You are on page 1of 315

Damage and Failure of Composite Materials

Understanding damage and failure of composite materials is critical for reliable


and cost-effective engineering design. Bringing together materials mechanics and
modeling, this book provides a complete guide to damage, fatigue, and failure of
composite materials. Early chapters focus on the underlying principles governing
composite damage, reviewing basic equations and mechanics theory, before
describing mechanisms of damage such as cracking, breakage, and buckling. In
subsequent chapters, the physical mechanisms underlying the formation and
progression of damage under mechanical loads are described with ample experi-
mental data, and micro- and macro-level damage models are combined. Finally,
fatigue of composite materials is discussed using fatigue-life diagrams. While there
is a special emphasis on polymer matrix composites, metal and ceramic matrix
composites are also described. Outlining methods for more reliable design of
composite structures, this is a valuable resource for engineers and materials
scientists in industry and academia.

Ramesh Talreja is a Professor of Aerospace Engineering at Texas A&M University.


He earned his Ph.D. and Doctor of Technical Sciences degrees from the Technical
University of Denmark. He has contributed extensively to the fields of damage,
fatigue, and failure of composite materials by authoring numerous books and
book chapters as well as by editing several encyclopedic works.
Chandra Veer Singh is an Assistant Professor of Materials Science and Engineering at
the University of Toronto. He earned his Ph.D. in aerospace engineering from Texas
A&M University, and worked as a post-doctoral Fellow at Cornell University. His
research expertise is in damage mechanics of composite materials, atomistic
modeling, and computational materials science. His industry experience includes
R&D at GE Aircraft Engines.
Damage and Failure
of Composite Materials
RAMESH TALREJA
Texas A&M University

C H A N D R A V E E R SI N G H
University of Toronto
cambridge university press
Cambridge, New York, Melbourne, Madrid, Cape Town,
Singapore, São Paulo, Delhi, Mexico City
Cambridge University Press
The Edinburgh Building, Cambridge CB2 8RU, UK

Published in the United States of America by


Cambridge University Press, New York

www.cambridge.org
Information on this title: www.cambridge.org/9780521819428

# R. Talreja and C. V. Singh 2012

This publication is in copyright. Subject to statutory exception


and to the provisions of relevant collective licensing agreements,
no reproduction of any part may take place without
the written permission of Cambridge University Press.

First published 2012

Printed in the United Kingdom at the University Press, Cambridge

A catalogue record for this publication is available from the British Library

Library of Congress Cataloging-in-Publication Data


Talreja, R.
Damage and failure of composite materials / Ramesh Talreja, Chandra Veer Singh.
p. cm.
Includes bibliographical references.
ISBN 978-0-521-81942-8 (Hardback)
1. Composite materials–Fatigue. 2. Composite materials–Fracture.
I. Singh, Chandra Veer. II. Title.
TA418.9.C6T338 2012
620.10 126–dc23
2011035578

ISBN 978-0-521-81942-8 Hardback

Cambridge University Press has no responsibility for the persistence or


accuracy of URLs for external or third-party internet websites referred to
in this publication, and does not guarantee that any content on such
websites is, or will remain, accurate or appropriate.
Contents

Preface page ix

1 Durability assessment of composite structures 1


1.1 Introduction 1
1.2 Historical development of damage mechanics of composites 3
1.3 Fatigue of composite materials 5
References 7

2 Review of mechanics of composite materials 9


2.1 Equations of elasticity 9
2.1.1 Strain–displacement relations 9
2.1.2 Conservation of linear and angular momenta 10
2.1.3 Constitutive relations 11
2.1.4 Equations of motion 15
2.1.5 Energy principles 15
2.2 Micromechanics 17
2.2.1 Stiffness properties of a unidirectional lamina 18
2.2.2 Thermal properties of a unidirectional lamina 19
2.2.3 Constitutive equations for a lamina 20
2.2.4 Strength of a unidirectional lamina 21
2.3 Analysis of laminates 24
2.3.1 Strain–displacement relations 25
2.3.2 Constitutive relationships for the laminate 26
2.3.3 Stresses and strains in a lamina within a laminate 28
2.3.4 Effect of layup configuration 28
2.4 Linear elastic fracture mechanics 29
2.4.1 Fracture criteria 30
2.4.2 Crack separation modes 31
2.4.3 Crack surface displacements 32
2.4.4 Relevance of fracture mechanics for damage analysis 33
References 34
vi Contents

3 Damage in composite materials 36


3.1 Mechanisms of damage 37
3.1.1 Interfacial debonding 37
3.1.2 Matrix microcracking/intralaminar (ply) cracking 39
3.1.3 Interfacial sliding 39
3.1.4 Delamination/interlaminar cracking 41
3.1.5 Fiber breakage 42
3.1.6 Fiber microbuckling 42
3.1.7 Particle cleavage 44
3.1.8 Void growth 44
3.1.9 Damage modes 45
3.2 Development of damage in composite laminates 46
3.3 Intralaminar ply cracking in laminates 49
3.4 Damage mechanics 50
References 52

4 Micro-damage mechanics 57
4.1 Introduction 57
4.2 Phenomena of single and multiple fracture: ACK theory 58
4.2.1 Multiple matrix cracking 61
4.2.2 Perfectly bonded fiber/matrix interface: a modified shear
lag analysis 65
4.2.3 Frictional fiber/matrix interface 67
4.3 Stress analysis (boundary value problem) for cracked laminates 68
4.3.1 Complexity and issues 68
4.3.2 Assumptions 71
4.4 One-dimensional models: shear lag analysis 73
4.4.1 Initial shear lag analysis 74
4.4.2 Interlaminar shear lag analysis 77
4.4.3 Extended shear lag analysis 79
4.4.4 2-D shear lag models 80
4.4.5 Summary of shear lag models 80
4.5 Self-consistent scheme 84
4.6 2-D stress analysis: variational methods 87
4.6.1 Hashin’s variational analysis 87
4.6.2 Effect of residual stresses 96
4.6.3 [0m/90n]s vs. [90n/0m]s laminates 97
4.6.4 Improved variational analysis 97
4.6.5 Related works 101
4.6.6 Comparison between 1-D and 2-D stress-based models 101
4.7 Generalized plain strain analysis – McCartney’s model 104
Contents vii

4.8 COD-based methods 110


4.8.1 3-D laminate theory: Gudmundson’s model 111
4.8.2 Lundmark–Varna model 117
4.9 Computational methods 119
4.9.1 Finite element method (FEM) 120
4.9.2 Finite strip method 121
4.9.3 Layerwise theory 123
4.10 Other methods 124
4.11 Changes in thermal expansion coefficients 125
4.12 Summary 126
References 126

5 Macro-damage mechanics 134


5.1 Introduction 134
5.2 Continuum damage mechanics (CDM) of composite materials 138
5.2.1 RVE for damage characterization 139
5.2.2 Characterization of damage 141
5.2.3 A thermodynamics framework for materials response 144
5.2.4 Stiffness–damage relationships 148
Case 1: Cracking in one off-axis orientation 152
Case 2: Cross-ply laminates 152
Evaluation of material constants 153
5.3 Synergistic damage mechanics (SDM) 155
5.3.1 Two damage modes 156
5.3.2 Three damage modes 165
5.4 Viscoelastic composites with ply cracking 170
5.5 Summary 176
References 177

6 Damage progression 179


6.1 Introduction 179
6.2 Experimental techniques 180
6.3 Experimental observations 185
6.3.1 Initiation of ply cracking 185
6.3.2 Crack growth and multiplication 187
6.3.3 Crack shapes 189
6.3.4 Effect of cracking 189
6.3.5 Loading and environmental effects 191
6.3.6 Cracking in multidirectional laminates 193
6.4 Modeling approaches 194
6.4.1 Strength-based approaches 194
6.4.2 Energy-based approaches 198
6.4.3 Strength vs. energy criteria for multiple cracking 210
viii Contents

6.5 Randomness in ply cracking 211


6.6 Damage evolution in multidirectional laminates 217
6.7 Damage evolution under cyclic loading 223
6.8 Summary 229
References 230

7 Damage mechanisms and fatigue-life diagrams 237


7.1 Introduction 237
7.2 Fatigue-life diagrams 237
7.3 On-axis fatigue of unidirectional composites 238
7.4 Effects of constituent properties 241
7.5 Unidirectional composites loaded parallel to the fibers 242
7.5.1 Polymer matrix composites (PMCs) 242
7.5.1.1 Experimental studies of mechanisms 247
7.5.2 Metal matrix composites (MMCs) 250
7.5.3 Ceramic matrix composites (CMCs) 252
7.6 Unidirectional composites loaded inclined to the fibers 257
7.7 Fatigue of laminates 259
7.7.1 Angle-ply laminates 260
7.7.2 Cross-ply laminates 261
7.7.3 General multidirectional laminates 263
7.8 Fatigue-life prediction 265
7.8.1 Cross-ply laminates 266
7.8.2 General laminates 273
7.9 Summary 273
References 274

8 Future directions 276


8.1 Computational structural analysis 276
8.2 Multiscale modeling of damage 278
8.2.1 Length scales of damage 280
8.2.2 Hierarchical multiscale modeling 282
8.2.3 Implication on multiscale modeling: Synergistic damage
mechanics 286
8.3 Cost-effective manufacturing and defect damage mechanics 287
8.3.1 Cost-effective manufacturing 288
8.3.2 Defect damage mechanics 291
8.4 Final remarks 296
References 298

Author index 301


Subject index 303
Preface

The field of composite materials has advanced steadily from the early develop-
ments during the 1970s when laminate plate theory and anisotropic failure criteria
were in focus to today’s diversification of composite materials to multifunctional
and nanostructured composite morphologies. Throughout the 1970s and 1980s
several books appeared along with courses that were developed and taught at
advanced levels dealing with mechanics of composite materials and structures.
The failure analysis was mostly limited to descriptions of strength that extended
previous continuum descriptions of metal yielding and failure. Beginning around
the mid-1980s, micromechanics and continuum damage mechanics were applied
to multiple cracking observed in composite materials. Under the overall descrip-
tion of “damage mechanics” a flurry of activities took place as evidenced by
conferences and symposia. Other than several conference proceedings that
recorded such activities, a collection of seminal contributions to the field appeared
in a volume (Damage Mechanics of Composite Materials, R. Talreja, ed., Composite
Materials Series, R.B. Pipes, series ed., Vol. 9, Amsterdam: Elsevier Science
Publishers, 1994). The two main avenues of approach to damage in composite
materials and its effect on materials response, now referred to as micro-damage
mechanics (MIDM) and macro-damage mechanics (MADM), were presented in a
balanced form in that volume. In the years since then, many developments have
taken place that have brought this field to such level of maturity that a book
coherently presenting the material was felt to be timely. It is hoped that this book
will help provide impetus for teaching advanced courses in composite damage at
universities as well as support short courses for professional development of
engineers in industry. The wealth of material covered can also help new research-
ers in advancing the field further. To this end, the last chapter provides some
guidance in identifying gaps and needs for further work.
The structure of the book is as follows. Chapter 1 lays down the overall strategy
for durability assessment of composite structures, emphasizing the needs and
motivating the content of the book to follow. Chapter 2 provides an easy reference
to the basic continuum mechanics topics that are felt to be relevant to the
subsequent treatment. Chapter 3 describes the mechanisms of damage that under-
lie the phenomena aimed for modeling. Many of the physical observations
described there are viewed to be vital to developing proper understanding of the
complex field of damage in composite materials. Chapters 4 and 5 deal with the
x Preface

two main approaches stated above, i.e., the MIDM and the MADM. Selected
works from the literature, including the authors’ own, are given as much treatment
as was found justified to generate coherency without overly including details.
While these two chapters focus on descriptions of damage and the constitutive
property changes caused by it, Chapter 6 is devoted to the progression of damage.
The crack multiplication is a distinctive feature of damage in composite materials
that distinguishes it from single crack growth in monolithic materials, and there-
fore justifies treatment in a chapter by itself. Chapter 7 is on fatigue of composite
materials. This field suffers from the historical treatments of metal fatigue and is
unfortunately the least understood part of damage in composite materials. Multi-
axial fatigue illustrates the situation well where the literature displays little under-
standing of the mechanisms underlying failure. While a separate book on fatigue
of composite materials is needed to do full justice to the field, a single chapter here
is added to draw attention to the mechanisms-based concepts for proper interpret-
ation and modeling. Finally, Chapter 8 presents a summary of the book and
points to the directions in which further advances are seen to be necessary.
Particular emphasis is given to the computational incorporation of damage mod-
eling in durability assessment as well as taking account of the manufacturing-
induced defects in an integrated manner.
Although the authors have written this book, the credit goes to many research-
ers who have worked on various aspects of damage in composite materials. Their
collective contributions have made it possible for us to present what we have seen
as a coherent story at this time. The field is evolving, and future versions of the
story will hopefully spur further development and, most importantly, transfer of
this knowledge to industry will take place.
1 Durability assessment of composite
structures

1.1 Introduction

Composite structures for mechanical and aerospace applications are designed to


retain structural integrity and remain durable for the intended service life. Since
the early 1970s important advances have been made in characterizing and model-
ing the underlying mechanical behavior and developing tools and methodologies
for predicting the fracture and fatigue of composite materials. This book provides
an exposition of the concepts and analyses related to this area and presents recent
results. The next chapters treat damage in composite materials as observed by a
variety of techniques, followed by modeling at the micro and macro levels. Fatigue
is treated separately because of its particular complexities that require systematic
interpretation schemes developed for the purpose. A chapter is added in the
beginning to provide convenient access to the mechanics concepts needed for the
modeling analyses in later chapters.
Here we present an overview of the durability assessment process for composite
structures. Figure 1.1 depicts the connectivity and flow of the elements of this
process. To begin, one usually conducts stress analysis of the component using the
“initial” constitutive behavior of the composite along with the service loading on
the component as input. In contrast to monolithic materials, such as metals, the
constitutive behavior of a composite can change due to damage incurred in
service. The stress analysis combined with prior experience allows identifying
critical sites (“hot spots”) in the component that are prone to be the sites of
failure. Further examination of these sites in terms of the local stress/strain/
temperature excursions combined with the composite material composition at
those sites helps to identify the possible mechanisms of damage that can result.
Examples of such mechanisms are microcracking of the matrix, delamination
(separation of layers at interfaces), aging (of the polymer matrix), etc. Chapter 3
describes these mechanisms in some detail. The next step is to analyze the conse-
quences of the mechanisms on the material response and in turn on the structural
performance. Chapters 4 and 5 deal with different models to predict the damage-
induced material response changes. Since the scales at which damage occurs are
small in comparison to the characteristic geometrical size of the “hot spots,”
models must account for the multiple length (or size) scales. The differentiation
of scales is conventionally described as “micro” (the scale of damage) and
2 Durability assessment of composite structures

Figure 1.1. A durability analysis scheme for composite structural components.

“macro” (the scale at which structural response is characterized). Since connectiv-


ity between these scales must be established, an intermediate scale called “meso” is
defined as needed by the particular model used. In micromechanics the concept of
“representative volume element” (RVE) has been proposed. The size of this
element is commonly taken to be the meso scale. Chapters 4 and 5 describe
the three scales in the context of different models. Chapter 6 is focused on the
initiation and progression of damage. Together the three chapters provide the
content of the subject known as “damage mechanics,” which as indicated in
Figure 1.1 is central to durability assessment.
The common output of the damage mechanics models is a description of the
material response, often described as “stiffness degradation,” caused by damage.
This description necessarily involves averaging over the so-called RVE. Thus the
materials response, or averaged constitutive behavior description, forms the new
input to the stress analysis that was conducted initially using pristine (undamaged)
material properties. The resulting iterative process of stress analysis should be an
inherent feature of composite structural analysis, although the industry practice
currently does not fully implement this procedure. Another output of the damage
mechanics analysis is “strength degradation,” i.e., reduction in the load-bearing
capability of the structure due to damage. Depending on the functional requirements
of a given structure, degradation of stiffness or strength would be the path to loss
of structural integrity. A typical example of a stiffness-critical structure is an
aircraft wing that must deform appropriately to perform its aerodynamic function,
while a fuselage is strength-critical as its design requirement is to contain the pressure
within it.
While monolithic materials such as metals fail due to unstable growth of a
crack, the heterogeneous internal structure of a composite leads to formation of
multiple cracks. A generic heterogeneous solid is illustrated in Figure 1.2 in three
states: pristine (undamaged) to the left in the figure shows a representative region
1.2 Historical development of damage mechanics of composites 3

t t
t u1
u u2

Figure 1.2. A heterogeneous solid in pristine (undamaged) state (left) and in two possible
multiple cracking states (middle and right).

of the solid within which heterogeneities (reinforcements) are indicated symbolic-


ally as filled circles, and two states that have multiple cracks resulting from
debonding of reinforcements (middle figure) and from local failure of the matrix
induced by defects and/or stress concentrations. Consider the external loading on
a composite structure resulting in tractions t on the surface bounding the repre-
sentative region of the composite shown. If the response to these tractions in terms
of the bounding surface displacements is given by u in the pristine state, then the
surface displacements of the multiple cracks (commonly expressed as crack
opening displacements, COD, and crack sliding displacement, CSD) within
the volume will change this to u1 or u2 depending on the type of damage (see
Figure 1.2). The local environment around the cracks influences the COD and
CSD of distributed cracks within the volume. This local environment is typically
described as a “constraint” (i.e., moderation) to the crack surface displacements
and is expressed in terms of the variables of heterogeneities. If the heterogeneous
solid with multiple cracks is homogenized over the representative region, then the
stress–strain response averaged over the RVE is given by the averaged stiffness
properties that change (degrade) with increasing number of cracks and the con-
straint to the crack surface displacements. This stiffness degradation is the subject
of damage mechanics, as discussed above in describing the durability assessment
procedure depicted in Figure 1.1.

