them respectively with the axial, in-plane transverse and
through-thickness directions of the lamina. It should be noted that the up-
per case subscripts A and T are associated only with in-plane directions,
while the lower case subscripts are associated with the through-thickness
direction. The relations (144) are either obtained by modifying directly the
relations (143) for the 0o plies, or by using the relations (27)-(40) to ro-
tate the ply by an angle ± 90o . The thermoelastic constants of individual
226 N. McCartney

plies in a laminate are usually assumed to be transverse isotropic so that


Et = ET , νa = νA , μa = μA , αt = αT and ET = 2μt (1 + νt ).

4.1 Free Surface, Interface, Edge and Symmetry Conditions


In order that the field equations can be solved uniquely, it is necessary
to impose a sufficient number of boundary and interface conditions. The
free surface (x3 = ±h) and interface (x3 = ±a) conditions will first be
considered. On the free surfaces
0
σ33 0
= σt , σ13 0
= σ23 = 0, on x3 = ±h, (145)
and on the interfaces
0 90 0 90 0 90
σ33 = σ33 , σ13 = σ13 , σ23 = σ23 ,
on x3 = ±a (146)
u01 = u90
1 , u02 = u90
2 , u03 = u90
3 ,

The edges x2 = ±W are such that in-plane transverse displacement is uni-


form having the following values

2 = ±W εT ,
u02 = u90 on x2 = ±W, (147)
where εT is the in-plane transverse strain that is uniform everywhere in the
laminate when generalised plane strain conditions are imposed. The edges
x2 = ±W are assumed to have zero shear stresses so that
0
σ12 90
= σ12 0
= 0, σ23 90
= σ23 = 0, on x2 = ±W (148)
For the above boundary conditions, and because of the symmetric nature of
the laminate, there will be symmetry about x3 = 0 of the stress, strain and
displacement distributions such that the following conditions are satisfied
0 90 0 90
σ13 = σ13 = 0, σ23 = σ23 = 0, u03 = 0, onx3 = 0 (149)
When applying laminate edge conditions applied on planes normal to the
x1 -axis, two possible approaches can be made. Consider first of all the RVE
shown in Fig. 6(a) which can be used for undamaged laminates, and for
damaged laminates where a ply crack in the 90o ply is located at x1 = 0.
The edges x1 = ±L are such that in-plane axial displacement is uniform
having the following values

1 = ±LεA ,
u01 = u90 on x1 = ±L, (150)
where εA is the effective axial applied strain. The edges x1 = ±L are
assumed to have zero shear stresses so that
0
σ12 90
= σ12 = 0, 0
σ13 90
= σ13 = 0, on x1 = ±L (151)
Analytical Methods of Predicting Performance… 227

These conditions imply that there is symmetry about the plane x1 = 0 so


that
0 90 0 90
σ12 = σ12 = 0, σ13 = σ13 = 0, u01 = u90
1 = 0, on x1 = 0 (152)
Consider now the RVE shown in Fig. 6(b) where the 90o ply cracks are
located on the planes x1 = ±L. The boundary conditions applied are given
by
0
σ12 90
= σ12 0
= 0, σ13 90
= σ13 90
= 0, σ11 = 0, u01 = ±LεA , on x1 = ±L (153)
It is clear from both the RVEs shown in Fig. 6 and the boundary conditions
applied on planes normal to the x1 -axis that for uniformly arrays of ply
cracks there is symmetry about the planes x1 = 0 and x1 = ±L.

4.2 Key Results for Undamaged Laminates


For undamaged laminates and in regions away from laminate edges, the
axial and transverse strains in both 0o and 90o plies have the uniform values
ε̂A and ε̂T respectively. The ‘hat’ symbol is attached to the in-plane strains
to distinguish them from the differing values εA and εT that will arise when
the laminate is damaged. The through-thickness stress in both plies has
the uniform value σt . For undamaged laminates it is useful to define the
following laminate constants for the 0o and 90o plies

0

0

1 1 0 2 ET 1 1 0 2 ET
= 0 1 − (ν A ) 0 , = 1 − (ν A ) ,
ẼA 0 EA EA ẼT0 ET0 EA0

ν̃a0 νa0 + νt0 νA0


ν̃t0 νt0 νa0 νA0
= 0 , = 0 + 0 , (154)
ẼA 0 EA ẼT0 ET EA
0
0 ET 0
α̃0A = α0A + νA α , α̃0 = α0T + νA 0 0
αA ,
EA T T
0


90

90

1 1 90 2 ET 1 1 90 2 ET
= 90 1 − (νA ) 90 , = 90 1 − (νA ) 90 ,
90
ẼA EA EA ẼT90 ET EA
90 90 90 90 90 90 90 90
ν̃a ν ν ν ν̃t ν ν ν
= a90 + t 90A , = t90 + a 90A , (155)
90
ẼA E A E A ẼT90 E T EA
90
90 90 ET 90
α̃90
A = αA + νA 90 αT , α̃90 90
T = αT + νA αA
90 90
EA
From a consideration of mechanical equilibrium the uniform ply stresses can
be used to define, for an undamaged laminate, the effective axial stress σ̂A
and the effective in-plane transverse stress σ̂T as follows
0 90
hσ̂A = bσ̂11 + aσ̂11 , (156)
228 N. McCartney
0 90
h σ̂T = bσ̂22 + aσ̂22 (157)
The ‘hat’ symbol is used to distinguish these effective stresses from those
that will result when the laminate is damaged. Corresponding to the uni-
form through-thickness stress σt , an effective through-thickness strain ε̂t
can be defined by the relation
hε̂t = bε̂033 + aε̂90
33 (158)
It should be noted that the value σt for the through-thickness stress of an
undamaged laminate corresponds to the effective value when the laminate
is damaged. It can be shown that on defining the constants
b 0 a b 0 0 a 90 90 b a
A = ẼA + ẼT90 , B = νA ẼT + νA ẼT , C = ν̃a0 + ν̃t90 ,
h h h h h h
b 0 a 90 b 0 a 90
F = ẼT + ẼA , G = ν̃t + ν̃a , (159)
h h h h
b 0 0 a b 0 0 a 90 90
P = ẼA α̃A + ẼT90 α̃90
T ,Q = Ẽ α̃ + ẼA α̃A ,
h h h T T h
the thermoelastic constants of an undamaged cross-ply laminate are given
by
(L) B2 (L) B2 (L) B
EA = A − , ET = F − , νA = ,
F A F
(L) BG (L) BC
νa = C − , νt = G − , (160)
F A
 
1   (L)
(L) (L) (L) 1 (L) ET
αA = (L) P − νA Q , αT = (L) Q − νA (L)
P ,
EA ET EA
such that the in-plane non-shear stress-strain relations for an undamaged
laminate are
(L) (L)
1 νA νa (L)
ε̂A = σ̂ −
(L) A
σ̂ −
(L) T
σ + αA ΔT,
(L) t
EA EA EA
(L) (L) (161)
ν 1 ν (L)
ε̂T = − A(L) σ̂A + (L) σ̂T − t(L) σt + αT ΔT
EA ET ET
It should be noted that
(L)
(L) b 0 a 90 (L) ET B b ν 0 Ẽ 0 + aνA90 90
ẼT
ẼA = A = Ẽ + Ẽ , νA == A T0 ,
h A h T (L)
EA A bẼA + a ẼT90
(L) b a 90 (L) B bν 0 Ẽ 0 + aνA 90 90
ẼT
ẼT = F = ẼT0 + Ẽ , νA = = A T0 ,
h h A F bẼT + aẼA 90