1.2 Historical development of damage mechanics of composites

Although the field of solid mechanics applied to heterogeneous solids was developed
in the late 1950s and early 1960s, and became known as micromechanics, the specific
situations encountered in composite materials such as those with continuous fiber
reinforcements were not addressed until much later. The concepts developed in
micromechanics turned out to be useful for multiple cracking in composite materials
and are recommended as essential background (see the text by Nemat-Nasser and
Hori [1]). However, the first pioneering work that clarified the phenomenon of
multiple cracking in the presence of fiber/matrix interfaces in reinforced composites
was by Aveston et al. [2] published as a conference proceedings paper in 1971.
4 Durability assessment of composite structures

This work, which became known as the ACK theory, treated multiple parallel cracks
normal to fibers in a matrix with all fibers in one direction loaded in tension along
fibers. The model produced an expression for the overall strain at which multiple
cracking occurs based on a simplified stress analysis and energy balance concepts.
The expression provided a basis for assessing the roles of fiber and matrix properties,
their volume fractions, and the fiber diameter in resisting multiple cracking.
The ACK model was motivated by the observation of multiple cracking in brittle
matrix composites such as cement reinforced with steel wires. For polymer matrix
composites, the application of the ACK theory was at first not clear since a ply with
unidirectional fibers of glass or carbon does not have the right conditions for multiple
cracking when loaded in tension along fibers. Garrett and Bailey in 1977 [3] found
that the multiple cracking observed in cross-ply laminates of glass fiber-reinforced
polyester under axial tension could in fact be described well by the case of fully
bonded interfaces treated by Aveston and Kelly [4] in a follow on paper to the ACK
model. This required replacing the matrix and fibers in that model by the transverse
and longitudinal plies, respectively. Garrett and Bailey then repeated with appropri-
ate modification the one-dimensional stress analysis and energy balance consider-
ations used in [4]. Thus began a long series of works that applied the one-dimensional
stress analysis, known as shear lag analysis, which assumes axial load transfer from
cracked to uncracked plies by the shear stress at the interfaces.
The inadequacy of the shear lag analysis to properly provide stresses in the
cracked cross-ply laminate was a severe limitation until a variational analysis-
based two-dimensional approximation appeared in the English literature [5]. This
spurred further work of more accuracy [6] and extension to partially debonded
frictional interfaces [7], while extension to cracked plies of other than transverse
orientation required other approaches [8]. The analyses that use local ply stress
solutions to evaluate overall stiffness degradation are grouped together in “micro-
damage mechanics” (MIDM) and are treated in Chapter 4.
In some ways parallel to the MIDM emerged another approach that became
known as continuum damage mechanics (CDM). Its beginnings are not attrib-
uted to composite materials but to metals undergoing creep. Kachanov in 1958
[9] put forth a concept of a field of internal material discontinuity responsible
for distributed local stress enhancement leading to overall creep strain. Later,
the internal state was called damage and a (hidden) scalar variable D was
associated with it. The continuum now had an internal damage state and
because of its irreversible nature its treatment required thermodynamics, in
particular the Second Law, which places conditions on the entropy changes.
Kachanov’s work stayed relatively unknown until Lemaitre and Chaboche [10]
applied it to analysis of various structural materials with distributed cavities
and cracks. Krajcinovic [11] further enhanced the field by connecting it to
concepts known from fracture mechanics and plasticity and by elaborating
the thermodynamics implications. For composite materials of technological
interests that are constructed with specific symmetries such as orthotropic, the
first work to apply CDM was by Talreja [12] and its companion paper that
1.3 Fatigue of composite materials 5

Stationary microstructure

Evolving
microstructure RVE

Homogenization
of
stationary
microstructure a
n
P
Damage entity

Homogenized continuum
with damage

Figure 1.3. Illustration of the CDM concept for composites.

validated the stiffness degradation relationships by experimental data [13]. The


CDM concept for composites is illustrated in Figure 1.3. The reinforcements in
a composite are regarded as a stationary microstructure and are homogenized
as an anisotropic medium in which the damage entities such as cracks are
embedded. Further homogenization is done by smearing out the damage
entities into an internal field, which is represented by a pair of vectors, whose
dyadic product averaged over all damage entities in a RVE provides the
characterization of damage. Since the early papers [12, 13] the CDM field for
composite materials has developed steadily, more recently in a version named
as synergistic damage mechanics (SDM) where the micromechanics is judi-
ciously applied to enhance the applicability of CDM. All this is the subject of
Chapter 5 on macro-damage mechanics (MADM).
As depicted in Figure 1.1, the stress analysis of critical structural sites
requires stress–strain relationships that reflect the presence of damage. These
relationships are developed by combining stiffness degradation and damage evo-
lution. The subject of damage evolution is complex with its own challenges.
Therefore Chapter 6 is devoted exclusively to its treatment.

1.3 Fatigue of composite materials

It is natural to assume that the complexities of damage in composite materials


observed under quasi-static loading would be enhanced when the loading is
applied in a cyclic manner. The experience with metal fatigue indicates that the
fracture surface of a sample failed in fatigue shows distinctly different features
than if failed in the application of a monotonically increasing load. The fracture
surface of a unidirectional fiber-reinforced composite loaded along fibers mono-
tonically or cyclically does not give clear indication of mechanisms preceding
6 Durability assessment of composite structures

failure in either case. In more general fiber architectures, such as laminates and
woven fabric composites, following the events from the first (initiation) to the last
(separation by breakage) is generally difficult. However, as advances in nondes-
tructive observation techniques are made, increasing clarity in mechanisms is
emerging. In the early years of composite fatigue studies in the 1970s, little was
understood of mechanisms and consequently the assumptions made in predictive
models were speculative at best.
One study of a unidirectional glass/epoxy composite made assumptions of
fatigue mechanisms that led to reasonable explanation of the trends in fatigue
life [14]. Following that work, a systematic conceptual framework for interpret-
ation of fatigue damage and failure was proposed by Talreja [15] for more
general cases of loading as well as for more general fiber orientations. The
framework took the form of a two-dimensional plot called a “fatigue life dia-
gram” in which regions of dominant mechanisms were separated. The diagram is
not meant to be a data-fitted S-N curve (historically known as a Wöhler dia-
gram) but as a means of interpreting the roles of fibers, matrix, and interfaces as
well as of laminate configuration parameters such as ply orientation, sequence,
and thickness.
Since the unidirectional composite (or ply) is a basic unit in laminates, the
fatigue life diagram for this composite under tension–tension loading forms the
baseline diagram from which more general cases evolve. This diagram is illustrated
in Figure 1.4 and discussed in detail in Chapter 7. As shown, the vertical axis of
the diagram is the maximum strain attained at the first application of maximum
stress in a load controlled fatigue test. This quantity forms a proper reference to
the loading condition and provides upper and lower limits to the fatigue behavior.
Thus the strain to failure (of fiber) forms the upper limit while the strain corres-
ponding to the fatigue limit (primarily a matrix property) forms the lower limit.
These strain values can always be converted to applied stress, but plotting these in
the diagram allows a systematic and proper interpretation of the roles of the
constituents. The regions indicated in the fatigue life diagram provide clarity of
the governing mechanisms dictated by the constituent properties. The construc-
tion of the diagram for unidirectional composites was initially based on systematic
arguments and logical deduction. Physical evidence to support the diagram was
later presented by an elaborate and tedious experimental study [16].
The fatigue life diagram can also serve the purpose of facilitating mechanisms-
based life prediction modeling. For cross-ply laminate this was demonstrated
in [17]. Generally the path to predictive modeling with account of the underlying
damage mechanisms is long and hard. Consequently, the literature has a prepon-
derance of studies that resort to “failure criteria” that are mostly extensions of
those for static failure with assumed procedures without fundamental validation.
The models are therefore not reliable enough to extend beyond the cases that
formed the impetus for the proposed schemes.
Chapter 7 treats the subject of composite damage with emphasis on mechan-
isms. It is not exhaustive in the sense of including the literature on models for life
References 7

Figure 1.4. Fatigue life diagram of a unidirectional fiber-reinforced composite subjected to


cyclic tension in the fiber direction.

prediction. A recent paper [18] has a fairly thorough examination of the main
models for multiaxial fatigue. It reveals the frustrating situation of lack of reliabil-
ity of the models. In Chapter 7 the main findings of this review are discussed and a
mechanisms-based methodology is proposed.

References

1. S. Nemat-Nasser and M. Hori, Micromechanics: Overall Properties of Heterogeneous


Materials. (Amsterdam: Elsevier, 1993).
2. J. Aveston, G. A. Cooper, and A. Kelly, Single and multiple fracture. In The Properties
of Fiber Composites. (Surrey, UK: IPC Science and Technology Press, National Phys-
ical Laboratory, 1971), pp. 15–26.
3. K. W. Garrett and J. E. Bailey, Multiple transverse fracture in 90 degrees cross-ply
laminates of a glass fiber-reinforced polyester. J Mater Sci, 12:1 (1977), 157–68.
4. J. Aveston and A. Kelly, Theory of multiple fracture of fibrous composites. J Mater Sci,
8:3 (1973), 352–62.
5. Z. Hashin, Analysis of cracked laminates: a variational approach. Mech Mater, 4:2
(1985), 121–36.
6. J. Varna and L. A. Berglund, Multiple transverse cracking and stiffness reduction in
cross-ply laminates. J Compos Tech Res, 13:2 (1991), 97–106.
7. N. V. Akshantala and R. Talreja, A mechanistic model for fatigue damage evolution in
composite laminates. Mech Mater, 29 (1998), 123–40.
8. L. N. McCartney, Model to predict effects of triaxial loading on ply cracking in general
symmetric laminates. Compos Sci Technol, 60 (2000), 2255–79 (see Errata in Compos Sci
Technol, 62:9 (2002), 1273–4).
9. L. M. Kachanov, On the creep rupture time. Izv Akad Nauk SSR, Otd Tekhn Nauk,
8 (1958), 26–31.
10. J. Lemaitre and J. L. Chaboche, Mechanique des Materiaux Solide. (Paris: Dunod, 1985).
11. D. Krajcinovic, Continuous damage mechanics. Appl Mech Rev, 37 (1984), 1–5.
12. R. Talreja, A continuum-mechanics characterization of damage in composite-
materials. Proc R Soc London A, 399:1817 (1985), 195–216.
8 Durability assessment of composite structures

13. R. Talreja, Transverse cracking and stiffness reduction in composite laminates.


J Compos Mater, 19 (1985), 355–75.
14. C. K. H. Dharan, Fatigue failure mechanisms in a unidirectionally reinforced compos-
ite material. In Fatigue in Composite Materials, ASTM STP 569. (Philadelphia, PA:
ASTM, 1975), pp. 171–88.
15. R. Talreja, Fatigue of composite materials: damage mechanisms and fatigue-life dia-
grams. Proc R Soc London A, 378 (1981), 461–75.
16. E. K. Gamstedt and R. Talreja, Fatigue damage mechanisms in unidirectional carbon
fibre-reinforced plastics. J Mater Sci, 34 (1999), 2535–46.
17. N. V. Akshantala and R. Talreja, A micromechanics based model for predicting fatigue
life of composite laminates. Mater Sci Eng A, 285 (2000), 303–13.
18. M. Quaresimin, L. Susmel, and R. Talreja, Fatigue behaviour and life assessment of
composite laminates under multiaxial loadings. Int J Fatigue, 32 (2010), 2–16.
2 Review of mechanics of
composite materials

In this chapter the fundamental aspects of elasticity, strength, and fracture of


composite solids are reviewed. Although this information is available in numerous
texts, more comprehensively and in greater detail than here, a brief exposition is
provided for convenient reference. For further in-depth treatment, the reader may
consult, e.g., [1–5] for theory of elasticity and continuum mechanics, [6–12] for
mechanics of composite materials, and [13–17] for fracture mechanics.

2.1 Equations of elasticity

2.1.1 Strain–displacement relations


Figure 2.1 illustrates the initial and deformed configurations of a body whose
representative material point P is described with respect to a fixed rectangular
Cartesian frame by coordinates Xj and xi respectively, j, i = 1, 2, 3. The compon-
ents of displacement of the point are given by
ui ¼ x i Xj dij ; ð2:1Þ

where Xj are the coordinates of the material point in the initial undeformed
configuration, xi are the coordinates of the material point in the final deformed
configuration, and dij is the Kronecker delta. The Lagrangian description of
displacement at time t is expressed in terms of the Xj coordinates as
ui ¼ xi ðX1 ; X2 ; X3 ; tÞ Xj dij : ð2:2Þ
The components of the Green–Lagrange strain tensor are given by
1 
Eij ¼ ui;j þ uj;i þ ui;k uj;k ; ð2:3Þ
2
@ui
where ui;j ¼ ; etc., and repeated indices imply summation.
@X
  j
When ui;j   1, Eij reduces to the infinitesimal strain tensor eij given by
 
1 @ui @uj 1 
eij ¼ þ  ui;j þ uj;i : ð2:4Þ
2 @xj @xi 2
10 Review of mechanics of composite materials

Deformed configuration

P⬘

Initial configuration
u

X3
P x

O X2

X1

Figure 2.1. Initial and deformed geometry of a continuum body.

From Eq. (2.4) it is seen that the strain tensor is symmetric. Thus, there are six
independent strain components, which in the infinitesimal version are three normal
strains (e11, e22, and e33), and three shear strains (e12 = e21, e23 = e32, and e13 = e31).
To ensure single-valued displacements ui, the strain components eij cannot be
assigned arbitrarily but must satisfy certain integrability or compatibility condi-
tions, given by
eij;kl þ ekl;ij eik;jl ejl;ik ¼ 0: ð2:5Þ
Of the 81 equations included in Eq. (2.5), only six are independent. The remainder
are either identities or repetitions due to symmetry of eij. For the special case of
plane stress conditions, the only surviving compatibility equation is
e11;22 þ e22;11 2e12;12 ¼ 0: ð2:6Þ

2.1.2 Conservation of linear and angular momenta


In general, the forces exerted on a continuum body are body forces and surface forces.
Body forces, such as gravitational and magnetic forces, act on all particles within the
volume of the body and are described in terms of force intensity per unit mass or per
unit volume, while surface forces are contact forces that act across an internal surface
or an external (bounding) surface. The continuum description of surface forces is
given by the traction vector t acting on a surface element dS with a unit normal n (see
Figure 2.2(a)). Let dP be the total force exerted on dS by the material points on the
side of dS toward which n is pointing. The traction vector t is then defined as
dP
t ¼ lim : ð2:7Þ
dS!0 dS

At an internal point P there are infinitely many surface elements, each with a different
unit normal vector. According to the Cauchy theorem a traction vector on any of these
2.1 Equations of elasticity 11

(a) (b) s33


dS
t s32
s23
s31
n
s13
P s22

s12 s21

s11

Figure 2.2. (a) Traction vector; (b) a volume element with components of stress tensor.

planes can be expressed in terms of the traction vectors on three orthogonal planes
passing through the point P. In a Cartesian reference frame, the three planes are chosen
parallel to the coordinate planes and the resultant traction vectors on these planes are
decomposed along the three coordinate axes. These 3  3 = 9 components taken
together form the components of the second rank stress tensor associated with the
considered point P. They are indicated in Figure 2.2(b) where their positive directions
are shown. In index notation, they are denoted by sij, where the first index refers to the
direction of the unit normal on the surface (the face of the cube in Figure 2.2(b)) and the
second index stands for the direction of the resolved traction component. The stress
components with two equal indices, e.g., s11, are called normal stresses while those with
unequal indices, e.g., s23, are termed shear stresses.
The traction vector components are related to the stress tensor components by
the following equation
ti ¼ sij nj ; ð2:8Þ
where nj are components of the normal vector associated with the traction vector.
The conservation of linear momentum at a material point inside the continuum
body gives the following relation
sji; j þ fi ¼ r€
ui ; ð2:9Þ
where fi are components of the body force vector, and r is the mass density. For
quasi-static problems the right-hand side of Eq. (2.9) vanishes, and if the body
forces are neglected, the equations of equilibrium reduce to
sji; j ¼ 0: ð2:10Þ
When there are no body moments, the conservation of angular momentum results
in the symmetry of the stress tensor, i.e.,
sij ¼ sji : ð2:11Þ

2.1.3 Constitutive relations


For an elastic material there exists a positive-definite, single-valued, potential
function of strains ekl, defined as
12 Review of mechanics of composite materials

ð ekl
U¼ sij deij : ð2:12Þ
0

This function is termed as the “strain energy density.” U is independent of the


loading path and thus a function of final strains only. Differentiating Eq. (2.12)
with respect to the strains, the stress tensor can be written as
@U
sij ¼ : ð2:13Þ
@eij
If we consider a linear elastic material, then U can be written as a quadratic
function in ekl
1
U ðekl Þ ¼ Cijkl eij ekl ; ð2:14Þ
2
where Cijkl is a fourth-order tensor of material stiffness coefficients known as
the stiffness tensor. Using Eqs. (2.13) and (2.14), one obtains the generalized
Hooke’s law
sij ¼ Cijkl ekl : ð2:15Þ
A potential function of stresses known as the complementary energy density is
defined as

U  sij ¼ sij eij U:


 
ð2:16Þ

Differentiation of Eq. (2.16) with respect to stress tensor yields the relation
@U 
eij ¼ : ð2:17Þ
@sij
Analogous to Eq. (2.14), U* can also be represented as a quadratic function as
  1
U  sij ¼ Sijkl sij skl ; ð2:18Þ
2
where Sijkl are components of the compliance tensor. Using Eqs. (2.17) and (2.18),
one obtains the inverse constitutive law
eij ¼ Sijkl skl : ð2:19Þ
In all, the stiffness matrix Cijkl has 81 coefficients. However, not all of these
coefficients are independent. Note first that the symmetry of the strain compon-
ents (ekl = elk) leads to Cijkl = Cijlk, which reduces the number of coefficients from
81 to 54. Similarly, the symmetry of the stress tensor further reduces the number of
these coefficients to 36. Finally, differentiating Eq. (2.14) twice with respect to
strains, one obtains

@2U
Cijkl ¼ : ð2:20Þ
@eij @ekl
2.1 Equations of elasticity 13

Since the order of differentiation in the above equation is arbitrary, one infers
that
Cijkl ¼ Cklij ; ð2:21Þ
which reduces the number of independent material coefficients to 21.
The coefficient matrix Cijkl is expressed in compact form by using the Voigt
notation, in which stress and strain tensor components are denoted using a single
subscript, whereas two subscripts are used to denote the stiffness tensor. With this,
the constitutive relation, Eq. (2.15), can be written as sp = Cpqeq; p, q = 1,2, . . ., 6,
or in expanded matrix form as
8 9 2 38 9
>
> s1 >
> C11 C12 C13 C14 C15 C16 > > e1 >>
> s2 >
>
> > 6 C21 C22 C23 C24 C25 C26 7>
> > e2 >
> >
>
> > 6 7 < >
>
s3 C C C C C C e
< = 6 =
31 32 33 34 35 36 7 3
7
¼66 C41 C42 C43 C44 C45 C46 7> e4 >; ð2:22Þ
>
> s4 >
> 6 7> >
s >> 4 C51 C52 C53 C54 C55 C56 5> e >
> > > >
: 5> : 5>
>
> > >
> >
s6 C61 C62 C63 C64 C65 C66 e6
; ;

where Cpq = Cqp, and

s1 ¼ s11 ; s2 ¼ s22 ; s3 ¼ s33 ; s4 ¼ s23 ; s5 ¼ s31 ; s6 ¼ s12


e1 ¼ e11 ; e2 ¼ e22 ; e3 ¼ e33 ; e4 ¼ 2e23 ; e5 ¼ 2e31 ; e6 ¼ 2e12 ð2:23Þ
C11 ¼ C1111 ; C22 ¼ C2222 ; : : : ; etc:

These constitutive relationships are for an anisotropic material. If material sym-


metry exists, then further reduction occurs in the number of independent coeffi-
cients of the stiffness matrix. It should be noted that the stiffness matrix in the
Voigt notation does not follow the transformation rule for tensors. The fourth-
order stiffness tensor Cijkl transforms as
0
Cijkl ¼ ‘ip ‘jq ‘kr ‘ls Cpqrs ; ð2:24Þ

where ‘ij is the matrix of direction cosines associated with coordinate trans-
formation from one coordinate system (x1, x2, x3) to another ðx01 ; x02 ; x03 Þ.
A material with one plane of symmetry is called monoclinic, and if this plane is
parallel to the x1–x2 plane then it can be shown that the constitutive relation is
given by
8 9 2 38 9
>
> s1 >
> C11 C12 C13 0 0 C16 > > e1 >
> >
s C C C 0 0 C e
> > > 6 >
2 21 22 23 26 2>
>
> > 7>> >
< >
> = 6 7>< >
s3 C31 C32 C33 0 0 C36 7 e3
6 7 =
¼6 : ð2:25Þ
> 6 0
> s4 > 0 0 C44 C45 0 7 > e4 >
> 6
7> >
> s5 >
> 4 0 0 0 C54 C55 0 5> > e5 >
>
> > > >
>
> > > >
s6 C61 C62 C63 0 0 C66 e6
: ; : ;

Here, the stiffness matrix has 13 independent material coefficients. If a mater-


ial has two mutually orthogonal planes of symmetry, then the plane
14 Review of mechanics of composite materials

orthogonal to these planes is also a plane of symmetry. In this case, the


material symmetry is described as orthotropic, and the number of independent
constants in the stiffness matrix reduces to nine. The stress–strain relations
when the symmetry planes are parallel to the three coordinate planes take the
following form

8 9 2 38 9
>
> s1 > C11 C12 C13 0 0 0 > e1 >
> > > >
s 6 C21 C22 C23 0 0 0 7 e
> > >
2 2>
>
> >
> >
> >
< >
> = 6 < >
7>
s3 6 C31 C32 C33 0 0 0 7 e3
7 =
¼6 : ð2:26Þ
> s4 >> 6 0 0 0 C44 0 0 7> e4 >
> 6 7
> >
4 0 0 0 0
> s5 > C55 0 5> e >
> > > >
: 5>
>
> > > >
> >
s6 0 0 0 0 0 C66 e6
: ; ;

In terms of the engineering elastic constants the inverse strain–stress relations for
the orthotropic case become as follows

1 n21 n31
2 3
0 0 0 7
6
6 E1 E2 E3 7
6 n12 1 n32 7
8 9 6 0 0 0 7 8 9
>
> e1 >
>
6
6 E1 E2 E3 7 > s1 >
7> >
>
> e2 > 6
>
> n13 n23 1 7>
> s2 >
> >
>
0 0 0 7
>
> >
< > = 6 >
e3 E1 E2 E3 7 s3 =
6 <
¼6 7> s4 >;
7 ð2:27Þ
e4 > 6
6 0 1
0 0 0 0 7
>
> > > >
e > 6 > s5 >
> > 7>
: 5>
>
>
>
>
; 6 G23 7>> >
>
>
e6 6
6 0 1 7: s6 ;
0 0 0 0 7
7
6
6 G31 7
4 1 5
0 0 0 0 0
G12

where E1, E2, E3 are Young’s moduli in the three material symmetry directions
(x1, x2, x3) respectively, nij ; i 6¼ j; are the six Poisson’s ratios defined in the
conventional way, e.g., n12 = e2 / e1 with s1 applied, and G23, G31, and G12 are
shear moduli in the x2–x3, x1–x3, and x1–x2 planes, respectively. The compliance
matrix in Eq. (2.27), being the inverse of a symmetric matrix, is also symmetric.
From this symmetry follows the “reciprocal” relationship,
nij nji
¼ ðno sum on i; jÞ; ð2:28Þ
E i Ej
which can be used to eliminate three of the six Poisson’s ratios.
If the material is isotropic in a plane, i.e., with same elastic properties in all
directions in the plane, it is called transversely isotropic. Let the x2–x3 plane be the
E2
plane of isotropy, i.e., E3 ¼ E2 ; n31 ¼ n12 ; G31 ¼ G12 ; G23 ¼ : The com-
2ð1 þ n23 Þ
pliance tensor is then given by
2.1 Equations of elasticity 15

1 n12 n12
2 3
0 0 0 7
6 E1 E1 E1
n12 1 n23
2 3 6 7
S11 S12 S12 0 0 0 6
6 0 0 0 7
7
6 S12 S22 S23 0 0 0 7 6 E1 E2 E2 7
6
6 S12 S23 S22 0 0 0
7 6 n12 n23 1 7
0 0 0 7:
7 6 7
½S Š ¼ 6 7¼6
E1 E2 E2
6 0 0 0 2ðS22 S23 Þ 0 0 7 6 7
6
4 0 0 0 0 S66 0
7 6
6 0 E2 7
5 0 0 0 0 7
2ð1 þ n23 Þ
6 7
0 0 0 0 0 S66 6 7
1
4 0 0 0 0 0 5
6 7
G12
1
0 0 0 0 0 G12

ð2:29Þ
As seen above, a transversely isotropic material has five independent stiffness
coefficients, viz. E1, E2, n23, n12, and G12. For a completely isotropic material there
are only two independent material coefficients, namely the Young’s modulus (E)
and Poisson’s ratio (n) or, alternatively, the Lame constants (l and m). The
constitutive relations can now be written as

sij ¼ lekk dij þ 2meij ; ð2:30Þ

where dij is the Kronecker delta. Alternatively,

1 
eij ¼ ð1 þ nÞsij nskk dij : ð2:31Þ
E

2.1.4 Equations of motion


The equations governing the motion of a deformable body can be obtained by
combining kinematic relations, Eq. (2.4), equilibrium equations, Eq. (2.10), and
the constitutive relations, Eq. (2.15). For the particular case of linear elastic
isotropic materials, they can be written as
ðl þ mÞuj; ji þ mui; jj þ fi ¼ r€
ui : ð2:32Þ
These equations are known as Navier’s equations. The displacement field obtained
from these equations is unique and results into the determination of strains and
stresses by use of kinematic and constitutive relations.

2.1.5 Energy principles


Energy principles for a continuum body allow formulating the relationships
between stresses, strains or deformations, displacements, material properties,
and external effects in the form of energy or work done by internal and external
forces. They are also useful for obtaining approximate solutions of complex
boundary value problems, e.g., finite element methods. Detailed treatment of these
concepts can be found in [18–20].
16 Review of mechanics of composite materials

ni

tj

Su
fi

V
ui St

Figure 2.3. A continuum body loaded with body forces inside its volume, and traction and
displacement on the boundary.

Principle of virtual work


In the context of an elastic boundary value problem, consider a solid continuum
body (Figure 2.3), occupying a volume V and bounded by surface S = St þ Su, to
be in static equilibrium under prescribed body forces fi over volume V, surface
tractions ti on St, and displacements ui over remaining portion of the boundary Su.
For a statically admissible stress field s ~ij; j ¼ 0 in V, and ~ti ¼ s
~ij (such that s ~ij nj
on St) and a kinematically admissible displacement field u^i (such that
^eij ¼ 12 u^i; j þ u^j;i ), the principle of virtual work states

ð ð ð
~ti u^i dS þ fi u^i dV ¼ ~ij^eij dV:
s ð2:33Þ
S V V

~ij are
It should be noted that the displacement field u^i and the stress field s
completely independent of each other.

Principle of minimum potential energy


For a kinematically admissible displacement field u^i , the potential energy of a
linear elastic continuum body under the action of conservative forces fi and
prescribed surface tractions ti on St is defined as
ð ð ð
1
ðu^i Þ ¼ ^ij^eij dV
s ti u^i dS fi u^i dV: ð2:34Þ
2 V S V

The principle of minimum potential energy states that among all the kinematically
admissible displacement fields the actual displacement field minimizes the poten-
tial energy. Thus, if ui represents the actual displacement field, then

ðu^i Þ  ðui Þ : ð2:35Þ


2.2 Micromechanics 17

P, D P
Õ*

Figure 2.4. A typical load–displacement diagram.

Principle of minimum complementary energy


^ij , the complementary potential energy of a
For a statically admissible stress field s
linear elastic body is defined as
ð ð

  1
 s ^ij ¼ ^ij^eij dV
s ^ti ui dS; ð2:36Þ
2 V S

where ^ti ¼ s
^ij nj is the reaction on Su. The principle of minimum complementary
energy states that among all the statically admissible stress fields the actual stress
field minimizes the complementary potential energy. Thus, if ui represents the
actual displacement field, then

 s^ij   sij :
   
ð2:37Þ

For actual stress, strain, and displacement fields, addition of Eqs. (2.34) and (2.36) yields
ð ð ð

 
ðui Þ þ  sij ¼ sij eij dV ti ui dS fi ui dV: ð2:38Þ
V S V

The right-hand side of Eq. (2.38) vanishes by virtue of the principle of virtual
work. Hence,

ðui Þ ¼  sij :
 
ð2:39Þ

Using Eqs. (2.35), (2.37), and (2.39), we obtain the lower and upper bounds to the
potential energy of a continuum body

ðu^i Þ  ðui Þ ¼  sij   s


   
^ij : ð2:40Þ

For the purpose of illustration, the potential and complementary energies for a
typical load–displacement response are shown in Figure 2.4.

2.2 Micromechanics

Micromechanics is a well-developed advanced field that treats the response of a


heterogeneous solid based on the behavior of its constituents and their geometrical
configurations. For a detailed exposition the reader may refer to, e.g., [8]. Here a
brief summary of simple micromechanics estimates of the linear elastic properties
18 Review of mechanics of composite materials

(a) (b) (c)

A unidirectional lamina Lamination Laminate

Figure 2.5. Stacking of a number of laminae makes up a laminate.

of a unidirectional fiber-reinforced composite is provided. These estimates are


useful in selecting fibers and matrix materials and their volume fractions. In many
structural applications a unidirectional composite, fabricated as a thin layer,
called lamina or ply, is used as a basic unit and a laminate is constructed by
stacking these layers as illustrated in Figure 2.5.

2.2.1 Stiffness properties of a unidirectional lamina


Linear elastic properties of a lamina can be referred to a coordinate system (x1, x2,
x3) where the x1-axis is aligned with fibers, x2-axis is transverse to fibers in
the plane of the lamina, and the x3-axis is normal to the plane of lamina
(see Figure 2.6). Noting that the lamina has orthotropic symmetry, the nine
independent elastic constants, as described in Section 2.1.3 above, in this reference
system are the three Young’s moduli (E1, E2, E3), the three Poisson’s ratios (n12,
n13, n23), and the three shear moduli (G12, G13, G23). For a subset of these constants
that represents in-plane properties, i.e., E1, E2, n12, n21, and G12, in the x1–x2 plane,
the following expressions hold
E1 ¼ Ef V f þ E m V m ; ð2:41Þ
n12 ¼ nf Vf þ nm Vm ; ð2:42Þ
1 Vf Vm
¼ þ ; ð2:43Þ
E2 Ef Em
1 Vf Vm
¼ þ ; ð2:44Þ
G12 Gf Gm
where E, n, G, and V stand for the Young’s modulus, Poisson’s ratio, shear
modulus, and volume fraction, respectively, with the subscripts f and m indicating
fibers and matrix, respectively. The minor Poisson’s ratio n21 can be estimated
using the reciprocal relationship n21 = n12 (E2/E1).
Equations (2.41) and (2.42) have the form of the familiar rule of mixtures and
Eqs. (2.43) and (2.44) follow that rule for the inverse of the respective properties.
2.2 Micromechanics 19

x1
y

x 3, z

x2

Figure 2.6. Coordinate systems for a unidirectional ply. The material system is denoted by
x1, x2, x3; while the laminate system is denoted by x, y, z.

The first two expressions predict the experimental properties usually well while the
third and fourth expressions are found to be less accurate. Halpin and Kardos [21]
and Halpin and Tsai [22] proposed semi-empirical relationships based on numer-
ical computations. These relations can be used in place of Eqs. (2.43) and (2.44),
and are together expressed as
p 1 þ xVf
¼ ; ð2:45Þ
pm 1 Vf
where
pf
1
pm
 ¼ pf : ð2:46Þ
þx
pm
Here p represents E2 or G12, and pf and pm are the corresponding moduli for fiber
and matrix, respectively. The fitting parameter x needs to be determined by
comparing predictions with experimental data.
More advanced micromechanics approaches, such as Hashin–Shtrikman vari-
ational bounds [23–29], Mori–Tanaka model [30], composite sphere and cylinder
assemblage model [31, 32], self-consistent method [33], method of cells [34–36], etc.
have also been developed in the past four decades. Interested readers are referred to
texts on micromechanics, e.g., [8, 37, 38], for detailed treatment of these approaches.

2.2.2 Thermal properties of a unidirectional lamina


Simple micromechanics estimates for the linear coefficient of thermal expansion of
a lamina can be obtained in the same way as the linear elastic properties. The
expressions obtained are as follows
1
a1 ¼ ðaf Ef Vf þ am Em Vm Þ;
E1 ð2:47Þ
a2 ¼ ð1 þ nf Þaf Vf þ ð1 þ nm Þam Vm a1 n12 ;
20 Review of mechanics of composite materials

where a1 and a2 are the thermal expansion coefficients in the fiber and transverse
directions, respectively, and E1 and n12 are given by Eqs. (2.41) and (2.42).

2.2.3 Constitutive equations for a lamina


A lamina is thin compared to other dimensions of the entire laminate. Therefore,
the lamina can be assumed to be in a state of generalized plane stress. Conse-
quently, all the through-thickness stress components are zero, i.e., s4 = s5 =
s6 = 0. In such a case, the constitutive relation for an individual lamina referred
to the three axes of symmetry can be written in Voigt notation as
8 9 2 38 9
< s1 = Q11 Q12 0 < e1 =
s ¼ 4 Q12 Q22 0 5 e2 ; ð2:48Þ
: 2;
s6 0 0 Q66 e6
: ;

with
E1 E2 n12 E2 n21 E1
Q11 ¼ ; Q22 ¼ ; Q12 ¼ ¼ ; Q66 ¼ G12 : ð2:49Þ
1 n12 n21 1 n12 n21 1 n12 n21 1 n12 n21
The inverse constitutive relation for the lamina is given by
1 n21
2 3
8 9 6 E 0 8 9
< e1 = 6 n1 E2 7< s1 =
7
12 1
eij ¼ Sijkl skl ) e2 ¼ 6 0 7 s2 : ð2:50Þ
6 7
: ; 6 E1 E 2
e6 1 5 s6
7: ;
4
0 0
G12
The above constitutive relations are written in the lamina coordinate system (i.e.,
with x1 along the fiber direction, x2 normal to the fiber direction, and x3 along the
lamina thickness). The constitutive relation for the lamina in another coordinate
system (x–y–z), which, for instance, could be aligned with the coordinate system
chosen for the laminate, is

Q11 Q12 Q16 < exx =


8 9 2 38 9
< sxx =
syy ¼ 4 Q12 Q22 Q26 5 eyy ; ð2:51Þ
:
sxy
;  
Q16 Q26 Q66  :
2e xy
;

where Qij are known as reduced stiffness coefficients. These are related to Qij ,
defined by Eq. (2.49), by the transformation rules for stresses and strains. Thus,
8 9 2 38 9 8 9
< s1 = m2 n2 2mn < sxx = < sxx =
s2 ¼ 4 n2 m2 2mn 5 syy ¼ ½T Š syy ð2:52Þ
2
s6 mn mn m n2 sxy sxy
: ; : ; : ;

8 9 2 38 9 8 9
< e1 = m2 n2 mn < exx =
T < exx =
1
e ¼ n2 m2 mn 5 eyy ¼ ½T Š eyy ð2:53Þ
: 2;
4
e6 2 2 :e ; exy
2mn 2mn m n
: ;
xy
2.2 Micromechanics 21

where m = cos y, n = sin y, where y is the angle between the x- and x1-axes (Figure
2.6). Then by inverting Eqs. (2.52) and (2.53), substituting these in Eq. (2.51), and
on using Eq. (2.48) one obtains

1
Q ¼ ½T Š 1 ½QŠ ½T ŠT
 
: ð2:54Þ

[T] 1 is simply given by changing y to –y, i.e., [T(y)] 1


= [T( y)]. Expanding the
above relation, we have
Q11 ¼ Q11 m4 þ 2ðQ12 þ 2Q66 Þm2 n2 þ Q22 n4 ;
Q22 ¼ Q11 n4 þ 2ðQ12 þ 2Q66 Þm2 n2 þ Q22 m4 ;
Q12 ¼ ðQ11 þ Q22 4Q66 Þm2 n2 þ Q12 m4 þ n4 ;
 
ð2:55Þ
Q16 ¼ ðQ11 Q12 2Q66 Þm3 n þ ðQ12 Q22 þ 2Q66 Þmn3 ;
Q26 ¼ ðQ11 Q12 2Q66 Þmn3 þ ðQ12 Q22 þ 2Q66 Þm3 n;
Q66 ¼ ðQ11 þ Q22 2Q66 Þm2 n2 þ Q66 m4 þ n4 :
 
2Q12
The transformation rules described above enable us to express engineering moduli
for the lamina referred to arbitrary in-plane axes (x–y) in terms of moduli in the
principal (x1–x2) directions as
m 4 n4
 
1 1 2n12
¼ þ þ m 2 n2 ;
E x E 1 E2 G12 E1
n4 m4
 
1 1 2n12
¼ þ þ m 2 n2 ;
Ey E1 E2 G12 E1
  ð2:56Þ
nxy n12 1 þ 2n12 1 1
¼ þ m 2 n2 ;
Ex E1 E1 E2 G12
 
1 1 1 þ n 12 1 þ n21 1
¼ þ 4m2 n2 þ :
Gxy G12 E1 E2 G12
To account for thermal stresses, we need to modify strains in Eq. (2.51) to include
thermal strains, as
9 8 0 9 8 th 9
< exx > < exx >
8
< exx = > = > =
0
eyy ¼ eyy þ eth
yy ; ð2:57Þ
2exy : 2e0 > : 2eth >
: ; > ; > ;
xy xy

where the superscripts 0 and th denote mechanical and thermal strains, respectively, with
8 9
th 8 th 9
< exx >
8 9
> = < e1 = < a1 =
eth ¼ ½ T Š e th ¼ ½T Š a T : ð2:58Þ
yy e
: 2th ; e
: 2;
: 2eth > 0
>
xy
; 2e 12

2.2.4 Strength of a unidirectional lamina


Phenomenological failure (strength) criteria that use experimental data to deter-
mine material constants have been proposed for composite materials along the
22 Review of mechanics of composite materials

lines of those used for metals such as the von Mises yield criterion. Failure
mechanisms in composite materials are, however, significantly more complex,
resulting in a large number of criteria. Here, some common criteria will be stated
for reference; the interested reader is encouraged to consult [39] for more in-depth
treatment.
For a unidirectional fiber-reinforced lamina, the five basic strength parameters
under in-plane loading are as follows:
X = ultimate tensile strength in the fiber direction
X0 = ultimate compressive strength in the fiber direction
Y = ultimate tensile strength transverse to fibers
Y0 = ultimate compressive strength transverse to fibers
S = ultimate shear strength in the lamina plane.
These parameters are obtained by experimental testing. See, e.g., [39, 40] for
further details.

Maximum stress theory


According to this theory, a lamina fails if
X s1 > 0;

s1 ¼
X0 s1 < 0;
Y s2 > 0;

ð2:59Þ
s2 ¼
Y0 s2 < 0;
js6 j ¼ S:

For combined loading, theoretical predictions of the theory are inaccurate because
the maximum stress criterion does not account for stress interactions. For an off-
axis normal loading, i.e., loading axis inclined to fibers, this theory can be applied
by transforming the stresses to the principal material directions and then using the
criteria in Eq. (2.59).

Maximum strain theory


This theory states that failure occurs when
Xe e1 > 0;

e1 ¼
X0 e e1 < 0;
Ye e2 > 0;

ð2:60Þ
e2 ¼
Y0e e2 < 0;
je6 j ¼ Se ;

where Xe ¼ X=E1 ; Xe0 ¼ X0 =E1 ; Ye ¼ Y=E2 ; Ye0 ¼ Y 0 =E2 ; and Se ¼ S=G12 are the
ultimate failure strains analogous to the stress-based parameters mentioned above.
2.2 Micromechanics 23

Distortional energy (Tsai–Hill) criterion


This criterion is based on the distortional energy failure (yield) theory of
von Mises. Hill [41] further developed this yield criterion for anisotropic
materials and Azzi and Tsai [42] modified it to describe failure of a composite
lamina as follows

s21 s1 s2 s22 s26


þ 2 þ 2 ¼ 1; ð2:61Þ
X2 X2 Y S
where s1 and s2 are the tensile normal stresses along fibers and normal to
fibers, respectively, and s6 is the in-plane shear stress. When the normal
stresses are compressive, the compressive strength values in Eq. (2.61) are to
be used.

Tsai–Wu criterion
A polynomial function of stress components can be formulated with the multiply-
ing terms of the polynomial expressing strength properties. Restricted to quadratic
terms of in-plane stress components, such a function is known as the Tsai–Wu
criterion [43] and can be expressed as

F1 s1 þ F2 s2 þ F11 s21 þ F22 s22 þ F66 s26 þ 2F12 s1 s2 ¼ 1: ð2:62Þ

The product terms s1s6 and s2s6 are not present in Eq. (2.62) because the
multiplying coefficients to these terms can be shown to vanish. Also, the linear
term in s6 is absent because of the shear strength being independent of the sign
of shear stress, which renders the coefficient of this term to be zero.
The six material constants in the Tsai–Wu criterion require two tests (tension
and compression) in the fiber direction, two similar tests normal to fibers, an in-
plane shear test, and a biaxial normal load test.