(162)
Analytical Methods of Predicting Performance… 229

where
(L) (L)
(L) EA (L) ET
ẼA = (L) (L) (L)
, ẼT = (L) (L) (L)
(163)
1 − (νA )2 ET /EA 1 − (νA )2 ET /EA

To calculate the through-thickness properties the following constants are


first defined


 b νa0 + νt0 νA0
0 a νt90 νa90 νA
90
b a
A = 0 ẼA + + ẼT90 = ν̃a0 + ν̃t90 ,
h EA h ET90 EA 90 h h


 b νt0 νa0 νA0
0 a νa90 + νt90 νA90
90 b a
B = + Ẽ T + ẼA = ν̃t0 + ν̃a90 ,
h ET0 EA 0 h EA 90 h h



 b 1 νa0 ν̃a0 νt0 ν̃t0 a 1 νt90 ν̃t90 νa90 ν̃a90
C = − 0 − 0 + − − ,
h Et0 EA ET h Et90 ET90 EA 90


b ν0 0 0 ν0
P = α0t + a0 ẼA α̃A + t0 ẼT0 α̃0T
h EA ET

90

a 90 ν a 90 90 νt90 90 90
+ αt + 90 ẼA α̃A + 90 ẼT α̃T
h EA ET
(164)
It can be shown that the through-thickness properties are given by
(L)
(L) ET
= A − B  νA , = B  − A νA
(L) (L) (L)
νa νt (L)
,
EA
(L) (L) (165)
1 νa νt
= C  + A + B = P  − A αA − B  αT ,
(L) (L) (L)
(L) (L) (L)
, αt
Et EA ET

such that
(L) (L)
νa νt 1 (L)
ε̂t = − σ̂ −
(L) A (L)
σ̂T + (L)
σt + αt ΔT, (166)
EA ET Et
It can be shown that the same values result for the minor Poisson’s ratios
(L) (L)
νa and νt when using the relations (160) or (165).

4.3 Effective Applied Stresses and Strains


Consider a symmetric damaged laminate of length 2L, width 2W and
total thickness 2h, for which the plies are uniformly thick and perfectly
bonded together. In-plane loading is applied by imposing uniform axial and
transverse displacements, denoted by ±UA and ±UT , on the edges of the
laminate thus defining an effective axial strain εA = UA /L and an effective
transverse strain εT = UT /W . The faces of the laminate are subjected to
230 N. McCartney

a uniform applied stress denoted by σt . Corresponding to the effective in-


plane strains εA and εT , axial and transverse effective applied stresses can
be defined for damaged laminates by
)W )h )L )h
1 1
σA = σ11 dx2 dx3 , σT = σ22 dx1 dx3 (167)
4W h 4Lh
−W −h −L −h

Corresponding to the applied through-thickness stress σt an effective through-


thickness strain εt can be defined for damaged and undamaged laminates
by
)W )L
1
εt = [u3 (x1 , x2 , h) − u3 (x1 , x2 , −h)] dx1 dx2 (168)
8LW h
−W −L

For an undamaged laminate subject to these displacements, the axial and


transverse strains will be uniform throughout all plies of the laminate having
the values εA and εT respectively, and the through-thickness stress will be
uniform throughout the laminate having the value σt in each ply. It then
follows from (167) and (168) that σA = σ̂A , σT = σ̂T and εt = ε̂t where σ̂A ,
σ̂T and ε̂t are defined by the relations (156)-(158).

4.4 Generalised Plane Strain Solution


Generalised plane strain conditions are assumed so that the stress and
strain distributions do not depend on the x 3 -coordinate. This situation
occurs when the displacement field is of the form

u1 = u1 (x1 , x3 ), u2 = εT x2 , u3 = u3 (x1 , x3 ), (169)

where ε22 = εT is the uniform in-plane transverse strain in both of the 0o


and 90o plies. In addition, it is assumed that on the laminate faces x3 = ±h
that σ330
≡ σ33
90
≡ σt where σt is a uniform applied through-thickness stress
and σ13 ≡ σ23 ≡ 0. If εT = 0 then the laminate is highly constrained in
the transverse direction leading to well known conditions of plane strain
deformation. It should be noted that the solution for undamaged laminates
derived in the previous section is automatically one of generalised plane
strain as the transverse strain εT is uniform everywhere in those regions of
the laminate that are sufficiently well away from the edges.
It can be shown that for the 0o plies
0 0
νA ET 0
0 σ11 + νt σ33 − ET αT ΔT + ET εT ,
0 0 0 0 0 0
σ22 = (170)
EA
Analytical Methods of Predicting Performance… 231

1 0 ν̃a0 0 0
0 ET
ε011 = σ
0 11
− σ
0 33
+ α̃0
A ΔT − ν A 0 εT , (171)
ẼA ẼA EA
ν̃a0 0 1 0
ε033 = − 0
σ11 + 0 σ33 + α̃0t ΔT − νt0 εT , (172)
ẼA Ẽt
where
1 1 (ν 0 )2
= 0 − t 0 , α̃0t = α0t + νt0 α0T (173)
0
Ẽt Et ET
For the 90o ply
90
σ22 90 90
= νA σ11 + νa90 σ33
90
− EA
90 90 90
αA ΔT + EA εT , (174)

1 90 ν̃ 90 90
ε90
11 = 90
σ11 − t90 σ33 T ΔT − νA εT ,
+ α̃90 90
(175)
ẼT ẼT
ν̃t90 90 1 90
33 = −
ε90 t ΔT − νa εT ,
+ α̃90 90
σ11 + 90 σ33 (176)
ẼT90 Ẽt
where
1 1 (νa90 )2
= 90 − 90 , α̃90 90 90 90
t = αt + νa αA (177)
Ẽt90 Et EA

4.5 Key Results for Damaged Laminates


The easiest approach is to consider the stress and displacement repre-
sentation involving an unknown stress transfer function C(x1 ) and to show
that it satisfies required field equations and boundary conditions. For gen-
eralised plane strain conditions where it is assumed σ12 = σ23 = 0 and
u2 = εT x2 , the stress distribution is assumed to have the following form
b
0
σ11 0
(x1 , x3 ) = C(x1 ) + σ̂11 , 90
σ11 (x1 , x3 ) = − C(x1 ) + σ̂11
90
, (178)
a
b 
0
σ13 (x1 , x3 ) = C  (x1 ) (h − x3 ) , 90
σ13 (x1 , x3 ) = C (x1 )x3 , (179)
a
0 0
νA ET 1 0  2
0 C(x1 ) + 2 νt C (x1 ) (h − x3 ) + σ̂22 ,
0 0
σ22 (x1 , x3 ) = (180)
EA
b 90 b  
90
σ22 (x1 , x3 ) = − νA C(x1 ) + νa90 C  (x1 ) ah − x23 + σ̂22
90
, (181)
a 2a
1  2
0
σ33 (x1 , x3 ) = C (x1 ) (h − x3 ) + σt ,
2 (182)
b   
90
σ33 (x1 , x3 ) = C (x1 ) ah − x23 + σt ,
2a
232 N. McCartney

and the corresponding displacement representation is


 
ν̃ 0
ν̃ 90
b 2
− (h − x )2
C  (x1 )
3
u01 (x1 , x3 ) = a
(x3 − a)2 − t90 b(x3 − a) +
2ẼA 0 ẼT 2μ0a