Hashin’s criterion
Hashin [44] formulated three-dimensional failure criteria for unidirectional fiber
composites in terms of quadratic stress polynomials. The terms used in the
polynomials were functions of the stress invariants for transversely isotropic
symmetry. Thus the cross-sectional plane of a unidirectional fiber composite was
assumed as an isotropic plane. For relatively thick layers this may be a good
assumption.
A unidirectional fiber composite was assumed to fail in one of four possible
separate modes: tensile fiber mode (s1 > 0), compressive fiber mode (s1 < 0),
tensile matrix mode (s2 + s3 > 0), and compressive matrix mode (s2 + s3 < 0)
For a thin unidirectional fiber composite layer (lamina), the four failure criteria
are given by
24 Review of mechanics of composite materials

s
2 s
2
2 6
þ ¼ 1; s1 > 0;
X S
s1 ¼ X0 ; s1 < 0;
s
2 s
2
2
þ
6
¼ 1; s2 > 0; ð2:63Þ
Y S
"  #
s
2
2 Y0 2 s2 s6
2
0
þ 0
1 0þ ¼ 1; s2 < 0;
2S 2S Y S

where S0 is the strength in transverse shear, while S here is the same in axial shear.
The difference between the two shear strengths is not fully unambiguous.
Over the years, a wide variety of failure criteria have been proposed. There is no
single failure theory that seems to capture all the complexities of composite failure.
A world-wide failure exercise was conducted to evaluate applicability of most
theories by comparing their predictions with test data [45].

2.3 Analysis of laminates

Laminates used in most engineering applications are fabricated by stacking plies in


different orientations. An example of a laminate with layup [0/90/45]s is shown in
Figure 2.7. A commonly used method of determining stresses and strains for such
laminates is based on the classical laminate plate theory (CLPT). More advanced
theories are treated in [9, 46]. The geometrical conditions needed for the application
of the CLPT are: (a) the individual plies are of uniform thickness, (b) they are
perfectly bonded to their neighboring plies, and (c) the total thickness of the
laminate follows the so-called thin plate assumption, which states that the thickness
dimension is much smaller than other structural dimensions (width and length).
The kinematic assumptions of the CLPT derive from the Kirchhoff assumptions,
which state that (a) a line element normal to the mid-plane in the undeformed state

z0
z1
z2


90˚ x
45˚ q
45˚ zn
90˚

z
y

Figure 2.7. Stacking of unidirectional plies in different orientations to make a


multidirectional [0/90/45]s laminate. The subscript s denotes that the laminate is symmetric
about the mid-plane.
2.3 Analysis of laminates 25

of the plate remains straight and perpendicular to the mid-plane after deformation,
and (b) such a line element does not change its length when the plate deforms.

2.3.1 Strain–displacement relations


The Kirchoff assumptions stated above lead to the x-, y-, and z-displacements u, v,
and w, respectively, in the coordinate system shown in Figure 2.7 as follows
@w0 ðx; yÞ
uðx; y; zÞ ¼ u0 ðx; yÞ z ;
@x
@w0 ðx; yÞ ð2:64Þ
vðx; y; zÞ ¼ v0 ðx; yÞ z ;
@y
wðx; y; zÞ ¼ w0 ðx; yÞ;

where (u0, v0, w0) are the displacements of the laminate mid-plane. The corres-
ponding strain–displacement relations are given by

@u @u0 @ 2 w0
exx ¼ ¼ z 2 ;
@x @x @x
@v @v0 @ 2 w0
eyy ¼ ¼ z 2 ;
@y @y @y
@w
ezz ¼ ¼ 0;
@z
ð2:65Þ
@ 2 w0
   
1 @u @v 1 @u0 @v0
exy ¼ þ ¼ þ z ;
2 @y @x 2 @y @x @x@y
 
1 @u @w
exz ¼ þ ¼ 0;
2 @z @x
 
1 @v @w
eyz ¼ þ ¼ 0:
2 @z @y

The nonzero equations can be written in the following form


8 9 8 0 9
< exx >
8 9
< exx = > = < kxx =
eyy ¼ e0yy þ z kyy ; ð2:66Þ
:g ; > : g0 > k
: ;
xy xy
;
xy

where e0xx ; e0yy ; g0xy are the mid-plane strains and kxx ; kyy ; kxy are the laminate
 

curvatures, given by

@ 2 w0 >
8 9 8 9
> @u 0 > >
> >
8 9 > @x2 >
> > >
@x
> > >
0
< exx >
> > 8 9 > >
kxx = >
> > > >
2
> = > > >
0
< @v0 = < < @ w 0
=
eyy ¼ and kyy ¼ : ð2:67Þ
@y @y2 >
: g0 > k
> ; > > : ; >
xy
> > > >
xy
>
> @u0 @v0 > > >
> >
@ 2 w0 >
>
þ
> > >
2
> > > >
@y @x
: ; >
: >
;
@x@y
26 Review of mechanics of composite materials

2.3.2 Constitutive relationships for the laminate


Using the lamina constitutive relations described earlier, the constitutive equation
for the kth (k = 1, 2, . . .) layer of the laminate can be written as
 ðkÞ
fsgðkÞ ¼ Q fegðkÞ : ð2:68Þ

In the above equation, the square bracket represents a 33 matrix and the curly
bracket is for a 31 vector. The strains in the kth ply are given by

fegðkÞ ¼ e0 þ zfkg:

ð2:69Þ

The thermal strains can be added to these strains, such that

fegðkÞ ¼ e0 þ zfkg

fagk T: ð2:70Þ

The kth ply stresses on using Eq. (2.68) can now be written as
 ðkÞ  0  ðkÞ
fsgðkÞ ¼ Q fagk T þ z Q fkg:

e ð2:71Þ

At the laminate level the force and moment resultants are defined as
8 9 8 9
< Nxx >
> = ð h=2 >< sxx >
=
½NŠ ¼ Nyy ¼ syy dz;
> > h=2 > >
Nxy sxy
: ; : ;
8 9 8 9 ð2:72Þ
< Mxx >
> = ð h=2 > < sxx >
=
½MŠ ¼ Myy ¼ syy z dz:
> > h=2 > >
Mxy sxy
: ; : ;

In terms of ply stresses that generally vary from ply to ply, we have
8 9 8 9
< Nxx = X N ð zkþ1 < sxx =
Nyy ¼ syy dz; ð2:73Þ
; k¼1 zk :
Nxy sxy
: ;

which gives us
8 9 8 th 9
< Nxx >
> = > < Nxx >=
th
Nyy þ Nyy
> >
; > : th >
Nxy Nxy
: ;
38 0 9
Q11 Q12 Q16 > Q11 Q12 Q16 >
38 9
< exx >
2 2
X N ð zkþ1 = N ð zkþ1 < kxx >
=
6 7 0 X
4 Q12 Q22 Q26 5 eyy dz þ
6
¼ 4 Q12 Q22 Q26 5 kyy z dz;
7
k¼1 zk k¼1 zk
Q16 Q26 Q66 Q16 Q26 Q66
>
: 0 > > >
gxy kxy
; : ;

ð2:74Þ
where the force resultants due to thermal stresses are given by
2.3 Analysis of laminates 27

8 9
th
< Nxx > Q11 Q12 Q16 < ax T =
2 38 9
> = X N ð zkþ1
th
Nyy ¼ 4 Q12 Q22 Q26 5 ay T dz: ð2:75Þ
; k¼1 zk
: N th > Q16 Q26 Q66 0
> : ;
xy

The relations in Eq. (2.74) can be rewritten in more compact form by using
matrices [A] and [B] as follows
9 8 th 9 2 38 0 9 2
< Nxx > < exx >
8 38 9
< Nxx = > = A11 A12 A16 > = B11 B12 B16 < kxx =
th
Nyy þ Nyy ¼ 4 A12 A22 A26 5 e0yy þ 4 B12 B22 B26 5 kyy :
Nxy : N th > A16 A26 A66 : g0xy ; B16 B26 B66 kxy
: ; > ; > > : ;
xy

ð2:76Þ
Similarly,
8 9 8 9
< Mxx = XN ð zkþ1 < sxx =
Myy ¼ syy zdz; ð2:77Þ
; k¼1 zk :
Mxy sxy
: ;

i.e.,
8 9 8 th 9
< Mxx >
> = > < Mxx > =
th
Myy þ Myy
> ; >
> : th >
Mxy Mxy
: ;
38 0 9
Q11 Q12 Q16 > Q11 Q12 Q16 >
38 9
< exx >
2 2
N ð zkþ1 = N ð zkþ1 < kxx >
=
0
X 6  X
¼ 4 Q12 Q22 Q26 5 eyy zdz þ 4 Q12 Q22 Q26 5 kyy z2 dz;
7 6 7
k¼1 zk zk
Q16 Q26 Q66 k¼1
Q16 Q26 Q66
> > > >
0 kxy
g
: ; : ;
xy

ð2:78Þ
where
8 9
th
< Mxx > Q11 Q12 Q16 < ax T =
2 38 9
> = XN ð zkþ1
th
Myy ¼ 4 Q12 Q22 Q26 5 ay T zdz: ð2:79Þ
; k¼1 zk
: Mth > Q16 Q26 Q66 0
> : ;
xy

Introducing a new matrix [D], Eq. (2.78) can be rewritten as


9 8 th 9 2 38 0 9 2
< Mxx > < exx >
8 38 9
< Mxx = > = B11 B12 B16 > = D11 D12 D16 < kxx =
th
Myy þ Myy ¼ 4 B12 B22 B26 5 e0yy þ 4 D12 D22 D26 5 kyy :
Mxy : Mth > B16 B26 B66 : g0xy ; D16 D26 D66 kxy
: ; > ; > > : ;
xy

ð2:80Þ
The material coefficients (Aij, Bij, Dij) are known as the extensional stiffness, the
extension–bending coupling stiffness, and the bending stiffness coefficients, respect-
ively. These are given by
ðh
2
Qij 1; z; z2 dz;
   
Aij ; Bij ; Dij ¼ ð2:81Þ
h
2
28 Review of mechanics of composite materials

or
N
X
Aij ¼ Qij ðzkþ1 zk Þ ;
k¼1
N
1X
Qij z2kþ1 z2k ;
 
Bij ¼ ð2:82Þ
2 k¼1
N
1X
Qij z3kþ1 z3k :
 
Dij ¼
3 k¼1

The laminate constitutive relations can now be written in compact form as


   0 
fNg ½AŠ ½BŠ e
¼ ð2:83Þ
f Mg ½BŠ ½DŠ fkg
where {N}, {M} include thermal resultants.

2.3.3 Stresses and strains in a lamina within a laminate


The laminate constitutive relations in Eq. (2.83) can be reverted to yield the mid-
plane strains and curvatures in terms of the stress and moment resultants.
A partial inversion is first done by inverting Eq. (2.76) and substituting into
Eq. (2.80) to obtain
0  
A B
 
e N
¼ ; ð2:84Þ
M C D k

where

A ¼ A 1 ; B ¼ A 1 B; C ¼ BA 1
¼ ðB  ÞT ; D ¼ D BA 1 B; ð2:85Þ
and the brackets for matrix/vector representation have been dropped for conveni-
ence. Solving Eq. (2.84) for k and its substitution back gives
0  0
A B0
 
e N
¼ ; ð2:86Þ
k B0 D0 M

where

A0 ¼ A þ B ðD Þ 1 ðB ÞT ; B0 ¼ B ð D  Þ 1 ; D 0 ¼ ðD  Þ 1 : ð2:87Þ


Once mid-plane strains and curvatures are known, the strains and stresses in each
lamina can be determined using Eqs. (2.70) and (2.68), respectively.

2.3.4 Effect of layup configuration


The sequence of ply layup has a significant impact on the stiffness properties of the
designed laminate. Some interesting ply configurations are described below.
2.4 Linear elastic fracture mechanics 29

 Balanced laminate: If for each +y-ply, we have another identical ply of


the same thickness, but –y orientation, we have A16 = A26 = 0. Such
laminates are known as balanced laminates. If additionally these plies are
at the same distance about the mid-plane (one above and another below
the mid-plane), then D16 = D26 = 0. An example of a balanced
laminate is [0/+45/ 45/902/0]T, where the subscript T denotes “total” laminate
sequence.
 Symmetric laminate: If a laminate has plies stacked in such a way that through
its thickness the plies are symmetrical about the mid-plane, then Bij = 0. Thus,
such laminates will not exhibit any extension–bending coupling, e.g., [0/30/
452/902/452/30/0]T  [0/30/452/90]s, where the subscript s represents sym-
metry about mid-plane.
 Cross-ply laminate: If the plies are stacked in two orthogonal directions, e.g., in
longitudinal (0 ) and transverse (90 ) directions, the laminate thus built is
known as a cross-ply laminate, e.g., [02/904/0]T.
 Quasi-isotropic laminate: If plies of identical properties and thickness are
oriented in such a way that the angle between any two adjacent plies is equal
to p/n, where n is the number of plies equal to or greater than three, then [A]
becomes directionally independent, thereby showing isotropy in the in-plane
material properties. This construction does not imply that the matrices [B] and
[D] are also isotropic, e.g., [0/45/90]s.
For the special case of a symmetric balanced laminate, the in-plane engineering
moduli can be obtained from the [A], [B], and [D] matrices using

A212 A212
   
Ex ¼ 1h A11 ; Ey ¼ 1h A22 ;
A22 A11 ð2:88Þ
A12 A66
nxy ¼ ; Gxy ¼ ;
A22 h
where h is the total laminate thickness.
The laminate analysis summarized here is inadequate at and near free edges in
laminates. The stress state in a laminate near a free boundary is three-dimensional
and cannot be assumed to be well described by plane stress or plane strain
assumptions. The through-thickness normal and shear stresses can, in some cases,
be significant and could cause laminate failure.

2.4 Linear elastic fracture mechanics

The basic concepts of fracture mechanics are useful in analyzing damage and
failure in composite materials. Here we will briefly review those concepts and
list some of the commonly used results from the linear elastic fracture mechanics.
For more detailed treatment of fracture mechanics comprehensive texts, e.g.,
[13, 15, 17, 47] are suggested.
30 Review of mechanics of composite materials

2.4.1 Fracture criteria


The traditional strength of the materials approach to structural design and mater-
ial selection is based on the notion of yield or failure stress (strength) of a given
material. The fracture mechanics approach instead recognizes the presence of
material flaws whose unstable growth could cause catastrophic failure. To deter-
mine conditions for this type of failure the local stress field in the vicinity of flaws
(modeled as cracks) is analyzed. The singularity of this stress field is characterized
by the so-called stress intensity factor and its critical value is associated with
unstable crack growth. Alternatively, energy balance considerations are used to
find the so-called energy release rate, per unit increment in the crack surface, and
its critical value is associated with the condition of unstable crack growth. For
linear elastic materials undergoing brittle failure the two approaches produce the
same failure criteria.

The stress intensity criterion


Consider an infinite plate with a through thickness edge crack of size a subjected
to a remote tensile stress as shown in Figure 2.8. For a linear elastic plate material
the stress field in close vicinity of the crack tip is given by

     
KI y y 3y
sxx ¼ pffiffiffiffiffiffiffiffi cos 1 sin sin ;
2pr 2 2 2
     
KI y y 3y
syy ¼ pffiffiffiffiffiffiffiffi cos 1 þ sin sin ; ð2:89Þ
2pr 2 2 2
     
KI y y 3y
txy ¼ pffiffiffiffiffiffiffiffi cos sin cos ;
2pr 2 2 2

where r and y are as shown in the figure and KI, known as the stress intensity
factor, is given by
pffiffiffiffiffiffi
KI ¼ s pa: ð2:90Þ
where the subscript I denotes the opening mode (mode I). It can be noted that the
stress field is singular at the crack tip with a r1/2 singularity.
The condition of failure, i.e., unstable crack growth, is assumed when
KI  KIC : ð2:91Þ
KIC, known as the critical stress intensity factor, or fracture toughness, is a
parameter representing the material resistance to fracture, and can be obtained
experimentally.

The energy criterion


In the energy-based approach, one considers a cracked body and examines the
changes brought about by an incremental crack growth in the potential energy of
2.4 Linear elastic fracture mechanics 31

txy
sxx
y
r

q syy

Figure 2.8. Edge crack in a plate in tension.

applied forces – the stored elastic strain energy and the crack surface energy. The
condition for unstable crack growth is then expressed as

G  GC ; ð2:92Þ

where G is the energy available for crack growth per unit of crack surface area,
called the energy release rate, and GC is its critical value, which depends on the
material in which the crack is advancing. GC is viewed as the resistance to crack
growth induced by the material. For a linear elastic material undergoing small-
scale yielding at the crack front, the energy release rate is found to be related to the
stress intensity factor, described above, as

KI2
G¼ ; ð2:93Þ
E0
E
where E0 = E for plane stress condition, and E0 ¼ for plane strain condition.
1 n2

2.4.2 Crack separation modes


A crack is activated, i.e., it produces stresses at its front, when the two crack
surfaces separate. The separation can take place in combination of three
32 Review of mechanics of composite materials

(a) (b) (c)

x3

x2

x1

Figure 2.9. Crack separation modes: (a) opening; (b) sliding; and (c) tearing.

independent modes, denoted as modes I, II, and III, illustrated in Figure 2.9. In
mode I, also called the crack opening mode, the two crack surfaces separate
symmetrically about the crack plane. Mode II is a sliding mode, in which the two
crack surfaces remain in contact and slide past each other in the crack plane.
Finally, mode III, described as the tearing mode, is driven by out-of-plane shear,
resulting in displacement of the two crack surfaces in the x3-direction.
Any displacement of the crack surfaces for a general loading can be viewed as
a superposition of these three modes. Denoting the stress intensity factors in individ-
ual modes as KI, KII, and KIII, the energy release rate for mixed-mode is given by

KI2 KII2 1 þ n 2
G¼ þ 0 þ KIII ; ð2:94Þ
E0 E E
where
pffiffiffiffiffiffi pffiffiffiffiffiffi pffiffiffiffiffiffi
KI ¼ s11 pa; KII ¼ s12 pa; KIII ¼ s13 pa; ð2:95Þ
where the stresses s11, etc. refer to the axes shown in Figure 2.9.

2.4.3 Crack surface displacements


The displacement jump across the two crack surfaces, expressed as

ui ¼ uþ
i ui ; ð2:96Þ
where uþ i and ui represent the displacements of the upper and lower crack
surfaces, respectively, is a quantity of interest in fracture analysis. For the opening
mode of crack separation, mode I, illustrated in Figure 2.10, i = 2, this quantity is
described as crack opening displacement (COD). For an infinite isotropic homo-
geneous medium the COD value is given by
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
x
2
1
u2 ¼ k 1 ; ð2:97Þ
a

which describes an elliptical crack opening profile.


2.4 Linear elastic fracture mechanics 33

x2
+
u2

+
x1

u 2–

Figure 2.10. Crack opening displacement for a crack of size “2a.”