(h − x3 )4 − b4 + 4b3 (x3 − a) ab (2h + b) (x3 − a)
− + C  (x1 ) + A(x1 ),
24Ẽt0 6Ẽt90
  (183)
90
b 1 ν̃
u90
1 (x1 , x3 ) = − t90 (x23 − a2 )C  (x1 )
2a μ90 t ẼT
b 1  4 
+ x3 − a4 − 6ah(x23 − a2 ) C  (x1 ) + A(x1 )
24a Ẽt
90

  (184)
90 0
ν̃ ν̃
u03 (x1 , x3 ) = ε̂033 (x3 − a) + ε̂90 33 a + b 90 −
t a
0
(x3 − a) C(x1 )
ẼT ẼA
 
1 1  3  ab
+ b − (h − x 3 )3
+ (2h + b) C  (x1 ),
6 Ẽt0 Ẽt90
(185)
b ν̃t90 b 1   
90
u3 (x1 , x3 ) = C(x1 )x3 + C (x1 ) 3ah − x3 x3 + ε̂33 x3 (186)
2 90
a ẼT90 6a Ẽt90
It follows that
0 90 0 90
hσA = bσ11 (x1 , x3 ) + aσ11 (x1 , x3 ) = bσ̂11 + aσ̂11 = hσ̂A , (187)
so that the effective applied axial stress for damaged laminate σA is equal to
the effective axial stress σ̂A for the corresponding undamaged laminate. The
function A(x1 ) appearing in (183) and (184) is for the moment arbitrary.
The representation automatically satisfies the equilibrium equations (142)
and the required interface continuity conditions. In addition, all the stress-
strain relations (143) and (144), except for the two relations (143)1 and
(144)1 involving the axial strains ε011 and ε90 11 respectively, for any functions
C(x1 ) and A(x1 ). It is possible, however, to satisfy these axial relations after
they are averaged through the thickness of the 0o and 90o plies respectively,
as will now be described.
Assuming symmetry about the mid-plane x3 = 0, the average of any
quantities f0 (x1 , x3 ) and f90 (x1 , x3 ) associated with the 0o and 90o plies are
defined respectively by
)h )a
1 1
f¯0 (x1 ) = f0 (x1 , x3 ) dx3 , f¯90 (x1 ) = f90 (x1 , x3 ) dx3 (188)
b a
a 0
Analytical Methods of Predicting Performance… 233

On averaging (183) and (184) using (188) it can be shown that


 
ν̃a0 ν̃t90 1
0
ū1 (x1 ) = − + 0 b2 C  (x1 )
6ẼA 0 2ẼT90 3μa

(189)
1 a (2a + 3b) 4 
− + b C (x1 ) + A(x1 ),
20Ẽt0 12b2 Ẽt90
 
a ν̃t90 1
90
ū1 (x1 ) = − 90 b2 C  (x1 )
3b ẼT90 μt
(190)
a2 (5b + 4a) 1 4 
+ b C (x1 ) + A(x1 )
30b3 Ẽt90
On substituting into the averaged axial stress-strain relations for the 0o and
90o plies, and on eliminating the terms σ̂11
0 90
and σ̂11 , and the function A(x1 ),
it can be shown that the stress transfer function must satisfy the following
fourth order ordinary differential equation
F b4 C  (x1 ) − Gb2 C  (x1 ) + HC(x1 ) = 0, (191)
where

1 2 a a2 5 a 15
F = + + + > 0,
20Ẽt0 15Ẽt90 b b2 2b 8
 
1 1 1 a ν̃a0 ν̃t90 2a + 3b
G= + 90 + 0 − 90 , (192)
3 μ0a μt b ẼA ẼT b
1 1 b
H = 0 + 90 > 0
ẼA ẼT a
The function A(x1 ) can be calculated using either of the following two equiv-
alent relations


1 a(2a + 3b) 4 
A(x1 ) = + b C (x1 )
20Ẽt0 12b2 Ẽt90
  (193)
ν̃a0 ν̃t90 1 2  1
− − + b C (x1 ) + C̄(x1 ) + ε̂ x
A 1 ,
3ẼA 0 2ẼT90 3μ0a 0
ẼA

a2 (5b + 4a) 4 


A(x1 ) =− b C (x1 )
30b3 Ẽt90
 
a 4a + 3b ν̃t90 1 b 1
− − 90 b2 C  (x1 ) − C̄(x1 ) + ε̂A x1 ,
3b 2a ẼT90 μt a ẼT90
(194)
234 N. McCartney

where
)x1
C̄(x1 ) ≡ C(x)dx. (195)
0

When there is symmetry about x1 = 0, it follows that C  (x1 ) = 0 and


C  (x1 ) = 0. The integration constant has been selected so that A(x1 ) = 0.
On substituting (193) in (189)

1 ν̃ 0
ū01 (x1 ) = 0
C̄(x1 ) − a0 b2 C  (x1 ) + ε̂A x1 , (196)
ẼA 6ẼA

and on substituting (194) in (190)

b 1 2a + 3b ν̃t90 2 
1 (x1 ) = −
ū90 C̄(x1 ) − b C (x1 ) + ε̂A x1 (197)
a ẼT
90 6b ẼT90

It should be noted that ū01 (0) = ū90


1 (0) = 0 consistent with the conditions
(152)3 . Because the function A(x1 ) is now known in terms of C(x1 ), the
u1 displacement distributions (183) and (184) are fully specified in terms of
the stress transfer function C(x1 ).
On defining

G H
r= , s= , (198)
2F F
the most general solution of the differential equation (191) satisfying the
symmetry condition C(x1 ) ≡ C(−x1 ) is given by
• if s > r:
px1 qx1 px1 qx1
C(x1 ) = P cosh cos + Q sinh sin , (199)
b b b b
• if s = r:
px1 x1 px1
C(x1 ) = P cosh + Q sinh , (200)
b b b
• if s < r:
px1 qx1 px1 qx1
C(x1 ) = P cosh cosh + Q sinh sinh , (201)
b b b b
where  
1 1
p= (r + s), q= |r − s| (202)
2 2
Analytical Methods of Predicting Performance… 235