2.4.4 Relevance of fracture mechanics for damage analysis


Fracture mechanics developed and matured well before damage mechanics
emerged. For both fields, the impetus came from the need to analyze failure of
metals. Fracture mechanics initially addressed brittle failure from sharp defects
based on idealized stress analysis of cracks. In contrast, damage mechanics was
concerned with the effect of distributed voids and cracks on the average response of
a solid. For composite materials, the complexity of failure processes involving
a multitude of cracks gave rise to further development of damage mechanics. Today,
damage mechanics of composite materials stands on its own as a mature field solidly
founded in thermodynamics and having a variety of analytical and computational
methodologies associated with it. Fracture mechanics has aided the development of
damage mechanics of composite materials in providing energy-based concepts for
addressing evolution of failure states. However, the stress analysis of cracks, char-
acterized by stress intensity factors, is less relevant to composite damage analysis.
Other than a few cases where single crack growth is a dominant failure mechanism,
such as delamination emanating from free edges in laminates, crack front singular-
ities are of little interest. Indeed, individual cracks constituting damage modes are
usually arrested at interfaces. Therefore, their growth is of little interest. Instead,
energy dissipation occurs due to crack multiplication. Therefore, an appropriate
energy-based analysis, needed to treat this type of situation, does not resort to stress
intensity factors, as is the case in brittle fracture of metals.
34 Review of mechanics of composite materials

References

1. Y. C. Fung, A First Course in Continuum Mechanics. (Englewood Cliffs, NJ: Prentice-


Hall, Inc., 1977).
2. L. E. Malvern, Introduction to the Mechanics of a Continuous Medium. (Englewood
Cliffs, NJ: Prentice-Hall, Inc., 1969).
3. I. S. Sokolnikoff, Mathematical Theory of Elasticity, 2nd edn. (New York: McGraw-Hill, 1956).
4. A. E. H. Love, A Treatise on the Mathematical Theory of Elasticity. (New York: Dover
Publications, 1944).
5. S. P. Timoshenko and J. N. Goodier, Theory of Elasticity. (New York: McGraw-Hill, 1951).
6. R. M. Jones, Mechanics of Composite Materials, 2nd edn. (Philadelphia, PA: Taylor &
Francis, 1999).
7. R. M. Christensen, Mechanics of Composite Materials. (Dover Publications, 2005).
8. J. Qu and M. Cherkaoui, Fundamentals of Micromechanics of Solids. (Hoboken, NJ:
Wiley, 2006).
9. J. N. Reddy, Mechanics of Laminated Composite Plates and Shells: Theory and Analysis,
2nd edn. (Boca Raton, FL: CRC Press, 2004).
10. I. M. Daniel and O. Ishai, Engineering Mechanics of Composite Materials. (Oxford:
Oxford University Press, 1994).
11. A. K. Kaw, Mechanics of Composite Materials. (Boca Raton, FL: CRC Press, 1997).
12. L. P. Kollar and G. S. Springer, Mechanics of Composite Structures. (Cambridge:
Cambridge University Press, 2003).
13. D. Broek, Elementary Engineering Fracture Mechanics, 4th revised edn. (Dordrecht,
Netherlands: Kluwer Academic Publishers, 1991).
14. T. L. Anderson, Fracture Mechanics: Fundamentals and Applications, 2nd edn. (Boca
Raton, FL: CRC Press, 1994).
15. E. E. Gdoutos, Fracture Mechanics: An Introduction, 2nd edn. Solid Mechanics and its
Applications, Vol. 123. (Dordrecht: Norwell, MA: Springer, 2005).
16. K. Friedrich, Application of Fracture Mechanics to Composite Materials. (Amsterdam:
Elsevier, 1989).
17. R. W. Hertzberg, Deformation and Fracture Mechanics of Engineering Materials, 3rd
edn. (New York: Wiley, 1989).
18. K. Washizu, Variational Methods in Elasticity and Plasticity, 2nd edn. (Oxford: Perga-
mon Press, 1974).
19. C. L. Dym and I. H. Shames, Solid Mechanics: A Variational Approach. (New York:
McGraw-Hill, 1973).
20. J. N. Reddy, Energy Principles and Variational Methods in Applied Mechanics, 2nd edn.
(Hoboken, NJ: Wiley, 2002).
21. J. C. Halpin and J. L. Kardos, Halpin–Tsai equations – review. Polym Eng Sci, 16:5
(1976), 344–52.
22. J. C. Halpin and S. W. Tsai, Environmental Factors Estimation in Composite Materials
Design. US Airforce Technical Report, AFML TR (1967). pp. 67–423.
23. Z. Hashin, On elastic behaviour of fibre reinforced materials of arbitrary transverse
phase geometry. J Mech Phys Solids, 13:3 (1965), 119–34.
24. Z. Hashin and S. Shtrikman, A variational approach to the theory of the
effective magnetic permeability of multiphase materials. J Appl Phys, 33:10 (1962),
3125–31.
References 35

25. Z. Hashin and S. Shtrikman, A variational approach to the theory of the elastic
behaviour of multiphase materials. J Mech Phys Solids, 11:2 (1963), 127–40.
26. Z. Hashin and S. Shtrikman, On some variational principles in anisotropic and non-
homogeneous elasticity. J Mech Phys Solids, 10:4 (1962), 335–42.
27. Z. Hashin and S. Shtrikman, A variational approach to the theory of the elastic
behaviour of polycrystals. J Mech Phys Solids, 10:4 (1962), 343–52.
28. Z. Hashin, Variational principles of elasticity in terms of the polarization tensor. Int
J Eng Sci, 5:2 (1967), 213–23.
29. Z. Hashin, Analysis of composite materials – a survey. J Appl Mech, Trans ASME, 50:3
(1983), 481–505.
30. T. Mori and K. Tanaka, Average stress in matrix and average elastic energy of
materials with misfitting inclusions. Acta Metall, 21:5 (1973), 571–4.
31. Z. Hashin, The elastic moduli of heterogeneous materials. J Appl Mech, T-ASME,
29 (1962), 143–50.
32. R. M. Christensen and K. H. Lo, Solutions for effective shear properties in 3-phase
sphere and cylinder models. J Mech Phys Solids, 27:4 (1979), 315–30.
33. R. Hill, A self-consistent mechanics of composite materials. J Mech Phys Solids,
13:4 (1965), 213–22.
34. J. Aboudi, Micromechanical analysis of composites by the method of cells. Appl Mech
Rev, 42:7 (1989), 193–221.
35. J. Aboudi, Micromechanical analysis of composites by the method of cells – update.
Appl Mech Rev, 49:10 Part 2 (1996), S83–91.
36. J. Aboudi, Micromechanical prediction of the effective coefficients of thermo-
piezoelectric multiphase composites. J Intell Mater Syst Struct, 9:9 (1999), 713–22.
37. S. Nemat-Nasser and M. Hori, Micromechanics: Overall Properties of Heterogeneous
Materials, 2nd edn. (Amsterdam: North Holland, 1999).
38. T. Mura, Micromechanics of Defects in Solids, 2nd edn. (New York: Springer, 1987).
39. C. T. Sun, Strength analysis of unidirectional composites and laminates. In Fiber
Reinforcements and General Theory of Composites, ed. T. W. Chou. (Amsterdam:
Elsevier, 2000), pp. 641–66.
40. L. A. Carlsson, D. F. Adams, and R. B. Pipes, Experimental Characterization of
Advanced Composite Materials. (Boca Raton, FL: CRC Press, 2003).
41. R. Hill, The Mathematical Theory of Plasticity. (New York: Oxford University Press,
1998).
42. V. D. Azzi and S. W. Tsai, Anisotropic strength of composites, Exp Mech, 5 (1965),
283–88.
43. S. W. Tsai and E. M. Wu, General theory of strength for anisotropic materials.
J Compos Mater, 5:January (1971), 58–80.
44. Z. Hashin, Failure criteria for unidirectional fiber composites. J Applied Mech, Trans
ASME, 47 (1980), 329–34.
45. P. D. Soden, A. S. Kaddour, and M. J. Hinton, Recommendations for designers and
researchers resulting from the world-wide failure exercise. Compos Sci Technol, 64:3–4
(2004), 589–604.
46. J. N. Reddy, Mechanics of Laminated Composite Plates: Theory and Analysis. (Boca
Raton, FL: CRC Press, 1997).
47. T. L. Anderson, Fracture Mechanics: Fundamentals and Applications, 3rd edn. (Boca
Raton, FL: Taylor & Francis, 2005).
3 Damage in composite materials

All structures are designed for a purpose. If the purpose is to carry loads, then a
designer must assure that the structure has sufficient load-bearing capacity. If the
structure is to function over a period of time, then it must be designed to meet its
functionality over that period without losing its integrity.
These are generic structural design issues irrespective of the material used. There
are, however, significant differences in design procedures depending on whether the
material used is a so-called monolithic material, e.g., a metal or a ceramic, or
whether it is a composite material with distinctly different constituents. The hetero-
geneity of microstructure as well as the anisotropy of properties provide signifi-
cantly different characteristics to composite materials in how they deform and fail
when compared to metals or ceramics. This chapter will review those characteris-
tics. However, before proceeding we need to introduce certain definitions.

Fracture: Conventionally, fracture is understood to be “breakage” of material,


or at a more fundamental level, breakage of atomic bonds, manifesting itself in
formation of internal surfaces. Examples of fracture in composites are fiber
breakage, cracks in matrix, fiber/matrix debonds, and separation of bonded
plies (delamination). The field known as fracture mechanics deals with condi-
tions for formation and enlargement of the surfaces of material separation.
Damage: Damage, on the other hand, refers to a collection of all the irreversible
changes brought about in a material by a set of energy dissipating physical or
chemical processes, resulting from the application of thermomechanical load-
ings. Damage may inherently be manifested by atomic bond breakage. Unless
specified differently, damage is understood to refer to distributed changes.
Examples of damage in composites are multiple fiber-bridged matrix cracking
in a unidirectional composite, multiple intralaminar cracking in a laminate,
local delamination distributed in an interlaminar plane, and fiber/matrix inter-
facial slip associated with multiple matrix cracking. These damage mechanisms
will be explained in some detail later in this chapter. The field of damage
mechanics deals with conditions for initiation and progression of distributed
changes as well as with consequences of those changes on the response of a
material (and by implication, a structure) to external loading.
Failure: The inability of a given material system (and consequently, a structure
made from it) to perform its design function. Fracture is one example of a
3.1 Mechanisms of damage 37

possible failure; but, generally, a material could fracture (locally) and still
perform its design function. Upon suffering damage, e.g., in the form of
multiple cracking, a composite material may still continue to carry loads and,
thereby, meet its load-bearing requirement but fail to deform in a manner
needed for its other design requirements, such as vibration characteristics and
deflection limits. It is a common practice for engineers to predict composite
failure based on any of the multitude of lamina failure criteria described in the
previous chapter. These criteria only predict the final event of failure, and
generally cannot characterize the damage mechanisms leading to the final
failure. In reality, the failure event in a composite structure is preceded and
influenced by the progressive occurrence and interaction of various damage
mechanisms.
Structural integrity: The ability of a load-bearing structure to remain intact and
functional upon the application of loads. In contrast to metals, remaining intact
(not breaking up in pieces) for composites is not necessarily the same as
remaining functional. For instance, composites can lose their functionality by
suffering degradation in their stiffness properties while still carrying significant
loads.
Durability: A term very close in meaning to structural integrity. Specifically,
durability is defined as the ability of a structure to retain adequate properties
(strength, stiffness, and environmental resistance) throughout its life to the
extent that any deterioration can be controlled and repaired [1]. The long-term
durability of a composite structure is an important design requirement in civil,
infrastructure, and aircraft industries.

3.1 Mechanisms of damage

The heterogeneous microstructure of composites, the large differences between


constituent properties, the presence of interfaces as well as directionality of
reinforcement that induces anisotropy in overall properties, are reasons for the
complexities observed in geometrical features of micro-level failure (microcracks)
in composites. Additionally, when interfaces are present, such as between fibers
and matrix and between plies in a laminate, the stress transfer via interfaces
provides conditions for multiple cracking (to be discussed later). The wealth of
observations reported in the literature on various cracking processes, collectively
referred to as “damage mechanisms,” are summarized below for the purpose of
treatments in later chapters related to deformation and failure of composite
materials at a “macro” level.

3.1.1 Interfacial debonding


The performance of a fiber-reinforced composite is markedly influenced by the
properties of the interface between the fiber and the matrix resin. The
38 Damage in composite materials

20 µm

Figure 3.1. Debonds in a fiber-reinforced composite. Reprinted, with kind permission, from
Compos Sci Technol, Vol. 59, E. K. Gamstedt and B. A. Sjögren, Micromechanisms in
tension-compression fatigue of composite laminates containing transverse plies, pp. 167–78,
copyright Elsevier (1999).

adhesion bond at the interfacial surface affects the macroscopic mechanical


properties of the composite. The interface plays a significant role in stress
transfer between fiber and matrix. For instance, if the fibers are weakly held
by the matrix, the composite starts to form a matrix crack at a relatively low
stress. On the other hand, if the fibers are strongly bonded to the matrix, the
matrix cracking is delayed and the composite fails catastrophically because of
fiber fracture as the matrix cracks. The constraint between the fiber and the
matrix also influences other damage mechanisms such as interfacial slipping,
and fiber pull-out. Controlling interfacial properties can thus provide a way to
control the performance of a composite structure. In unidirectional compos-
ites, debonding occurs at the interface between fiber and matrix when the
interface is weak. Figure 3.1 shows debond surfaces observed in a fiber-
reinforced composite [2].
The longitudinal interfacial debonding behavior of single-fiber composites has
been studied in detail by the use of the pull-out [3–7] and fragmentation [8–11]
tests. The mechanics of fiber/matrix interfacial debonding in a unidirectional
fiber-reinforced composite is depicted in Figure 3.2. When fracture strain of the
fiber is greater than that of the matrix, a crack originating at a point of stress
concentration, e.g., voids, air bubbles, or inclusions, in the matrix is either halted
by the fiber, if the stress is not high enough (Figure 3.2(b)), or it may pass
around the fiber without destroying the interfacial bond. As the applied load
increases, the fiber and matrix deform differentially, resulting in a buildup of
large local stresses in the fiber. This causes local Poisson contraction and even-
tually when the shear stress developed at the interface exceeds the interfacial
shear strength, debonding extending over a distance along the fiber results
(Figure 3.2(c)). Shear lag and cohesive zone models are commonly used
approaches to predict initiation of debonding and stress transfer at the interface
[3, 5, 12–17].
3.1 Mechanisms of damage 39

(a) (b) (c)

Figure 3.2. Mechanics of interfacial debonding in a simple composite: (a) perfect laminate;
(b) differential deformation of fiber and matrix crack causes high stresses at fiber/matrix
interface; (c) shear stress exceeds the interfacial shear strength nucleating a debond.
Reprinted from [18], with kind permission from Maney Publishing.

3.1.2 Matrix microcracking/intralaminar (ply) cracking


Fiber-reinforced composites offer high strength and stiffness properties in the
longitudinal direction. Their properties, however, in the transverse directions are
generally low. As a result, they readily develop cracks along fibers. These cracks
are usually the first observed form of damage in fiber-reinforced composites [19].
In laminates with plies in different fiber orientations, these cracks can form from
defects in a given ply and grow traversing the thickness of the ply and running
parallel to the fibers in that ply. The terms matrix microcracks, transverse cracks,
intralaminar cracks, and ply cracks are invariably used to refer to these very same
cracks. Such cracks are found to be caused by tensile loading, fatigue loading, as
well as by changes in temperature or by thermal cycling. They can originate from
fiber/matrix debonds or manufacturing-induced defects such as voids and inclu-
sions [20] (see Figure 3.3). Matrix cracks can also form in ceramic matrix compos-
ites (CMC), and in short fiber composites (SFC). The field of damage mechanics
deals with prediction of formation, growth, and effects of matrix cracks on overall
material behavior. Analysis, design, and behavior of composites subjected to
intralaminar cracking will be dealt with in detail in the subsequent chapters.
Figure 3.4 illustrates matrix cracks observed on the free edges of continuous
fiber and woven fabric polymer composite laminates induced due to fatigue
loading [21, 22]. Although matrix cracking does not cause structural failure by
itself, it can result in significant degradation in material stiffness and can also
induce more severe forms of damage, such as delamination and fiber breakage,
and give pathways for entry of fluids.

3.1.3 Interfacial sliding


Interfacial sliding between constituents in a composite can take place by differential
displacement of the constituents. One example of this is when fibers and matrix in a
composite are not bonded together adhesively but by a “shrink-fit” mechanism
due to difference in thermal expansion properties of the constituents. On
40 Damage in composite materials

(a) (b)

Matrix Void

Fiber
Debond

5 µm 10 µm

Figure 3.3. Matrix crack initiation from: (a) fiber debonds; (b) void results. Reprinted, with
kind permission, from Compos Sci Technol, Vol. 57, C. A. Wood and W. L. Bradley,
Determination of the effect of seawater on the interfacial strength of an interlayer E-glass/
graphite/epoxy composite by in situ observation of transverse cracking in an environmental
SEM, pp. 1033–43, copyright Elsevier (1997).

(a)

(b)

Figure 3.4. Examples of matrix cracks observed in (a) continuous fiber and (b) woven fabric
polymer composite laminates [22]. Part (a) reprinted, with kind permission, from Compos Sci
Technol, Vol. 68, D. T. G. Katerelos, J. Varna, and C. Galiotis, Energy criterion for modeling
damage evolution in cross-ply composite laminates, pp. 2318–24, copyright Elsevier (2008).

thermomechanical loading, the shrink-fit (residual) stresses can be removed,


leading to a relative displacement (sliding) at the interface. The relief of interfacial
normal stress can also occur when a matrix crack tip approaches or hits the
interface.
3.1 Mechanisms of damage 41

When the two constituents are bonded together adhesively, interfacial sliding
can occur subsequent to debonding if a compressive normal stress on the interface
is present. The debonding can be induced by a matrix crack, or it can result from
growth of interfacial defects. Thus, interfacial sliding that follows debonding can
be a separate damage mode or it can be a damage mode coupled with matrix
damage.
Interfacial sliding in ceramic matrix composites (CMCs) can be significant if the
temperature change imposed is high and the thermal expansion mismatch between
the fibers and matrix is also large. When the matrix in a CMC cracks, the resulting
interfacial debonding affects interfacial sliding, causing interactive effect on the
composite deformation [23].

3.1.4 Delamination/interlaminar cracking


Interlaminar cracking, i.e., cracking in the interfacial plane between two adjoining
plies in a laminate, causes separation of the plies (laminae) and is referred to as
delamination. In composite laminates, delamination can occur at cut (free) edges,
such as at holes, or at an exposed surface through the thickness. When loaded in
the plane, the laminate develops through-thickness normal and shear stresses at
the traction-free surface extending a short distance into the laminate plane. These
stresses can result in local cracking in the interlaminar planes. Delaminations can
also form as a result of low-velocity impact [24–26]. In contrast to metals, in
polymer composite laminates delamination can occur below the surface of a
structure under a relatively light impact, such as that from a dropped tool, while
the surface appears undamaged to visual inspection [25, 27, 28]. The growth of
delamination cracks under the subsequent application of external loads leads to a
rapid deterioration of the mechanical properties and may cause catastrophic
failure of the composite structure [29, 30]. Another source of delamination is the
local interlaminar cracking induced by ply cracks. This delamination can grow
and separate the region between two adjacent ply matrix cracks as illustrated in
Figure 3.5.
Delamination can be a substantial problem in designing composite structures
as it can diminish the role of strong fibers and make the weaker matrix
properties govern the structural strength [31]. In initiating delamination the
critical material property is the interlaminar strength, which is determined by
the matrix [26, 31]. Once the interlaminar cracks are formed, their growth is
determined by the interlaminar fracture toughness, which is also governed by
the matrix. If delamination is viewed as decohesion of the cohesive zone
between the separating plies, then both the matrix strength and the fracture
toughness act as material parameters [32]. As a design approach, delamination
can be reduced either by improving the interlaminar strength and fracture
toughness or by modifying the fiber architecture to reduce the driving forces
for delamination [33, 34].
42 Damage in composite materials

Figure 3.5. Interlaminar delamination crack formed due to joining of two adjacent matrix
cracks in a fiber-reinforced composite laminate.

3.1.5 Fiber breakage


The failure (separation) of a fiber-reinforced composite ultimately comes from
breakage of fibers. In a unidirectional composite loaded in tension along fibers the
individual fibers fail at their weak points and stress redistribution between fibers
and matrix occurs, affecting other fibers in the local vicinity of the broken fibers
and possibly breaking some. The fiber/matrix interface transfers the stress from
the broken fiber back to the fiber at a certain distance, making another fiber break
possible if the strength is exceeded by the stress. The fiber breakage process is of
a statistical nature because of the nonuniformity of fiber strength along the fiber
length and the stress redistribution. When plies of unidirectional fibers are stacked
in a laminate, the stress on fibers is enhanced in the vicinity of ply cracks in the
adjacent plies, causing a narrow distribution of fiber failure sites [35]. A greater
number of fiber breaks per unit volume is found closer to the interface where the
ply crack terminates than away from the interface where the local stress concen-
tration falls off [35].
The ultimate tensile strength of a ply within a general laminate is difficult to
predict from the tensile strength of fibers due to the statistical nature of fiber
failure and the progression of fiber failures [36, 37]. Fracture (crack growth)
properties such as the fracture toughness of a composite depend not only on the
failure properties of the constituents but significantly also on the efficiency of
bonding across the interface [38].

3.1.6 Fiber microbuckling


When a unidirectional composite is loaded in compression, the failure is governed
by a mechanism known as microbuckling of fibers. There are two idealized
basic modes of microbuckling deformation, denoted “extensional” and “shear”
modes [39], illustrated in Figure 3.6, depending upon whether the fibers deform
“out of phase” or “in phase” with one another. The corresponding compressive
strength for the onset of instability is given as
3.1 Mechanisms of damage 43

Extensional Mode Shear Mode

Figure 3.6. Extensional and shear modes of fiber microbuckling.

sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Vf E f Em
sc ¼ 2Vf ; ð3:1Þ
3ð1  Vf Þ

for the extension mode, and


Gm
sc ¼ ; ð3:2Þ
1Vf
for the shear mode, where E and G denote Young’s modulus and shear modulus,
respectively, and the subscripts f and m designate fiber and matrix, respectively.
These expressions for idealized deformation modes do not generally agree with
experimental data for compression strength. It has been argued that in practical
composites the manufacturing process tends to cause misalignment of fibers,
which can induce localized kinking of fibers. The kinking process is driven by
local shear, which depends on the initial misalignment angle f0 [40]. The critical
compressive stress corresponding to instability then is given by
ty
sc ¼ ; ð3:3Þ
f0
where ty represents the in-plane shear strength (yielding). Budiansky [41] con-
sidered the kink band geometry (see Figure 3.7) and derived the following estimate
for the kink band angle b in terms of the transverse modulus ET and shear
modulus G of a composite layer:

pffiffi G  sc G  sc
ð 2  1Þ2 < tan2 b< : ð3:4Þ
ET ET

To account for shear deformation effects, Niu and Talreja [42] modeled the fiber
as a generalized Timoshenko beam with the matrix as an elastic foundation. It was
observed that not only an initial fiber misalignment but also any misalignment in
the loading system can affect the critical stress for kinking.
44 Damage in composite materials

W
s

s
j0

Figure 3.7. Kink band geometry assumed in Budiansky [41]. Reprinted, with kind
permission, from Computers & Structures, Vol. 16, B. Budiansky, Micromechanics,
pp. 3–12, copyright Elsevier (1983).

3.1.7 Particle cleavage


If brittle particles (e.g., ceramics) are placed in a ductile but strong and tough
matrix, particle cleavage is the main mode of damage in initial stages of deform-
ation. This mode of damage is found in particulate metal matrix composites.
Cleavage refers to the breakage of the reinforcing particle. The cleavage crack
typically forms perpendicular to the global maximum principal stress. The damage
analysis has been performed assuming viscoplastic material behavior [43]. Failure
of many practically relevant particulate two-phase composites can typically be
attributed to cleavage fracture of the brittle particles followed by ductile crack
growth in the matrix [44]. To account for particle geometry and distribution,
statistical methods are employed to predict inclusion fracture. To fully character-
ize brittle fracture of a particle embedded in a ductile metallic matrix, careful
computational modeling (FEA) sometimes becomes necessary (see, e.g., [45–47]).

3.1.8 Void growth


A composite structure may contain an appreciable amount of manufacturing-
induced defects. For polymer matrix composites, the defects induced during
manufacturing can be in the fiber architecture, e.g., fiber misalignment, irregular
fiber distribution in the cross section, and broken fibers; in the matrix, e.g., voids;
or at the fiber/matrix interface, e.g., disbonds and delaminations. Voids are one of
the primary defects found virtually in all types of composite materials. The
formation of voids is controlled by manufacturing parameters, such as vacuum
pressure, cure temperature, cure pressure, and resin viscosity.
The presence of voids, even at low volume fractions, is found to have a
significant detrimental effect on the overall material behavior. The flexural, trans-
verse, and shear properties are affected the most. Their shape, size, and distribu-
tion also play role in material degradation. Micromechanics homogenization
methods, such as Mori–Tanaka [48], are commonly used to estimate the average
composite property assuming voids as inclusions with zero properties. More
sophisticated methods have also been developed to analyze the effect of voids on
overall composite elastic and failure properties [49].
Voids can also lead to appreciable inelastic deformations in the material locally,
which can act as precursors to initiation of damage processes, such as crazing,
3.1 Mechanisms of damage 45

shear yielding, fibrillation, and local fracture. These damage processes in the final
stage may have significant influence on the deformation response and failure
properties of the composite material.
In composites with metallic and polymer matrices, the matrix phase undergoes
ductile fracture due to nucleation, growth and coalescence of voids and cavities.
These voids grow and expand due to high local inelastic strains and high stress
triaxiality in the matrix. Ductile fracture models, such as Rice–Tracy [50] can be
used to model the initiation and growth of voids in ductile matrices [51]. These
voids can sometimes coalesce to form matrix cracks, and may also cause fiber
matrix debonds.

3.1.9 Damage modes


The damage mechanisms described above have different characteristics depending
on a variety of geometric and material parameters. Each mechanism has different
governing length scales and evolves differently when the applied load is increased.
Interactions between individual mechanisms further complicate the damage pic-
ture. As the loading increases, stress transfer takes place from a region of high
damage to that of low damage, and the composite failure results from the critical-
ity of the last load-bearing element or region. For clarity of treatment, the full
range of damage can be separated into damage modes, treating them individually
followed by examining their interactions.
Which damage mechanisms become active in a given life period of a composite
structure depends mainly on the properties of the base material (e.g., matrix),
architecture, orientation, distribution, and volume fraction of the reinforcing agent
(fiber), the properties of the interface, and loading and environmental conditions.
Intralaminar and interlaminar cracking, fiber fracture, and microbuckling are the
dominant damage mechanisms in long fiber composites. Short fiber composites
show three basic mechanisms of interfacial failure [52], as depicted in Figure 3.8:

Mode a: Localized matrix yielding at the interface due to the stress concen-
tration at the fiber end (see Figure 3.8(a)). Typically, this occurs in combination
with debonding of the fiber end and the formation of a penny-shaped crack.
Mode b: If the interface is relatively weak, an interface crack propagates from
the debonded fiber end (Figure 3.8(b)). This is different than the fiber end
penny-shaped crack and remains closed upon increase in tensile loading on
the composite, and the load transfer occurs by frictional stress transfer.
Mode g: If the interface is relatively strong, a conical matrix crack propagates
from the debonded fiber end at an angle yc to the fiber axis (Figure 3.8(c)). This
matrix crack opens with increasing applied load and suppresses load transfer
across the crack faces.
For particulate composites, the major damage mechanisms are dewetting
(debonding) of the particle and cavity nucleation [53] (see Figure 3.9). At a critical
tensile load, the particles separate from the matrix causing dewetting. Dewetting
46 Damage in composite materials

(b)

(a)

(c)

qc qc

Figure 3.8. Failure mechanisms of interfacial failure in short fiber/epoxy composites:


(a) mode a; (b) mode b; (c) mode g. Reprinted, with kind permission, from Compos Sci
Technol, Vol. 60, S. Sirivedin, D. N. Fenner, R. B. Nath, and C. Galiotis, Matrix crack
propagation criteria for model short-carbon fibre/epoxy composites, pp. 2835–47,
copyright Elsevier (2000).

of the particle eventually leads to cavity formation which grows on subsequent


loading. Dewetting introduces volume dilatation and results in nonlinearity in the
stress–strain behavior. For well-bonded particles, cavities and cracks may form
entirely within the matrix [54].
Damage modes in continuous fiber laminates are thus rich in complexity. These
will be described below in the context of their evolution with loading.

3.2 Development of damage in composite laminates

A schematic description of damage development in composite laminates in tension


is depicted in Figure 3.10, where the five identifiable damage mechanisms are
indicated in the order of their occurrence. Although the figure is developed on
the basis of fatigue experiments [55–60], it provides the basic details for quasi-
static loading as well.
In the early stage of damage accumulation, multiple matrix cracking dominates
in the layers which have fibers aligned transverse to the applied load direction.
3.2 Development of damage in composite laminates 47

s s s

Matrix

Debond
Particle

Cavity

s s s

Figure 3.9. Damage mechanisms in particulate composites. Reprinted, with kind permission,
from Int J Solids Struct, Vol. 32, G. Ravichandran and C. T. Liu, Modeling constitutive
behavior of particulate composites undergoing damage, pp. 979–90, copyright Elsevier (1995).

Static tensile tests on cross-ply laminates have shown that the transverse matrix
cracks can initiate as early as at 0.4% applied strain depending upon the laminate
configuration. They initiate from the locations of defects such as voids, or areas of
high fiber volume fraction or resin rich areas. Ply cracks grow unstably through
the width direction and quickly span the specimen width. As the applied load is
increased (or the specimen is loaded cyclically), more and more cracks appear. The
accumulation of ply cracks in a cracked ply is depicted in Figure 3.11. Initially
these cracks are irregularly spaced and isolated from each other, i.e., have no
interaction among themselves. However, as cracks become closer they start inter-
acting, i.e., the in-between tensile stresses diminish and can no longer build up to
earlier levels. Thus further increase in load is required to produce new cracks. This
is well illustrated in Figure 3.12 by plots of diminishing crack spacing versus load
or number of cycles. The configuration to which crack density saturates, often
reached only under fatigue loading, has been termed the “characteristic damage
state” (CDS) [57–59]. This state seems to mark the termination of the intralaminar
cracking. The uniqueness of the CDS for a given laminate irrespective of the
loading path has, however, not been found to hold in all cases [61].
Subsequent loading causes initiation of cracks transverse to the primary (intra-
laminar) cracks lying in plies adjacent to the ones with those primary cracks (see
Figure 3.10). These cracks, known as secondary cracks, are small in size and they
can cause interfacial debonding, thereby initiating interlaminar cracks. The inter-
laminar cracks are also initially small, isolated and distributed in the interlaminar
planes. Subsequently, some interlaminar cracks merge into strip-like zones leading
to large scale delaminations. This results into loss of the integrity of the laminate
48 Damage in composite materials

1. Matrix Cracking 3. Delamination


5. Fracture

0º 0º 0º 0º
DAMAGE

CDS

0º 0º 0º 0º
2. Crack coupling- 4. Fiber Breakage
Interfacial debonding

PERCENT OF LIFE 100

Figure 3.10. Development of damage in composite laminates [62].

Figure 3.11. Accumulation of intralaminar cracks in an off-axis ply of a composite laminate.


Based on X-ray radiographs reported in [64].

in those regions. Further development of damage is highly localized, increasing


unstably, and involving extensive fiber breakage. The final failure event is mani-
fested by the formation of a failure path through the locally failed regions and is
therefore highly stochastic.
The damage prior to localization is sometimes referred to as sub-critical
damage. The intralaminar (ply) cracking in this stage causes loss of stiffness
properties in the laminate and can by itself lead to loss of functionality (failure)
of the composite structure. The field of “damage mechanics” addresses the initi-
ation and progression of the sub-critical damage. Later chapters will be devoted to
this subject. The next section discusses the phenomenon of multiple cracking and
its effects on overall (average) laminate response.
3.3 Intralaminar ply cracking in laminates 49

10.0

8.0 Fatigue data


Crack spacing (mm)

Quasistatic data
6.0

4.0

2.0

0
0 0.2 0.4 0.6 0.8 1.0
Cycles (× 106)

0 100 200 300 400 500 600 700


Applied stress (MPa)

Figure 3.12. Spacing of cracks in 45 -plies of [0/90/ 45]s graphite/epoxy laminates as a
function of quasi-static and fatigue loading [57]. Reprinted, with kind permission, from
Damage in Composite Materials, ASTM STP 775, copyright ASTM International, 100 Barr
Harbor Drive, West Conshohocken, PA 19428.

3.3 Intralaminar ply cracking in laminates

One of the earliest observations of ply cracking in laminates was reported by


Broutman and Sahu [65]. However, the first major explanation of multiple matrix
cracking was proposed by Aveston, Cooper, and Kelly [66, 67]. They argued that
multiple fracture occurs in a fibrous composite when one of the constituents (fiber or
matrix) fractures at a much lower elongation than the other and when the unbroken
constituent is able to take the additional load; otherwise single fracture results. Later,
a group of experimentalists, Garret, Bailey, Parvizi, and colleagues [68–74] carried
out significant tests to analyze the ply cracking behavior in cross-ply laminates. These
experiments showed the initiation of microcracking in glass-fiber reinforced polyster
and epoxy cross-ply laminates. It was observed that for thick 90 -plies, transverse
cracks initiated at the edge of the specimen and propagated instantly through the
width of entire cross section. As the 90 -plies were made thinner, the strain to initiate
transverse cracks increased (see Figure 3.13). For very thin 90 -plies (< 0.1 mm),
cracks were suppressed and the laminates failed before crack initiation. One of these
experiments [74] involved a microscopy study into the origins of matrix cracks and
revealed that they nucleate from the processing flaws, voids and the regions of high
fiber volume fraction, and progress through fiber-matrix debonding.
The thickness effect on crack initiation can be explained in terms of the
“constraint” posed by uncracked plies over displacement of crack surfaces
in cracked plies (Figure 3.13). On one hand, as the thickness of (cracked) 90 -plies
50 Damage in composite materials

eFPF D
0° Ply failure
%

1.00
Gr./Ep. [04 / 90n]s

C
0.75

0.50 B
A

0.25
90° Ply failure

0
0 2.0 4.0 6.0 8.0 2n Number of
90° plies

Figure 3.13. The strain at first ply failure (eFPF) as a function of the number of transverse
plies in [04/90n]s laminates. Source: [75].

increases, the relative constraint from (uncracked) 0 -plies decreases leading to ply
cracking at lower applied strains. On the other hand, thicker 0 -plies exert a larger
constraint on opening of cracks in 90 -plies, thereby delaying the crack initiation
in those plies. For qualitative understanding, Talreja [75] classified this constraint
in four categories: A – no constraint; B – low constraint; C – high constraint; and
D – full constraint. The stress–strain behavior for each of these four cases varies
greatly and is illustrated in Figure 3.14. On one extreme, it resembles an elastic-
rigid plastic like deformation behavior for constraint type A, and on the other end
a linear elastic behavior for constraint type D.
Over the past four decades, numerous approaches to analyze ply cracking in
composite laminates have been developed. They can be categorized into two broad
categories: micromechanics-based models (Chapter 4), and continuum damage
models (Chapter 5). Based on the dimensionality of the boundary value
problem, micromechanics-based models can be sub-divided into one-dimensional
[59, 76–81], two-dimensional [82–85], and three-dimensional [86–88]. The next
chapter is devoted to these approaches.

3.4 Damage mechanics

Damage mechanics can be broadly defined as the “subject dealing with mechanics-
based analyses of microstructural events in solids responsible for changes in their
3.4 Damage mechanics 51

(a) (b)

s s

eFPF ec eFPF ec e
(c) (d)

s s

eFPF ec e ec e

Figure 3.14. Schematic stress–strain response of cross-ply laminates at different constraints to


transverse cracking: (a) single fracture, no constraint; (b) multiple fracture, low constraint;
(c) multiple fracture, high constraint; (d) multiple fracture, full constraint (with crack
suppression). Source: [75].

response to external loading.” The general objectives of damage mechanics analy-


sis are as follows:
1. Understand the conditions for initiation of the first damage event.
2. Predict the evolution of progressive damage.
3. Characterize and quantify damage in the structure.
4. Analyze the effect of damage on thermomechanical response, e.g., by express-
ing stiffness properties as a function of damage.
5. Assess failure (criticality of damage) and durability of the structure.
6. Provide input into overall structural analysis and design.
This chapter provided an overview of damage development in composites.
The next three chapters will describe analysis methods for quasi-static loading.
Chapter 4 will describe “micro-damage mechanics” (MIDM), whereas Chapter 5
will describe the “macro-damage mechanics” (MADM). Evolution of damage will
be covered in Chapter 6. Damage in fatigue and models for lifetime prediction will
be treated in Chapter 7.
52 Damage in composite materials

References

1. Military handbook, MIL-HDBK-17–3f: Composite Materials Handbook, Vol. 3, Depart-


ment of Defense, USA, 2002.
2. E. K. Gamstedt and B. A. Sjögren, Micromechanisms in tension-compression fatigue of
composite laminates containing transverse plies. Compos Sci Technol, 59:2 (1999), 167–78.
3. C. H. Hsueh, Interfacial debonding and fiber pull-out stresses of fiber-reinforced
composites. Mater Sci Eng A, 154 (1992), 125–32.
4. C. H. Hsueh, Elastic load transfer from partially embedded axially loaded fibre to
matrix. J Mater Sci Lett, 7:5 (1988), 497–500.
5. C. H. Hsueh, Modeling of elastic stress transfer in fiber-reinforced composites. Trends
Polym Sci, 3:10 (1995), 336–41.
6. B. Lauke, W. Beckert, and J. Singletary, Energy release rate and stress field calculation
for debonding crack extension at the fibre-matrix interface during single-fibre pull-out.
Compos Interfaces, 3 (1996), 263–73.
7. L. M. Zhou, J. K. Kim, and Y. W. Mai, Interfacial debonding and fiber pull-out
stresses. J Mater Sci, 27 (1992), 3155–66.
8. L. T. Drzal, The effect of polymer matrix mechanical properties on the fibre-matrix
interfacial strength. Mater Sci Eng A, 126 (1990), 2890–3.
9. J. M. Whitney and L. T. Drzal, Axisymmetric stress distribution around an isolated
fibre fragment. In Toughened Composites, ed. N. J. Johnston. (Philadelphia, PA: ASTM,
1987), pp. 179–196.
10. J. P. Fabre, P. Sigety, and D. Jacques, Stress transfer by shear in carbon fiber model
composites. J Mater Sci, 26 (1991), 189–95.
11. R. B. Henstenburg and S. L. Phoenix, Interfacial shear strength studies using single
filament composite test. Polym Compos, 10 (1989), 389–408.
12. C. H. Hsueh, Interfacial debonding and fiber pull-out stresses of fiber-reinforced
composites. Mater Sci Eng A, 123:1 (1990), 1–11.
13. S. Ghosh, Y. Ling, B. Majumdar, and R. Kim, Interfacial debonding analysis in
multiple fiber reinforced composites. Mech Mater, 32:10 (2000), 561–591.
14. P. Raghavan and S. Ghosh, A continuum damage mechanics model for unidirectional
composites undergoing interfacial debonding. Mech Mater, 37:9 (2005), 955–979.
15. N. Chandra and H. Ghonem, Interfacial mechanics of push-out tests: Theory and
experiments. Compos A, 32:3–4 (2001), 575–584.
16. G. Lin, P. H. Geubelle, and N. R. Sottos, Simulation of fiber debonding with friction in
a model composite pushout test. Int J Solids Struct, 38:46–47 (2001), 8547–62.
17. Y. F. Liu and Y. Kagawa, Analysis of debonding and frictional sliding in fiber-reinforced
brittle matrix composites: Basic problems. Mater Sci Eng A, 212:1 (1996), 75–86.
18. B. Harris, Micromechanisms of crack extension in composites Metal Sci, 14 (1980),
351–362.
19. J. A. Nairn, Matrix microcracking in composites. In Polymer Matrix Composites, ed.
R. Talreja and J. A. E. Manson. (Elsevier Science, 2000), pp. 403–32.
20. C. A. Wood and W. L. Bradley, Determination of the effect of seawater on the interfacial
strength of an interlayer E-glass/graphite/epoxy composite by in-situ observation of
transverse cracking in an environmental sem. Compos Sci Technol, 57:8 (1997), 1033–43.
21. D. T. G. Katerelos, J. Varna, and C. Galiotis, Energy criterion for modelling damage
evolution in cross-ply composite laminates. Compos Sci Technol, 68:12 (2008), 2318–24.
References 53