4.6 Solution for Ply Cracks


Consider now a uniform array of ply cracks, having density ρ = 1/(2L),
in the 90o ply of the cross-ply laminate as shown in Fig. 6(b). The tractions
on the ply crack surfaces must be zero so that from (179) and (178)

b 
90
σ13 (L, x3 ) = C (L)x3 = 0, implying C  (L) = 0, (203)
a

b a 90
90
σ11 (L, x3 ) = − C(L) + σ̂11
90
= 0, implying C(L) = σ̂ (204)
a b 11
On applying these conditions and on writing P = A + B, Q = A − B, the
parameters A and B must be selected so that
(p−q)L (p+q)L
a (p − q) tanh b a (p + q) tanh b
A=− 90
Λσ̂11 , B= 90
Λσ̂11 , (205)
b cosh (p+q)L
b
b cosh (p−q)L
b

where
1 (p + q)L (p − q)L
= (q + p) tanh + (q − p) tanh (206)
Λ b b
The only boundary condition for a damaged laminate that has not been
satisfied is given by (153)1 . It is clear from (183), (184) and (193) or (194)
that it is not possible for this boundary condition to be satisfied by the ap-
proximate solution derived. The boundary condition (153)1 is now replaced
by the following averaged condition

ū01 = ± LεA, on x1 = ±L (207)

It can then be shown using (196) and (203) that


90
a Φσ̂11
εA = 0
+ ε̂A (208)
LẼA

where
4Λpq (p + q)L (p − q)L
Φ= tanh tanh (209)
p −q
2 2 b b
On using (157) it can be shown using (180) and (181) that the effective
applied transverse stress σT defined by (167)2 is given by

0

ab 0 ET
σT = νA 0 − νA Φσ̂11
90 90
+ σ̂T (210)
Lh EA
236 N. McCartney

It is assumed that the values of σA = σ̂A , εT = ε̂T , σt and ΔT are known


so that the stress-strain relations (161) for an undamaged laminate may be
written
(L) (L)
1 νA νa (L)
ε̂A = σ −
(L) A
σ̂ −
(L) T (L) t
σ + αA ΔT ,
EA EA EA
(L) (L) (211)
νA 1 νt (L)
εT = − (L) σA + (L) σ̂T − (L) t
σ + αT ΔT
EA ET ET
On using (208) and (210) to eliminate ε̂A and σ̂T respectively, it can be
shown following extensive algebraic manipulation that
1 ν ν
εA = σA − A σT − a σt + αA ΔT, (212)
EA EA EA
νA 1 ν
εT = − σA + σT − t σt + αT ΔT, (213)
EA ET ET
where the thermoelastic constants of the damaged laminate are defined by

1 1  2 a Ẽ 90 Φ
(L) 90
= (L)
+ 1 − ν A ν A
T
, (214)
EA EA L E (L) Ẽ 0 ξ
A A

2 2 
0 0 90
1 1 0 ET b a ẼA ẼT Φ
= 1 + νA 0 − νA 90
, (215)
ET (L)
ET EA h2 L Ẽ (L) E (L) ξ
A T
(L)
implying ET = ξET ,

νA
(L)
νA  
E0
b a ẼT90 Φ
(L) 90
= (L)
+ 1 − ν A ν A ν 0
A
T
0 − ν 90
A (L) (L)
, (216)
EA EA EA hLE E ξ
A T

(L)
νa νa
= (L)
EA
EA  
(L)
ν 90
(L)
ν 90   (L) 90
νa νt (L) 90 a ẼA ẼT Φ
+ (L)
− t90 + νA
90
(L)
− a90 1 − νA νA ,
EA ET ET EA L E (L) ẼA
0 ξ
A
(217)
(L)
νt νt
= (L)
ET
ET  

(L) (L)
νa ν 90 νt ν 90 0
0 ET
90
90 b a ẼT Φ
− (L)
− t90 + νA
90
(L)
− a90 νA 0 − νA ,
EA ET ET EA EA h L E (L) ξ
T
(218)
Analytical Methods of Predicting Performance… 237

   a Ẽ (L) Ẽ 90 Φ
(L) (L) 90 (L) 90 (L)
αA = αA + 1 − νA νA αA + νA αT − α̃90
T
A T
, (219)
L E (L) ẼA0 ξ
A
 
E0
b a ẼT90 Φ
(L) (L) 90 (L)
αT = αT − αA + νA αT − α̃90T
0
νA T
0 − νA
90
, (220)
EA h L E (L) ξ
T
where
2 2
0 0 90
0 ET b a ẼA ẼT
ξ = 1 − νA 0 − ν 90
A 2
Φ. (221)
EA h L Ẽ E (L)
(L)
A T

The results (212)-(220) show that the stress-strain relations of a damaged


laminate are exactly of the same form as those for an undamaged lami-
nate. The formation of damage affects only the values of the thermoelastic
constants, and not the form of the stress-strain relations.

4.7 Through-Thickness Properties of Damaged Laminates


On applying (168) to the ply crack problem being considered
) L
1
εt = u03 (x1 , h)dx1 (222)
Lh 0

It can be shown again following extensive algebraic manipulation that through-


thickness stress-strain relation is obtained
νa ν 1
εt = − σA − t σT + σ + αt ΔT, (223)
EA ET Et t

where
νa νa
(L)   90
90 (L) b a ẼT Φ
= (L)
+ Ω 1 − ν ν
A A , (224)
EA EA h L E (L) ξ
A
(L)
0
2 0 90
νt νt 0 ET b a ẼA ẼT Φ
= (L)
− Ω ν A 0 − ν 90
A 2
, (225)
ET ET EA h L Ẽ E (L) ξ
(L)
A T
$  %
(L) (L)
1 1 νa νt90 νt νa90 b a 90 Φ
= (L) + Ω − 90 + νA 90
− 90 Ẽ , (226)
Et Et E
(L) ET E
(L) EA hL T ξ
A T
  ba Φ
(L) (L) (L)
αt = αt − Ω αA + νA
90
αT − α̃90
T Ẽ 90 , (227)
hL T ξ
and where

(L)
ν̃ 0 ν̃ 90 ν 0
0 ET
Ω = a0 − t90 − t(L) νA 0 − νA
90
(228)
ẼA ẼT ET EA
238 N. McCartney

The relations (217) and (218) are equivalent to the results (224) and (225)
because it can be shown that
 
(L) (L)
νa νt90 νt νa90 b ẼA 0

(L)
− 90 + νA90
(L)
− 90 = Ω (L)
(229)
E ET E EA h Ẽ
A T A

4.8 Example Predictions


The key results of this section are expressions for the various effective
thermoelastic properties of a simple cross-ply laminate having an array of
uniformly spaced ply cracks. For the example the following ply properties,
typical of a transverse isotropic carbon fibre reinforced composite, are used:

EA = 140.77 GPa, ET = 8.85 GPa, Et = 8.85 GPa,


νA = 0.28, νa = 0.28, νt = 0.43,
μA = 4.59 GPa, μa = 4.59 GPa, μt = 3.09441 GPa,
αA = 0.245x10−6K−1 , αT = 45.6x10−6K−1 , αt = 45.6x10−6K−1

When these ply properties are used in conjunction with the formulae
(214)-(220) and (224)-(227), for a set of ply crack densities in the range
0 – 4 cracks/mm, the results shown in Fig. 7 are obtained. The results
shown assume the following identifications:

EA ≡ EA, ET ≡ ET, Et ≡ Est,


νA ≡ nuA, νa ≡ nusa, νt ≡ nust,
μA ≡ muA, μa ≡ musa, μt ≡ must,
αA ≡ alA, αT ≡ alT, αt ≡ alst