22. R. Talreja, Damage in woven fabric polymer laminates. (unpublished).


23. D. B. Marshall and A. G. Evans, Failure mechanisms in ceramic-fiber/ceramic-matrix
composites. J Amer Ceram Soc, 68:5 (1985), 225–31.
24. W. J. Cantwell and J. Morton, The impact resistance of composite materials – a review.
Composites, 22 (1991), 347–62.
25. W. C. Chung, et al., Fracture behaviour in stitched multidirectional composites. Mater
Sci Eng A, 112 (1989), 157–73.
26. D. Liu, Delamination resistance in stitched and unstitched composite planes subjected
to composite loading. J Reinf Plast Compos, 9 (1990), 59–69.
27. J. C. Prichard and P. J. Hogg, The role of impact damage in post-impacted compres-
sion testing. Composites, 21 (1990), 503–11.
28. B. Z. Jang, M. Cholakara, B. P. Jang, and W. K. Shih, Mechanical properties in
multidimensional composites. Polym Eng Sci, 31 (1991), 40–6.
29. K. B. Su, Delamination resistance of stitched thermoplastic matrix composite lamin-
ates. In Advances in Thermoplastic Matrix Composite Materials, ASTM STP 1044.
(Philadelphia, PA: ASTM, 1989), pp. 279–300.
30. N. S. Choi, A. J. Kinloch, and J. G. Williams, Delamination fracture of multidirec-
tional carbon-fiber/epoxy composites under mode I, mode II and mixed-mode I/II
loading. J Compos Mater, 33:1 (1999), 73–100.
31. I. Verpoest, M. Wevers, P. DeMeester, and P. Declereq, 2.5D and 3D fabrics for
delamination resistant composite laminates and sandwich structure. SAMPE J, 25
(1989), 51–6.
32. D. J. Elder, R. S. Thomson, M. Q. Nguyen, and M. L. Scott, Review of delamination
predictive methods for low speed impact of composite laminates. Compos Struct, 66
(2004), 677–83.
33. W. S. Chan, Design approaches for edge delamination resistance in laminated compos-
ites. J Compos Tech Res, 14 (1991), 91–6.
34. A. Mayadas, C. Pastore, and F. K. Ko, Tensile and shear properties of composites by
various reinforcement concepts. In Advancing Technology in Materials and Processes,
SAMPE 30th National Meeting (1985), pp. 1284–93.
35. R. Jamison, On the interrelationship between fiber fracture and ply cracking in
graphite/epoxy laminates. In Composite Materials: Fatigue and Fracture, ASTM STP
907, ed. H. T. Hahn. (Philadelphia, PA: ASTM, 1986), pp. 252–73.
36. N. J. Pagano, On the micromechanical failure modes in a class of ideal brittle matrix
composites. Part 1. Coated-fiber composites. Compos B, 29:2 (1998), 93–119.
37. N. J. Pagano, On the micromechanical failure modes in a class of ideal brittle matrix
composites. Part 2. Uncoated-fiber composites. Compos B, 29:2 (1998), 121–30.
38. J. K. Kim and Y. W. Mai, High strength, high fracture toughness fibre composites with
interface control – a review. Compos Sci Technol, 41:4 (1991), 333–78.
39. B. W. Rosen, Tensile failure of fibrous composites. AIAA J, 2 (1964), 1985–91.
40. A. Argon, Fracture of composites. In Treatise on Materials Science and Technology,
ed. H. Herman. (New York, London: Academic Press, 1972), pp. 79–114.
41. B. Budiansky, Micromechanics. Computers & Structures, 16 (1983), 3–12.
42. K. Niu and R. Talreja, Modeling of compressive failure in fiber reinforced composites.
Int J Solids Struct, 37:17 (2000), 2405–28.
43. C. Broeckmann and R. Pandorf, Influence of particle cleavage on the creep behaviour
of metal matrix composites. Comput Mater Sci, 9 (1997), 48–55.
54 Damage in composite materials

44. T. Antretter and F. D. Fischer, Particle cleavage and ductile crack growth in a two-
phase composite on a microscale. Comput Mater Sci, 13 (1998), 1–7.
45. H. J. Böhm, A. Eckschlager, and W. Han, Multi-inclusion unit cell models for metal
matrix composites with randomly oriented discontinuous reinforcements. Comput
Mater Sci, 25 (2002), 42–53.
46. A. Eckschlager, W. Han, and H. J. Böhm, A unit cell model for brittle fracture of
particles embedded in a ductile matrix. Comput Mater Sci, 25 (2002), 85–91.
47. J. Segurado and J. Llorca, A new three-dimensional interface finite element to simulate
fracture in composites. Int J Solids Struct, 41 (2004), 2977–93.
48. T. Mori and K. Tanaka, Average stress in matrix and average elastic energy of
materials with misfitting inclusions. Acta Metall, 21:5 (1973), 571–4.
49. H. Huang and R. Talreja, Effects of void geometry on elastic properties of unidirec-
tional fiber reinforced composites. Compos Sci Technol, 65:13 (2005), 1964–81.
50. J. R. Rice and D. M. Tracey, On the ductile enlargement of voids in triaxial stress fields.
J Mech Phys Solids, 17 (1969), 201–17.
51. H. S. Huang and R. Talreja, Numerical simulation of matrix micro-cracking in short
fiber reinforced polymer composites: initiation and propagation. Compos Sci Technol,
66:15 (2006), 2743–57.
52. S. Sirivedin, D. N. Fenner, R. B. Nath, and C. Galiotis, Matrix crack propagation
criteria for model short-carbon fibre/epoxy composites. Compos Sci Technol, 60:15
(2000), 2835–47.
53. G. Ravichandran and C. T. Liu, Modeling constitutive behavior of particulate com-
posites undergoing damage. Int J Solids Struct, 32 (1995), 979–90.
54. L. R. Cornwell and R. A. Schapery, SEM study of microcracking in strained solid
propellant. Metallography, 8 (1975), 445–52.
55. K. L. Reifsnider and A. Talug, Analysis of fatigue damage in composite laminates. Int
J Fatigue, 2:1 (1980), 3–11.
56. W. W. Stinchcomb, K. L. Reifsnider, P. Yeung, and J. M. Masters, Effect of ply constraint
on fatigue damage development in composite material laminates. In Fatigue of Fibrous
Composite Materials, ASTM STP 723. (Philadelphia, PA: ASTM, 1981), pp. 64–84.
57. J. M. Masters and K. L. Reifsnider, An investigation of cumulative damage develop-
ment in quasi-isotropic graphite/epoxy laminates. In Damage in Composite Materials,
ASTM STP 775, ed. K. L. Reifsnider. (Philadelphia, PA: ASTM, 1982), pp. 40–62.
58. K. L. Reifsnider and R. Jamison, Fracture of fatigue-loaded composite laminates. Int
J Fatigue, 4:4 (1982), 187–197.
59. A. L. Highsmith and K. L. Reifsnider, Stiffness-reduction mechanisms in composite
laminates. In Damage in Composite Materials, ASTM STP 775, ed. K. L. Reifsnider.
(Philadelphia, PA: ASTM, 1982), pp. 103–117.
60. R. D. Jamison, K. Schulte, K. L. Reifsnider, and W. W. Stinchcomb, Characterization
and analysis of damage mechanisms in tension–tension fatigue of graphite/epoxy
laminates. In Effects of Defects in Composite Materials, ASTM STP 836. (Philadelphia,
PA: ASTM, 1984), pp. 21–55.
61. N. V. Akshantala and R. Talreja, A micromechanics based model for predicting fatigue
life of composite laminates. Mater Sci Eng A, 285:1–2 (2000), 303–13.
62. R. Talreja, Internal variable damage mechanics of composite materials. In Yielding,
Damage, and Failure of Anisotropic Solids, ed. J. P. Boehler. (London: Mechanical
Engineering Publications, 1990), pp. 509–33.
References 55

63. Z. Hashin, Analysis of damage in composite materials. In Yielding, Damage, and Failure
of Anisotropic Solids, ed. J. P. Boehler. (London: Mechanical Engineering Publications,
1990), pp. 3–31.
64. A. S. D. Wang, Fracture mechanics of sublaminate cracks in composite materials. In
Composites Technology Review. (Philadelphia, PA: ASTM, 1984), pp. 45–62.
65. L. J. Broutman and S. Sahu, Progressive damage of a glass reinforced plastic during
fatigue. In SPI, 24th Annual Technical Conference, Section 11-D. (Washington, DC: SPI,
1969).
66. J. Aveston, G. A. Cooper, and A. Kelly, Single and multiple fracture. In The Properties
of Fiber Composites. (Surrey, UK: IPC Science and Technology Press, National Phys-
ical Laboratory, 1971).
67. J. Aveston and A. Kelly, Theory of multiple fracture of fibrous composites. J Mater Sci,
8:3 (1973), 352–62.
68. K. W. Garrett and J. E. Bailey, Effect of resin failure strain on tensile properties of glass
fiber-reinforced polyester cross-ply laminates. J Mater Sci, 12:11 (1977), 2189–94.
69. K. W. Garrett and J. E. Bailey, Multiple transverse fracture in 90 cross-ply laminates
of a glass fibre-reinforced polyester. J Mater Sci, 12:1 (1977), 157–68.
70. A. Parvizi, K. W. Garrett, and J. E. Bailey, Constrained cracking in glass fibre-
reinforced epoxy cross-ply laminates. J Mater Sci, 13:1 (1978), 195–201.
71. M. G. Bader, J. E. Bailey, P. T. Curtis, and A. Parvizi, eds. The mechanisms of
initiation and development of damage in multi-axial fibre-reinforced plastic laminates.
Proc Third Int Conf Mech Behav Mater (ICM3), Vol. 3. (Cambridge, 1979), pp. 227–39.
72. J. E. Bailey, P. T. Curtis, and A. Parvizi, On the transverse cracking and longitudinal
splitting behaviour of glass and carbon fibre reinforced epoxy cross-ply laminates and
the effect of Poisson and thermally generated strain. Proc R Soc London A, 366:1727
(1979), 599–623.
73. J. E. Bailey and A. Parvizi, On fiber debonding effects and the mechanism of
transverse-ply failure in cross-ply laminates of glass fiber-thermoset composites.
J Mater Sci, 16:3 (1981), 649–59.
74. A. Parvizi and J. E. Bailey, Multiple transverse cracking in glass-fiber epoxy cross-ply
laminates. J Mater Sci, 13:10 (1978), 2131–6.
75. R. Talreja, Transverse cracking and stiffness reduction in composite laminates.
J Compos Mater, 19:4 (1985), 355–75.
76. P. W. Manders, T. W. Chou, F. R. Jones, and J. W. Rock, Statistical analysis of multiple
fracture in [0/90/0] glass fiber/epoxy resin laminates. J Mater Sci, 19 (1983), 2876–89.
77. H. Fukunaga, T. W. Chou, P. W. M. Peters, and K. Schulte, Probabilistic failure strength
analysis of graphite epoxy cross-ply laminates. J Compos Mater, 18:4 (1984), 339–56.
78. H. Fukunaga, T. W. Chou, K. Schulte, and P. W. M. Peters, Probabilistic initial failure
strength of hybrid and non-hybrid laminates. J Mater Sci, 19:11 (1984), 3546–53.
79. P. S. Steif, Parabolic shear lag analysis of a [0/90]s laminate. Transverse ply crack growth
and associated stiffness reduction during the fatigue of a simple cross-ply laminate. In
S. L. Ogin, P. A. Smith, and P. W. R. Beaumont (eds.), Report CUED/C/MATS/TR
105, Cambridge University, Engineering Department, UK (September 1984).
80. R. J. Nuismer and S. C. Tan, Constitutive relations of a cracked composite lamina.
J Compos Mater, 22:4 (1988), 306–21.
81. S. C. Tan and R. J. Nuismer, A theory for progressive matrix cracking in composite
laminates. J Compos Mater, 23:10 (1989), 1029–47.
56 Damage in composite materials

82. Z. Hashin, Analysis of cracked laminates: a variational approach. Mech Mater, 4:2
(1985), 121–36.
83. J. A. Nairn, The strain energy release rate of composite microcracking: a variational
approach. J Compos Mater, 23:11 (1989), 1106–29.
84. J. Varna and L. A. Berglund, Multiple transverse cracking and stiffness reduction in
cross-ply laminates. J Compos Tech Res, 13:2 (1991), 97–106.
85. L. N. McCartney, Theory of stress transfer in a 0-degrees-90-degrees-0-degrees cross-
ply laminate containing a parallel array of transverse cracks. J Mech Phys Solids, 40:1
(1992), 27–68.
86. P. Gudmundson and W. L. Zang, An analytic model for thermoelastic properties of
composite laminates containing transverse matrix cracks. Int J Solids Struct, 30:23
(1993), 3211–31.
87. E. Adolfsson and P. Gudmundson, Thermoelastic properties in combined bending and
extension of thin composite laminates with transverse matrix cracks. Int J Solids Struct,
34:16 (1997), 2035–60.
88. P. Lundmark and J. Varna, Constitutive relationships for laminates with ply cracks in
in-plane loading. Int J Damage Mech, 14:3 (2005), 235–59.
4 Micro-damage mechanics

4.1 Introduction

As explained in the previous chapter, damage affects the overall stress–strain


response of the solid continuum body. Damage mechanics pertains to the study
of this effect. Two widely different subfields have emerged over the years in this
field. One concerns study of damage directly at the scale of formation of cracks,
i.e., the microstructural scale, and hence can be called “micro-damage mechanics”
(MIDM). The other approach, on the contrary, looks at the overall response at
the macro or structural scale by using some internal variables to characterize
damage, and thus can be termed as “macro-damage mechanics” (MADM). These
terms were originally coined by Hashin [1]. MADM is the same as “continuum
damage mechanics” (CDM), which is still the commonly used terminology.
MIDM for composite materials is derived from an older and more mature field
called micromechanics that deals with overall properties of heterogeneous mater-
ials (see, e.g., [2]). In micromechanics one views heterogeneities such as inclusions
and voids as “microstructure” and estimates overall properties by various
methods, e.g., averaging schemes such as self-consistent and differential schemes,
or variational methods to obtain bounds to average properties. Microcracks are
treated as limiting geometry of microvoids, such as ellipsoidal voids with one
dimension much smaller than the other two. As illustrated in the previous chapter,
“damage” in composite materials has significant complexities concerning the
geometry as well as evolution characteristics such as multiplication of cracks
within a fixed volume. For these reasons a simple extension of micromechanics
to damage in composites is generally not possible. A separate field identified as
MIDM has therefore emerged. This chapter will treat the features of MIDM that
have been developed to specifically treat certain cases of damage in composite
materials. Since determining local (micro-level) stress or displacement fields is a
necessary feature of micromechanics, it is expected that not all cases within the
wide range of damage in composites can be handled by MIDM. However, this
limitation can be alleviated by incorporating computational solutions of the local
stress or displacement fields, thereby broadening classical micromechanics to
include so-called computational micromechanics. In the most recent versions of
MIDM this strategy has been used. More on this will be discussed toward the end
of this chapter.
58 Micro-damage mechanics

In the following we will first treat the aspect of damage in composite mater-
ials that is due to the presence of continuous interfaces between dissimilar
materials, such as fibers and matrix or plies oriented differently in a laminate.
In fact this aspect is fundamental to understanding damage in composite
materials. Historically, it was first analyzed in a classical work by Aveston,
Cooper, and Kelly [3] who explained conditions that lead to failure of a
composite from a single crack versus when multiple cracks precede the final
failure. Their work has become known as the ACK theory. Although the case
treated by them is of simple geometry and loading, namely a unidirectional
fiber-reinforced brittle matrix composite loaded in tension along fibers, it
explains the basic mechanism underlying multiple cracking in a wide range of
cases. The simplified stress analysis and the associated energy balance consider-
ations in the ACK paper have later been extended to include more accurate
solutions, but little further insight into the multiple cracking mechanism has
resulted by these efforts.

4.2 Phenomena of single and multiple fracture: ACK theory

The mode of failure (separation in two or more pieces) in homogeneous mater-


ials such as metals and ceramics may be described as “single fracture” in the
sense that the failure is attributable to a single source – a crack. Heterogeneous
materials, on the other hand, can fail in the mode of single fracture or sustain
multiple fractures of one of the phases before ultimately separating in two or
more pieces. The latter phenomenon is known as “multiple fracture” and can
commonly occur in fibrous composites with brittle matrices, such as cement
plaster, glass, etc.
Damage in composite materials usually initiates with matrix cracking.
Figure 4.1 depicts the isotropic stress–strain response of an unreinforced glass
and of a fiber-reinforced glass in tension loading along fibers. Other than the
enhancement of the stress–strain response in the fiber direction, a striking aspect is
the nonlinearity shown by the reinforced specimen. This nonlinearity is a type of
ductility which occurs due to multiple cracking of the matrix (glass) [4].
Although the phenomenon of multiple fracture was observed earlier, e.g., by
Cooper and Sillwood [5] it was systematically investigated in a landmark paper by
Aveston, Cooper, and Kelly [3]. Their work, the ACK theory, is the basis of the
treatment presented below.
Consider a unidirectional fibrous composite loaded in tension along fibers as
shown in Figure 4.2 and assume the following:
1. Fibers are of the same diameter and uniformly distributed in the matrix.
2. All the fibers are aligned parallel to one another.
3. There are no pre-existing flaws in the matrix such as voids and cracks.
4. Both the matrix and the fibers are linearly elastic.
4.2 Phenomena of single and multiple fracture: ACK theory 59

1000 Fibre-
reinforced
glass
800
Stress/MPa

600

400

200
Unreinforced glass

0 0.4 0.8 1.2


Strain (%)

Figure 4.1. The stress–strain curves for borosilicate glass alone (dotted line) and
reinforced with aligned carbon fibers. Reprinted from [4] with kind permission
from Royal Society Publishing, London.

sc

sm sf

Figure 4.2. A unidirectional fibrous composite loaded in tension.

Assuming that fibers and matrix have different failure strain in tension, when one
of the constituents fails, the other will either fail simultaneously or continue
deforming by carrying the additional load. In the latter case, the constituent that
failed first will fail again at a different site. Thus, there are two necessary condi-
tions for multiple fracture to occur in a composite:
1. One of the constituents has a lower failure strain than the other.
2. When the weaker constituent fails, i.e., when it no longer carries any load,
the stronger constituent must be able to carry the additional load thrown
upon it.
If Pc is the total tensile load on the composite and Pf and Pm represent the load
taken up by the fibers and matrix, respectively, then by force balance we have
Pc ¼ Pf þ Pm : ð4:1Þ
60 Micro-damage mechanics

(a) (b)
s s
sf σf
Multiple Single Single Multiple
fracture fracture fracture fracture
of fibers of matrix
sm
σ f′
A


A
sm
sm

Vf Vf

Figure 4.3. Single and multiple fractures in a unidirectional composite. Fracture stress is
plotted against the fiber volume fraction: (a) case 1: emu > efu; (b) case 2: emu < efu.

Dividing by the composite’s cross-sectional area A, we get


Pc P f Pm
¼ þ ; ð4:2Þ
A A A
or
P c Pf Af Pm Am
¼ þ : ð4:3Þ
A Af A Am A
Assuming unit composite length, we obtain
sc ¼ sf Vf þ sm Vm ; ð4:4Þ
where Vf and Vm are the volume fractions of fibers and matrix, respectively.
Depending upon which phase fails first, two cases as shown in Figure 4.3 arise. Case
1 is when fibers have a lower breaking strain than the matrix (efu < emu), while Case 2 is
for the opposite (emu < efu). In the first case, fibers will undergo multiple fracture if the
matrix is able to carry the additional load thrown upon it due to fiber failure, i.e., if
Pmu > Pc jefu ;
0 ð4:5Þ
i:e:; if smu Vm > sfu Vf þ sm Vm
where Pmu is the maximum load that can be carried by the matrix, sfu and smu
are the tensile strength values for fibers and matrix, respectively, and
0
sm ¼ Em efu ¼ ðsmu =emu Þefu is the stress in the matrix required to produce a strain equal
to the breaking strain of the fibers. In this case, the fibers will be successively fractured
into shorter lengths until the matrix attains its failure strain and at that instant the whole
composite fails. On the other hand, for Case 2, the matrix will undergo multiple fracture if
Pfu > Pc jemu ;
0 ð4:6Þ
i:e:; if sfu Vf > smu Vm þ sf Vf ;
4.2 Phenomena of single and multiple fracture: ACK theory 61

x⬘

Pm

Figure 4.4. Mechanism of load transfer at the fiber/matrix interface.

where Pmu is the maximum load that can be carried by the matrix, sfu and smu
are the tensile strength values for fibers and matrix, respectively, and
0
sf ¼ Ef emu ¼ ðsfu =efu Þemu is the stress in the fibers required to produce a strain
equal to the breaking strain of the matrix.