It is noted that for ply crack densities exceeding 2/mm, the effective
properties no longer depend on the crack density. Also, it is seen that the
effective in-plane transverse modulus ET is hardly affected by ply cracking,
and that the effective axial thermal expansion coefficient is affected a great
deal by ply cracking.
Similar situations arise for ply cracking in the 90o plies of general sym-
metric laminates for a variety of laminate configurations considered in the
WWFE III International Exercise (Kaddour et al., 2013a) concerned with
the assessment of damage models for composite laminates. Results anal-
ogous to those derived here for cross-ply laminates have been derived for
general symmetric laminates (McCartney, 2013a,b) and assessed/discussed
by the organisers of the Exercise (Kaddour et al., 2013b). However, the
author recommends that a great deal of caution is applied when consider-
ing the comparison models as the information presented by the organisers
Analytical Methods of Predicting Performance… 239
1
Normalised properties

0.9
EA
ET
0.8
Est
nuA
0.7
nusa
nust
0.6 alA
alT
0.5 alst

0.4
0 0.5 1 1.5 2 2.5 3 3.5 4
Ply crack density (/[mm])
Figure 7. Predictions of the normalised effective properties of a simple
cross-ply laminate as a function of the density of a uniform distribution of
ply cracks in the 90o ply

(Kaddour et al., 2013b) relating to the generalised plane strain model devel-
oped by the author is wholly misleading, and conclusions are not justified
by the information presented to the exercise by the author (see McCartney
(2013b)).

5 Model of Composite Degradation Due to


Environmental Damage
The axial strength of unidirectional fibre reinforced composites is controlled
by the strength of the fibres. In cross-ply laminates the axial strength of the
laminate is controlled to a large degree by the strength of the fibres in the
0o plies. Fibre strength is statistical in nature due to the presence of defects
both on fibre surfaces, and in their interior. The effect of interface properties
on axial strength are of secondary importance, and modelling their effect
on axial strength requires the use of sophisticated stress transfer models
and Monte Carlo simulation techniques. For unprotected glass fibres, it is
well known that the environmental exposure of the composite leads to time
dependent reductions in fibre strength. The strength reduction of the fibres
results because of the progressive growth of fibre defects caused by stress
corrosion cracking at a microscopic level.
Environmental exposure, provided that it is saturated, can lead to a
240 N. McCartney

deterioration in interface properties. Given that the axial strength of a


unidirectional composite is not affected to a great extent by interface prop-
erties, it is reasonable to assume, when modelling the axial behaviour of
a unidirectional composite, that the interfaces in the composite following
prolonged exposure have no ‘strength’. This enables a relatively simply ap-
proach to be taken that will provide good insight into the axial behaviour
of a composite when exposed an aggressive environment.
Because of the dominance of fibre behaviour, earlier modelling work ap-
plied to glass fibre composites (McCartney, 1998; Broughton and McCart-
ney, 1998; Metcalfe et al., 1971; Kelly and McCartney, 1981; McCartney,
1982) regarded the unidirectional composite as a loose bundle of parallel
fibres having equal length, so that the relatively low load carrying capacity
of the matrix was ignored. The glass fibres in the bundle were assumed
to be attached to rigid supports which were able to share the applied load
equally between all surviving fibres. The objective of this report is to ex-
tend the existing loose bundle model so that the load carrying capacity of
the matrix is taken into account when considering the axial behaviour of
glass fibre reinforced composites subject to environmental exposure.

5.1 Model Geometry


Consider a unidirectional fibre reinforced composite having a fibre vol-
ume fraction Vf and matrix volume fraction Vm such that Vf + Vm = 1.
The composite has been wholly immersed in an aggressive environment for
a sufficient time for the composite to be fully saturated. The application of
axial load to the composite leads to the environmental growth of defects in
the fibres; a phenomenon well known to afflict glass fibres. The interfaces
between the fibres and matrix are regarded as being significantly weakened
by the environment to the extent that it can be assumed that the fibres
and matrix behave independently in regions of the composite that are well
away from the uniaxial loading mechanism where clamping effects become
important. This assumption means that the composite can be modelled as
a parallel bar model, as shown in Fig. 8.
The fibres in the composite are regarded as acting as a loose bundle
forming one bar of the model. The matrix material in the composite is
considered as being gathered together to form the other bar of the model
which is regarded as homogeneous, i.e. the bar is solid. When a fibre fails
the load it carried is shared between the surviving fibres in the bundle and
the matrix in such a way that all surviving fibres and matrix experience the
same axial strain increment. The fibres are assumed to have the same length
so that each surviving fibre has the same stress throughout the progressive
Analytical Methods of Predicting Performance… 241

F (t)

Fibre Matrix
bundle
Vf Vm

σ(t) σm (t)

Eb (t) Em

ε(t) ε(t)

Ef

F (t)

Figure 8. Schematic diagram of the parallel bar model of a unidirectional


composite for predicting effects of environmental exposure on axial compos-
ite properties

failure process. The composite is subject to a fixed applied load F for all
times t > 0, where t = 0 corresponds to the time when the fixed load F is
first applied. Environmental defect growth in the fibres leads to progressive
fibre failure until the bundle collapses. It is assumed that bundle collapse
corresponds to the catastrophic failure of the composite, i.e. the matrix
strength is insufficient to maintain the load when all the fibres have failed.
The objective is to develop the parallel bar model of a composite so that
it can predict the dependence of composite life tf on the fixed applied load
F, and the dependence of the residual strength F *(t ) of the composite on
elapsed time t from the instant of first loading.
242 N. McCartney

The behaviour of bundles of loose fibres subject to environmental degra-


dation has been modelled by Kelly and McCartney (1981); McCartney
(1982) for the case when the load applied to the bundle is fixed in time.
For the parallel bar model of the composite which is subject to a fixed load
F, the progressive failure of fibres in the bundle leads to a time dependence
of the effective bundle stiffness, and consequently to a time dependence of
the load applied to the fibre bundle. Thus, the earlier modelling requires
modification if it is to be applied to the prediction of the behaviour of a
uniaxially loaded unidirectional composite material having weak interfaces.

5.2 Basic Mechanics for Parallel Bar Model of a Composite


The analysis of the parallel bar model shown in Fig. 8 will neglect any
axial thermal stresses arising from a mismatch of the thermal expansion
coefficients of the fibres and the matrix. The area fraction of all fibres in
the bundle is denoted by Ab , and that of the matrix is Am . It follows that

Ab Am
Vf = , Vm = = 1 − Vf (230)
Ab + Am Ab + Am

The load applied to the fibre bundle at time t is denoted by Fb (t ), the stress
in each surviving fibre being denoted by σ(t). The cross-sectional area of
each of the fibres in the bundle is denoted by A, and the axial modulus of
each fibres is denoted by Ef which is assumed to be time independent. The
axial stress at time t in the matrix is denoted by σm (t). The modulus of
the matrix is denoted by Em which is assumed to be independent of time.
A time dependence could be included to account for visco-elastic effects, or
for time-dependence arising from matrix ageing.
The axial strain in all surviving fibres of the bundle and the matrix
has the same time dependent value that is denoted by ε(t). As thermal
expansion mismatch effects are neglected it follows that

σ(t) σm (t)
ε(t) = = (231)
Ef Em

The balance of forces in the parallel bar model leads to the equilibrium
relation
Fb (t) + Am σm (t) = F (232)
The number of surviving fibres in the bundle at time t is denoted by N (t )
so that the load applied to the bundle at time t may be written

Fb (t) = N (t)Aσ(t). (233)


Analytical Methods of Predicting Performance… 243

On substituting (233) into (232) followed by the elimination of σm (t) using


(231) it is easily shown that the number of fibres surviving at time t is related
to the fibre stress σ(t) through the following relation that quantitatively
characterises the load sharing that occurs when fibres in the composite fail
 
N (t) F Vm Em
+ α σ(t) = , where α = , (234)
N0 Ab Vf Ef

and where N 0 = N (0) and Ab = N 0 A. This is the generalisation to a com-


posite material of the relation used in the modelling a loose bundle of fibres
subject to environmental degradation (Kelly and McCartney, 1981; McCart-
ney, 1982) which is recovered from (234) on letting α → 0.
It is useful to relate the number of fibres surviving in the bundle at time
t to the effective axial modulus of the bundle Eb (t ). The effective stress
applied to the bundle is defined by

Fb (t)
σb (t) = , (235)
Ab

and since the axial strain of the bundle and the individual fibres has the
value σ(t)

σ(t) σb (t) Fb (t) N (t) σ(t)


ε(t) = = = = , (236)
Ef Eb (t) Ab Eb (t) N0 Eb (t)

where use has been made of (233) and (235). Clearly the effective axial
modulus of the fibre bundle is given by

N (t)
Eb (t) = Ef (237)
N0

The effective stress σapp applied to the composite is defined by

F
σapp = , (238)
Ab + Am

and it can be shown from (230) and (234), together with the fact that
Ab = N0 A, that

σapp (t) = Ec (t) ε(t), where Ec (t) = Vf Eb (t) + Vm Em , (239)

where Ec (t ) is the effective axial modulus of the composite defined by the


rule of mixtures, as to be expected.
244 N. McCartney

5.3 Accounting for Defect Growth


The objective here is to show how the analysis of Kelly and McCartney
(1981), developed for loose bundles exposed to an aggressive environment,
must be modified for application to a unidirectional composite having weak
interfaces. The analysis is based on the assumption that the strength of
individual fibres is determined by surface defects whose effective size and
distribution along the fibre surface is statistically distributed.
Fibre failure is assumed to be governed by a Griffith type of failure
criterion having the form
K 2 = y 2 σ 2 a = KIc
2
, (240)
where K is an effective stress intensity factor for a fibre defect of effective
size a subject to a fibre stress s, KIc is the effective fracture toughness of
the fibre material, and where y is a dimensionless parameter designed to
account for defect geometry. The aggressive environment leads to defect
growth when the fibre is under load. Such defect growth is assumed to be
governed by a growth law of the form
da
= CK n , (241)
dt
where C and n are material constants.
When a constant load is applied to a unidirectional composite, exposed
to an aggressive environment to the point of saturation, the fibre defects
grow in size according to the growth law (241) eventually leading to fibre
failure when the failure criterion (240) is satisfied. Thus fibres progressively
fail and the load carried by failed fibres is, for the parallel bar model under
discussion, transferred to the surviving fibres and matrix using the load
sharing rule (234). The stress in each fibre of the system is thus time
dependent. It is useful to present here the relationship that determines
the initial defect size X 0 (t ) that requires a time t to grow to the critical
size ac (t ), at which the fibre stress is s(t ), under the influence of a time
dependent fibre stress history σ(τ ); 0 < τ < t . The critical defect size at
time t is predicted by (240) to be
 2
KIc
ac (t) = , (242)
yσ(t)
and it can be shown on integrating (241) between the limits X 0 (t ) and ac (t )
that
⎡ ⎤ 2−n
2

2 )t
KIc
X0 (t) = 2 ⎣σ n−2 (t) + (n − 2)λ σ n (τ )dτ ⎦ , (243)
y
0
Analytical Methods of Predicting Performance… 245

where
1 n−2 2
λ= CKIc y (244)
2
On using (240) it follows from (243) that the initial strength σi (t) of the
fibres, that fail at time t when their stress is σ(t), is given by
⎡ ⎤ n−2
1
)t
σi (t) = ⎣σ n−2 (t) + (n − 2)λ σ n (τ )dτ ⎦ (245)
0

The cross-sectional area of the sample of unidirectional composite is as-


sumed to be large enough for there to be a very large number of fibres.
It can then be assumed that the bundle of fibres used in the parallel bar
model contains every possible fibre strength that can arise in the statistical
distribution. It is assumed that the strength distribution of the fibres is
given by the two parameter Weibull distribution (Weibull, 1951) so that,
for a large bundle of N 0 fibres, the expected number of fibres N surviving
when the stress in each fibre is σ is given by

m 
σ
N = N0 exp − (246)
σ0
where σ0 is a scaling parameter that will depend on the length of the com-
posite.

5.4 Prediction of Static Strength


It is useful to investigate the prediction of the static strength of a uni-
directional composite assuming that the parallel bar model is valid. When
using (246) in conjunction with (234) it is easily shown that the total load
carried by the composite, when the stress in surviving fibres has the value
σ is given by  
m
F̂ = σ̂ α + e−σ̂ , (247)

where F̂ and σ̂ are a dimensionless normalised load and stress respectively


defined by
F σ
F̂ = , σ̂ = (248)
N0 σ0 A σ0
The static strength of the composite is the maximum value of the load that
can be carried by the composite. The maximum load occurs when F̂ has a
local maximum when plotted as a function of σ̂. The maximum fibre stress
σmax satisfies the transcendental equation
m
m
m σ̂max = 1 + α eσ̂max (249)
246 N. McCartney

The corresponding static strength for the composite is then obtained using

Fmax  m
 m+1
mσ̂max
= F̂max = σ̂max α + e−σ̂max = α m −1
, (250)
N0 A σ0 m σ̂max

which is consistent with the known result for a loose bundle Kelly and
McCartney (1981) when the limit α → 0 is taken.
The equation (249) governing the maximum fibre stress does not always
have a solution as is easily seen by examining the form of the LHS and RHS
m
of (249). On letting x = σ̂max the critical conditions defining the limit of
solutions to (249) may be written

mx = 1 + αex , m = αex (251)

These conditions correspond to the touching of the curves y = mx − 1 and


y = αex . It is easily seen that the critical condition occurs when
m 1+m
x = ln = (252)
α m
It is concluded that the equation (249) has a solution only if
1
α < me−(1+ m ) (253)

If this condition is not satisfied then it is deduced that the fibres progres-
sively fail until there is just one surviving fibre which will then fail, i.e. the
bundle does not suddenly collapse. The value of the Weibull modulus m for
fibres of interest is usually such that the condition (253) is satisfied so that
bundle collapse is always expected in practice.

5.5 Prediction of Progressive Damage


At time t the fibres that survive in the composite are those whose initial
strengths were greater than σi (t) defined by (245). It then follows from
(246) that the expected number of surviving fibres N (t ) at time t is given
by 
m 
N (t) σi (t)
= N̂ (t) = exp − (254)
N0 σ0
On substituting (245) in (254)
  n−2
m )t
1
ln = σ̂ n−2
(t) + (n − 2)η σ̂ n (τ )dτ, (255)
N̂ (t)
0
Analytical Methods of Predicting Performance… 247

where
1 n−2 2 2
η = λσ02 =
CKIc y σ0 , (256)
2
and where use has been made of the definitions (248), which when applied
to the load sharing rule (234) lead to


N̂ (t) = −α (257)
σ̂(t)

On substituting (257) into (255)


  n−2
m )t
σ̂(t)
ln = σ̂ n−2
(t) + (n − 2)η σ̂ n (τ )dτ. (258)
F̂ − ασ̂(t)
0

On differentiating (258) with respect to t, the following differential equation


governing the time dependence of the normalised fibre stress σ̂(t) is obtained
⎡   n−m−2 ⎤
m

⎣ 1 F̂ σ̂(t) dσ̂(t)
ln − σ̂ n−2 (t)⎦ = σ̂ n+1 (t) (259)
m F̂ − ασ̂(t) F̂ − ασ̂(t) d(ηt)

This differential equation is solved by standard numerical techniques subject


to the initial condition
σ̂(0) = s0 , (260)
where s 0 is the solution of the transcendental equation
 m

F̂ = s0 α + e−s0 , (261)

corresponding to (247), that must be solved numerically.

5.6 Predicting the Failure Stress and Time to Failure


The structure of the differential equation is such that dσ̂/dt → ∞ when
σ̂(t) → σ̂f where
  n−m−2
m
1 F̂ σ̂f
ln = σ̂fn−2 . (262)
m F̂ − ασ̂f F̂ − ασ̂f

The stress σ̂f in the surviving fibres when the composite fails can thus be
determined using numerical methods without having to solve the differential
equation (259). It should be noted that when F̂ = F̂max the solution of (262)
248 N. McCartney

is given by σ̂f = σ̂max where σ̂max and F̂max are given by (249) and (250)
respectively. The time to failure for the composite is denoted by tf .
The number of surviving fibres just before composite failure is obtained
using (254) and is given by

N (tf ) m σi (tf )
= N̂f = e−σ̂i , σ̂i = (263)
N0 σ0
The transcendental equation (262), that usually must be solved numerically,
involves the dimensionless loading parameter F̂ in a complicated way. It
is useful to unravel the dependence on this parameter by using the load
sharing rule (257) to express (262) in terms of Nf as follows
 n−1   n−m−2
1 N̂ f + α 1
m

F̂ n−2 = ln (264)
m N̂f N̂f

Having solved (264) to find N̂f using numerical methods, the normalised
failure stress is obtained, on making use of (234), from the relation


σ̂f = (265)
N̂f + α

The time to failure tf can be predicted only by solving the differential


equation (259) in the normalised stress range s0 ≤ σ̂(t) ≤ σ̂f .

5.7 Predicting Residual Strength


A key requirement concerning the effects of environment on composite
degradation is the prediction of the time dependence of the residual strength
of a composite. This has already been considered for the case of a loose bun-
dle of fibres (McCartney, 1982). The objective now is to extend the analysis,
and simplify it so far as is possible, so that the residual strength of a unidi-
rectional composite with weak interfaces can be predicted. After an elapsed
time t from the application of a fixed load F, the load is instantaneously
increased until the composite fails catastrophically. Just before the load is
suddenly increased the stress in the surviving fibres has the value σ(t) and
at any stage during the subsequent instantaneous load increase the value of
the stress in the fibres is denoted by s. When the fibre stress has the value
s the critical defect size has the following value specified by (240)

2
KIc
a∗c = (266)
ys
Analytical Methods of Predicting Performance… 249

It is necessary to calculate the original size X ∗ of the critical defect using


the relations (240) and (241). It is easily shown that

)t

2−n 2−n 1
(X ) 2
= (a∗c ) 2
+ C (n − 2)y n σ n (τ )dτ (267)
2
0

On using (240) the initial strength of the fibres that are critical at time t
when the fibre stress has the value s is denoted by si and is given, on using
(267), by the relation

)t
ŝn−2
i = ŝ n−2
+ (n − 2)η σ̂ n (τ )dτ, (268)
0

where use has been made of (256) and where


si s
ŝi = , ŝ = (269)
σ0 σ0
On using (258) the relation (268) may be written in the form

ŝn−2
i = ŝn−2 − k(t), (270)

where
  n−2
m
σ̂(t)
k(t) = σ̂ n−2 (t) − ln (271)
F̂ − ασ̂(t)
The load applied to the composite Fs , when the fibre stress has the value
s, is obtained from (248) and (257) so that
Fs  
= F̂s = ŝ α + N̂s , (272)
N0 σ0 A

where F̂s is the normalised applied load and where N̂s is the normalised
number of surviving fibres when the load on the composite is such that the
fibre stress has the value s. It follows from (246) that
m
N̂s = e−ŝi (273)

On substituting in (272) the following expression is derived for the nor-


malised load applied to the composite during a residual strength test
 m

F̂s = ŝ α + e−ŝi . (274)
250 N. McCartney

The residual strength S (t ) of the composite at time t is the maximum value


of Fs when s is varied, or alternatively the maximum value of F̂s when ŝ
is varied. Noting that k (t ) is independent of ŝ, the maximum value of F̂s
occurs when ŝi = x (t ) which satisfies the transcendental equation


m k(t)
1 + αex (t) = mxm (t) 1 + n−2 (275)
x (t)

On using (270) the stress σmax (t) in the surviving fibres just before the
composite fails during a residual strength test is obtained from

σmax (t)  n−2  1


= x (t) + k(t) n−2 = σ̂max (t). (276)
σ0

It then follows from (270) and (274) that the residual strength of the com-
posite S (t ) is obtained using

S(t)  m

= σ̂max (t) α + e−x (t) = Ŝ(t). (277)
σ0

When t = 0 it can be shown using (258) that k (0) = 0 in which case the
transcendental equation (275) reduces to the form (249) which needs to be
solved when calculating the static strength of the composite.

5.8 Example Prediction


In order to assess the properties of the model some example predictions
have been made to illustrate the principal characteristics. There are four
parameters that need to be specified in order to obtain predictions:
1. The Weibull exponent characterising the strength distribution of the
fibres before environmental exposure. This parameter, which often
has values in the range 4 – 8, appears in the relation (246) defining
the expected number of fibre failures for a given fibre stress in a static
loading test. The value m = 8 will be used here. The value of m is
usually obtained from single fibre strength tests.
2. The exponent n appearing in the defect growth law (241). This pa-
rameter, which usually has values in the range 3 – 30, is often obtained
from stress corrosion cracking tests carried out using monolithic glass
testpieces rather than fibres. The value n = 20 is used here.
3. The ratio α defined by (234) which takes approximate account of the
properties of the fibre and matrix, and also of the fibre volume fraction.
For many glass fibre composites of interest, the value of α lies in the
range 0 - 0.1. It will be assumed here that α = 0.025.
Analytical Methods of Predicting Performance… 251
1

0.8

0.6
F/Fm

F/Fm
F ∗ /Fm = 0.2
0.4
F ∗ /Fm = 0.4
F ∗ /Fm = 0.6
0.2

0
-10 -5 0 5 10 15
log10 (ηt)
Figure 9. Schematic diagram of the parallel bar model of a unidirectional
composite for predicting effects of environmental exposure on axial compos-
ite properties

4. The level of loading applied axially to the composite where the model
assumes that the ratio F/Fm is given where F is the axial load applied
to the composite and Fm is the static strength, i.e. the strength of the
composite before environmental exposure. The value of F/Fm always
lies in the range 0 – 1.
The Euler-Richardson solution technique (Churchhouse, 1981) is used to
solve the ordinary differential equation (259) where the normalised dimen-
sionless time η t may be regarded as an unknown function of σ̂. In other
words, the differential equation can be used directly to determine an incre-
ment in the value of η t for any given increment in σ̂. The initial condition
is specified by (260) and (261) and the range s0 ≤ σ̂ ≤ σ̂f is subdivided into
100 equal intervals when solving the differential equation. The upper limit
σ̂f is determined by the relations (264) and (265). Figure 9 shows the result
of solving the differential equation (259) to find the normalised time η tf
for various values of the loading ratio F/Fm . The normalising parameter
η is defined by (256). It is seen that as F/Fm → 1 the lifetime tends to
zero. Figure 9 also shows predictions of the normalised residual strength Ŝ
defined by (277), as a function of the normalised time η t.
The principal conclusion to be drawn from the results presented is that
the time dependence of the axial properties of a unidirectional fibre rein-
forced glass composite subject to environmental exposure under fixed load
252 N. McCartney

can be predicted using a parallel bar model of the composite where interface
bonding is neglected. The model enables the prediction of the stress history
of the fibre stress in surviving fibres from the point of first loading to the
occurrence of catastrophic failure. Results not shown indicate that the fibre
stress is almost independent of the matrix properties, a situation that arises
because Em  Ef .
The model can also be used to predict the time dependence of the resid-
ual strength of the composite, a property which does show some dependence
on matrix properties. However, results not shown indicate that, when the
residual strength is divided by the static strength, the resulting residual
strength ratio is virtually independent of the matrix properties. It is con-
cluded that the residual strength ratio for a unidirectional composite is
predictable (and therefore measurable) from the static strength of the com-
posite, and the time dependence of the residual strength of a loose bundle
of fibres.

6 Closing Remarks
A varied set of topics concerning the behaviour of composite materials has
been considered in this paper. They concern the estimation of the undam-
aged properties of plies in terms of fibre and matrix properties, the esti-
mation of the undamaged properties of general symmetric laminates, the
consideration of an elegant method of considering cracks in anisotropic ma-
terials using orthogonal polynomials, a detailed treatment of ply cracking
in a simple cross-ply laminate, and the modelling of the effects of envi-
ronmental exposure on the lifetime and residual strength of unidirectional
composites. Much of the work presented here has not been published be-
fore. For the analyses dealing with composite damage, example predictions
have been given to help readers understand the capabilities of the various
damage models.
It is hoped that readers of this paper will be convinced that analytical
modelling, which has been undertaken in some quite complex situations,
enables much deeper insight into the modelling of composite material sys-
tems than numerical solution methods permit, and provides opportunities
for convenient design methods based on relatively compact formulae rather
than on data tables and graphs that have to be generated when using nu-
merical methods such as finite element analysis.

c Crown Copyright. Reproduced by permission of the Controller of HMSO


and the Queen’s printer for Scotland
Analytical Methods of Predicting Performance… 253

Bibliography
W.R. Broughton and L.N. McCartney. Predictive models for assessing
long-term performance of polymer matrix composites. Technical Report
CMMT(A)95, NPL, Teddington, April 1998.
R.F. Churchhouse. Numerical methods. In W. Ledermann, editor, Handbook
of Applicable Mathematics, volume 3, pages 319–321, Chichester, 1981.
Wiley.
G.M.L. Gladwell and A.H. England. J. Mech. Appl. Math., 30:175, 1977.
Z. Hashin. Analysis of composite materials - a survey. Trans. ASME. J.
Appl. Mech., 50:481–505, 1983.
A.S. Kaddour, M.J. Hinton, P.A. Smith, and S. Li. The background to the
third world-wide failure exercise. J. Comp. Mater., 47(20-21):2417–2426,
2013a.
A.S. Kaddour, M.J. Hinton, P.A. Smith, and S. Li. A comparison between
the predictive capability of matrix cracking, damage and failure criteria
for fibre reinforced laminates: part a of the third world-wide failure
exercise. J. Comp. Mater., 47(20-21):2749–2779, 2013b.
A. Kelly and L.N. McCartney. Failure by stress corrosion of bundles of
fibres. Proceedings of the Royal Society of London, A374:475–489, 1981.
J.C. Maxwell. A Treatise on Electricity and Magnetism, volume 1. Claren-
don Press, Oxford, 1st edition, 1873.
L.N. McCartney. Time dependent strength of large bundles of fibres loaded
in corrosive environments. Fibre Science & Technology, 16:95–109, 1982.
L.N. McCartney. Theory of stress transfer in a 0-90-0 cross-ply laminate
containing a parallel array of transverse cracks. J. Mech. Phys. Solids,
40:27–68, 1992.
L.N. McCartney. Model of composite degradation due to environmental
damage. Technical Report CMMT(A)124, NPL, Teddington, September
1998.
L.N. McCartney. Maxwell’s far-field methodology predicting elastic prop-
erties of multiphase composites reinforced with aligned transversely
isotropic spheroids. Phil. Mag., 90:4175–4207, 2010.
L.N. McCartney. Derivations of energy-based modelling for ply cracking
in general symmetric laminates. J. Comp. Mater., 47(20-21):2641–2673,
2013a.
L.N. McCartney. Energy methods for modelling damage in laminates. J.
Comp. Mater., 47(20-21):2613–2640, 2013b.
L.N. McCartney and T.A.E. Gorley. Complex variable method of calculating
stress intensity factors for cracks in plates. In A.R. Luxmoore et al.,
editor, Proc. 4th Int. Conf. on Numerical Methods in Fracture Mechanics,
Swansea, 1987. Pineridge Press.
254 N. McCartney

L.N. McCartney and A. Kelly. Maxwell’s far-field methodology applied to


the prediction of properties of multi-phase isotropic particulate compos-
ites. Proc. Roy. Soc., A464:423–446, 2008.
A.G. Metcalfe, M.E. Gulden, and G.K. Schmitz. Glass Technology, 12:
15–23, 1971.
D.P. Rooke and D.J. Cartwright. Compendium of Stress Intensity. Hilling-
don Press, Uxbridge, 1st edition, 1976.
G.C. Sih and H. Liebowitz. Mathematical theories of brittle fracture. In
H. Liebowitz, editor, Fracture, volume II, page 67, San Diego, 1968.
Academic Press.
W. Weibull. A statistical distribution function of wide applicability. Journal
of Applied Mechanics, 19:293–297, 1951.

You might also like