4.2.1 Multiple matrix cracking


From this point on we will focus on the case of multiple matrix cracking, assuming
emu < efu. In addition to the assumptions described in the previous section, the
following analysis will assume:
1. Fibers remain intact throughout entire loading history.
2. Matrix cracks extend in the entire cross section.
3. Fibers debond completely between adjacent matrix cracks.
If we concentrate on the matrix region between two fibers, the force Pm shed by the
matrix, subsequent to its failure, is carried by fibers in the cracked cross section,
and is transferred back to the matrix over a distance x0 . This load transfer takes
place through shear at the fiber/matrix interface with constant shear stress t. The
mechanism of interfacial load transfer is illustrated in Figure 4.4. The load balance
between the total matrix load Pm and the total shear force at the fiber-matrix
interfaces yields

Pm ¼ smu Am ¼ t
2pr
x0
N; ð4:7Þ
where r is the fiber radius and N is the number of fibers in the composite cross
section of area A.
62 Micro-damage mechanics

Thus,
s
A
mu m
x0 ¼ : ð4:8Þ
t 2prN
Am Am =A Vm :r Vm :r
Now, ¼ ¼ ¼ : Hence,
2prN 2prN=A 2ðN:pr 2 =AÞ 2Vf
s
V r
mu m
x0 ¼ : ð4:9Þ
t Vf 2
The assumption of constant shear stress at the interface simplifies the analysis. However,
its inaccuracy can be noted by realizing that the shear stress at the point where the matrix
crack meets the interface must vanish if the crack surface is to remain traction free.

Stress distribution in fibers and matrix


The balance of forces for a piece of fiber of length Dx0 , over which the change in
the fiber stress is Dsf, yields

sf :pr 2 ¼ t
2pr
x0 : ð4:10Þ
Thus the rate of load transfer is a constant given by
sf t
¼ ; ð4:11Þ
x0 r
and consequently the fiber stress sf varies linearly, and correspondingly sm also
varies linearly, along the fiber axis. The maximum stress in fiber occurs at the
matrix crack and can be determined as
Pmu Am Vm
sf;max ¼ sf þ ¼ sf þ smu ¼ sf þ smu : ð4:12Þ
Af Af Vf
Additional strain due to cracking The strain in fibers increases from emu at
cracking to the maximum value given by
sf;max sf smu Vm smu Em Vm
ef; max ¼ ¼ þ ¼ ef þ ¼ emu ð1 þ aÞ; ð4:13Þ
Ef Ef Ef V f Em E f Vf
with
Em Vm
a¼ : ð4:14Þ
Ef Vf
The mean strain over crack spacing 2x0 is equal to
1 a

emean;2x0 ¼ ½emu þ emu ð1 þ aÞ ¼ emu 1 þ : ð4:15Þ


2 2
When the crack spacing reduces to x0 , the mean strain increases to
 
3a
emean;x0 ¼ emu 1 þ : ð4:16Þ
4
4.2 Phenomena of single and multiple fracture: ACK theory 63

Energy considerations in multiple cracking


Consider the unidirectional composite of Figure 4.2 at fixed applied load Pc =
scA. Let its initial configuration be denoted as state 1 and let state 2 refer to its
configuration with multiple matrix cracking. The energy changes in going from
state 1 to state 2 are described below.

“Supply” of energy
1. DW: the work done per unit cross-sectional area A by an external (fixed) load
through specimen extension caused by cracking is given by
1 a

W ¼ Pc
2x0 ¼ sc
2 emu x0 ¼ Ec emu
emu ax0
A 2 ð4:17Þ
¼ Ec e2mu ax0 :

2. DUm: the reduction in elastic strain energy of matrix over distance 2x0 is given by
ð x0  
1 1 x
2
Um ¼ 2 Em Vm e2mu  Em Vm emu 0 dx
0 2 2 x
2 ð4:18Þ
¼ Em Vm e2mu x0
3
E m Vm 3
¼ e ar:
3t mu

“Consumption” of energy
1. Energy spent in formation of matrix cracks. If gm is the surface energy per unit
area of crack surface, then the energy spent in formation of a matrix crack per
unit cross-sectional area A is
Am
2gm ¼ 2gm Vm : ð4:19Þ
A
2. Energy spent in fiber/matrix interfacial debonding. If we take GII to be the
energy released per unit area of the debond surface, then the debond energy gdb
per unit cross-sectional area A can be expressed as
gdb A ¼ GII
2pr
2x0
N; ð4:20Þ
i.e.,
2prN Af Vm smu
gdb ¼ GII

2x0 ¼ 2GII


r
A Ar Vf t ð4:21Þ
smu
¼ 2GII Vm :
t
3. Us: the energy spent in sliding of the matrix onto the fiber surface over a
distance 2x0 , per unit cross-sectional area A is given by
ð x0
1
Us ¼
N
2 v
t
2pr dx ð4:22Þ
A 0
64 Micro-damage mechanics

where Dn is the sliding displacement at x. This sliding displacement is equal to the


difference in displacements of the fiber and the matrix. It can be found by
integrating the strain in matrix and fiber:
ð x0 ð x0  
a x2 a
x2 x0
tv dx ¼ t emu ð1 þ aÞx  0  x0 1 þ þ dx
0 0 2x 2 x0 2
ð4:23Þ
temu x0 2
¼ ð1 þ aÞ:
6

Thus,
E f Em V m 3
Us ¼ emu rað1 þ aÞ: ð4:24Þ
6t

4. DUf : the increase in the elastic energy of fibers due to additional extension
caused by additional fiber stress, per cross-sectional area A, is given by
ð x0  n 
ð2Þ ð1Þ 1 x
o2 1
Uf ¼ Uf  Uf ¼ 2 Ef Vf emu a 1  0 emu  Ef Vf e2mu dx
0 2 x 2
a

2 0
¼ Ef Vf emu x a 1 þ ð4:25Þ
3
E f E m Vm 3 a

¼ emu ra 1 þ :
2t 3

Conditions for multiple matrix cracking


1. Stress in the matrix is greater than or equal to the matrix failure stress, i.e.,
sm  smu or em  emu : ð4:26Þ
2. The “supply” of energy in going from state 1 to state 2 is greater than or equal
to the “consumption” of energy, i.e.,
2gm Vm þ gdb þ Us þ Uf W þ Um : ð4:27Þ

Substituting Eqs. (4.17)–(4.25) derived above into Eq. (4.27), one obtains
smu
Ec Ef e3mu a2 r
2Vm gm þ GII : ð4:28Þ
t 6t
It can be argued that the energy term GII is much smaller than the other energy
contributions. Assuming GII = 0 then gives

Ec Ef e3mu a2 r
2Vm gm : ð4:29Þ
6t
Thus, the strain required to cause multiple matrix cracking is given by the
following expression
 1=3
12tgm Ef Vf2
emuc ¼ : ð4:30Þ
Ec Em V m r
4.2 Phenomena of single and multiple fracture: ACK theory 65

C D
sfuV f

Vf
Ef

A B
c
E

e fu e
O emu
a 3a
emu 1+ ∼ emu 1+ aemu aemu
2 4 e fu ∼ e fu
2 4

Figure 4.5. The stress–strain response subsequent to multiple fracture for a unidirectional
composite according to the ACK theory.

Stress–strain response
When the composite is loaded to an applied strain level equal to the failure strain of
the matrix, multiple cracking in the matrix starts occurring. If the matrix has a well-
defined single-valued breaking strain, the cracking will continue at a constant applied
stress Ecemu until the matrix is broken down into a set of blocks of length between x0
and 2x0 . The composite stress–strain behavior subsequent to multiple fracture of the
matrix is shown in Figure 4.5. During  multiple
 matrix
 cracking
 (Point A to Point B),
a 3a
the mean strain varies between emu 1 þ 2 and emu 1 þ 4 while going from a crack
spacing of 2x0 to x0 . The total strain at the limit of multiple cracking emc is therefore
 
a
3a
emu 1 þ <emc <emu 1 þ : ð4:31Þ
2 4
When the applied load is increased, fibers are stretched further and start slipping
through the blocks of matrix (point B to Point C). Since the matrix can no longer take
any load the Young’s modulus of the specimen is reduced to EfVf. The composite will
eventually fail at a stress sfuVf. The failure strain of the composite ecu is given by
aemu
aemu

efu  <ecu < efu  : ð4:32Þ


2 4

4.2.2 Perfectly bonded fiber/matrix interface: a modified shear lag analysis


The previous analysis was based on the assumption that the fibers debond com-
pletely from the matrix during the fracture process. In reality the displacements of
matrix and fibers are interrelated. The complete debonding scenario can be viewed
as one extreme, the other extreme being the fully bonded case. For the latter, a
modified shear lag model [6] is applicable.
66 Micro-damage mechanics

For both cases the fundamental equation governing load transfer between fibers
and matrix is obtained from a simple force balance and is, for discrete fibers of
radius r in a continuous matrix, given by

dF 2Vf ti
¼ ; ð4:33Þ
dy r

where dF is the load per cross-sectional area A transferred over the distance dy and
ti is the shear stress acting at the interface.
For the unbonded case, the load transfer from fiber to matrix is given by
Eq. (4.11), whereas for the bonded case the fibers in the plane of the first crack
will be subjected to an additional stress, Ds0, given by

sa
s0 ¼  Ef emu ; ð4:34Þ
Vf

where sa is the applied stress. This additional stress has its maximum value at the
plane of the matrix crack and decays with distance from the crack surface. Using a
modified shear lag analysis, Aveston and Kelly [6] showed that the variation of
this additional stress along the fiber is given by
pffiffi
sð yÞ ¼ s0 e w y ; ð4:35Þ
with
 
2Gm Ec 1
w¼  ; ð4:36Þ
Ef Em Vm r 2 ln Rr
where Gm is the shear modulus of the matrix.
Carrying out a force balance over an element of fiber of length dy,

ds
pr 2 þ ti
2pr
dy ¼ 0; ð4:37Þ

from which the rate of change of stress in the fiber is given by


ds 2
¼  ti : ð4:38Þ
dy r

Differentiating Eq. (4.35) w.r.t. y and substituting into Eq. (4.38), the shear stress
at the interface between the fibers and matrix is given by
r pffiffi
ti ¼ s0 we w y : ð4:39Þ
2
Substituting Eq. (4.39) into Eq. (4.33), and integrating, the load F per cross-
sectional area A transferred to the matrix over any distance l from the crack
surface can be found as
 pffiffi 
F ¼ Vf s0 1  e w l : ð4:40Þ
4.2 Phenomena of single and multiple fracture: ACK theory 67

If Ds0  smu(Vm/Vf), the matrix will undergo further cracking into blocks of
length between l and 2l, where l is obtained by setting F = smu Vm into Eq. (4.40),
 
1 smu Vm
l ¼  pffiffiffi ln 1  : ð4:41Þ
w s0 Vf
The energetic considerations in this case result into the crack initiation strain for
the matrix, as given by
 pffiffiffi
2gm Vm w 1=2
emu ¼ : ð4:42Þ
aEc
The effective Young’s modulus for the matrix can be determined by averaging the
stress distribution in the matrix, i.e.,
ð
1 1 s
Em ¼ sm ð yÞdy; ð4:43Þ
emu s 0
where “2s” is the mutual crack spacing, and the stress distribution in the matrix is
given by
Vf
sm ð yÞ ¼ Em emu  s : ð4:44Þ
Vm
Using Eqs. (4.34), (4.43), and (4.44), Em can be derived as [7]
Em  pffiffi 
Em ¼ Em þ pffiffiffi es w  1 : ð4:45Þ
s w
The effective modulus for the unidirectional composite can then be determined
using the rule of mixture
Ec ¼ Ef Vf þ Em Vm : ð4:46Þ

4.2.3 Frictional fiber/matrix interface


As suggested by Wang and Parvizi-Majidi [7], another important case is to assume
that the fiber/matrix interface is frictional. For this case, the load transfer between
fiber and matrix is assumed constant. After the development of matrix cracking,
the additional stress transferred to the fiber at the crack surface should be
transferred back to the matrix by a constant interfacial shear stress over the
distance of the limiting matrix crack spacing, x. Accordingly, the stress in the
matrix away from the crack surface will vary linearly, from a zero value at the crack
to a maximum of sm at a distance x from the crack. This stress will stay at its
maximum level for distances longer than x. The average stress and the stiffness
values in the matrix can therefore be represented by the following equations
2s  x
 m ¼ sm
s ;
x ð4:47Þ
2s  x
Em ¼ Em :
x
The overall stiffness for the composite can then be calculated from Eq. (4.46).
68 Micro-damage mechanics

The one-dimensional analyses of stress transfer used in the ACK theory and its
later version are inadequate to give accurate prediction, e.g., of the strain to onset
of multiple cracking. The main value of these works was in explaining the phe-
nomenon of multiple cracking. In subsequent sections we will address the problem
of multiple cracking more generally and describe a wide range of approaches
taken since the appearance of the ACK theory.

4.3 Stress analysis (boundary value problem) for cracked laminates

4.3.1 Complexity and issues


Damage analysis of cracked laminates of a general layup is a highly complex task. The
major issues in analyzing damage in a multidirectional laminate are discussed below:
1. Anisotropy and heterogeneity: The commonly used laminate analysis is based
on the assumption that the plies are homogeneous and anisotropic, with the
effective properties of a unidirectional composite. This is a valid assumption
for undamaged laminates for the usual cases of membrane force and moment
loading when the variation of the primary, in-plane, laminate stresses through
the ply thickness is constant or linear. But the presence of intralaminar cracks
may cause local conditions such as high stress gradients to invalidate assump-
tions needed for homogenization of plies [1]. Furthermore, an in-plane stress
analysis will be inadequate to account for interactions between cracks in
different plies.
2. Stress singularity: An assumption of ideal cracks within plies will produce
stress singularity. In reality, however, the crack tips will be blunted due to the
presence of finitely sized fibers near the ply interfaces and by local flow of the
matrix at crack tips. Hence, the actual stress field in the vicinity of the crack
tip may not be accurately given by the usual stress analysis of cracks in
homogeneous elastic medium. In such instances, numerical approaches such
as the finite element method may become necessary.
3. Interaction between cracks: At sufficient crack densities, the stress fields
around two adjacent cracks in a ply start interacting, thereby relaxing the
region between those two cracks. This crack interaction affects the crack
surface displacements as well as the overall stress fields. Accurate modeling
of crack interaction is a complex task.
4. Three-dimensionality of the boundary value problem: Following all the
previous points, and realizing that in a real scenario the ply cracks in a
general laminate may be curved, irregularly spaced, or not fully grown
through the laminate width, a complex 3-D boundary value problem
(BVP) arises. In the simpler case of cracked cross-ply laminate, assuming
that cracks are periodic, straight, and fully grown through the ply thickness
and width, the resulting BVP can be reduced to a generalized plane strain
problem. For the off-axis ply cracking, however, the stress analysis problem
4.3 Stress analysis (boundary value problem) for cracked laminates 69

is still a truly 3-D BVP, and any reduction to a 2-D problem will not produce
accurate predictions.
5. Defining RVE size: In homogenizing a heterogeneous solid such as a composite
ply the RVE size must be large enough to contain sufficiently many fibers to
provide average properties. With fibers of typically 0.01 mm in diameter, and a
ply thickness of typically 0.125 mm, the RVE extending across a ply thickness
may or may not suffice, depending on the fiber volume fraction and fiber
distribution irregularity, but it is implicitly assumed to suffice in the classical
laminate theory. However, when cracks appear within a ply, the local stress
gradients increase sharply, leading to a breakdown of the homogenized ply
properties. Away from the cracks, nevertheless, the properties hold. In
obtaining the overall (average) composite properties with multiple cracks, the
RVE size must be large enough to contain a representation of the cracks. To
satisfy this requirement the RVE must extend in the laminate length direction
while it is limited in the thickness direction by the laminate thickness.
6. Multiscale considerations: Connected with the RVE issue is the consideration
of the multiscale nature of the stress and failure analysis. The RVE scale is the
so-called meso (intermediate) scale, while the scale of discrete entities within
the RVE is the micro scale. The scale at which structural analysis is conducted
is the macro scale. Most of the MIDM is concerned with determining changes
of mesoscale properties using stress fields at the microscale. In fact the
analyses described in the following mostly assume uniform distribution of
cracks, giving repeating unit cells for BVP solutions.
7. Constraint effects: In a cracked laminate, stress perturbations are caused by
the surface displacements of the ply cracks in response to the applied loading.
These surface displacements do not occur freely, as they would if the cracks
were to lie in a homogeneous ply of infinite thickness, but are affected by
constraint from the neighboring plies. Understanding these constraint effects
is the key in determining the effective properties of cracked laminate. They
will be discussed in further detail in a subsequent chapter.
8. Complexity of off-axis ply cracking: Unlike in cross-ply laminates, intralami-
nar cracking in off-axis plies of orientations other than 90 can be complex.
Microscopic observations suggest that depending on the off-axis angle, ply
thickness, and orientation of neighboring plies these cracks may be partially
grown and erratic in shape and size, and of nonuniform distribution [8, 9].
Raman Spectroscopy experiments on [0/45]s laminates show that a crack
developing in the 45 -ply behaves differently from a similar crack in the 90 -
ply of a cross-ply laminate [10], which seemed to suggest that the initiation and
propagation strains for a 45 crack were different. For the laminates containing
a 90 -ply, the cracks usually initiate in that ply under axial tension, while cracks
in other off-axis plies initiate at higher loading. The observations on multidir-
ectional laminates indicate that the angle between the two adjacent plies may
have significant effects on damage initiation and progression. When this angle
is small, partially formed cracks are observed before propagation in the fiber
70 Micro-damage mechanics

(a)

90⬚ 10 mm
60⬚

σx= 315[MPa] εx = 0.71%


(b)

σx= 347[MPa] εx = 0.80%


(c)

σx= 388[MPa] εx = 0.89%


(d)

σx= 428[MPa] εx = 1.00%

Figure 4.6. Consecutive matrix cracking behavior in contiguous plies in a [0/602/90]s


laminate. Parts (a)–(d) show damage state at different levels of applied strain. Reprinted,
with kind permission, from Int J Solids Struct, Vol. 42, T. Yokozeki, T. Aoki, and
T. Ishikawa, Consecutive matrix cracking in contiguous plies of composite laminates,
pp. 2785–802, copyright Elsevier (2005).

direction. However, fully developed cracks mainly form in the cases of large
intersecting angles [11, 12]. The damage development in the 60 -ply of a [0/602/
90]s laminate is shown in Figure 4.6. Moreover, shear–extension coupling may
introduce some additional complexity in analysis of off-axis laminates [13, 14].
9. Randomness in cracking process: In general, damage models assume a uni-
form distribution of transverse cracks, i.e., they are assumed to be periodic
and self-similar. In reality the variation in crack spacing can be considerable,
particularly at large crack spacing. Recently, there have been some develop-
ments to account for the influence of the spatial nonuniformity of matrix
cracking on the stress transfer and the effective mechanical properties of
cracked cross-ply laminates [15–17].
4.3 Stress analysis (boundary value problem) for cracked laminates 71

10. Multiple damage mechanisms: In general, composite laminates can display


a multitude of damage modes. Here, our focus is on ply cracking, which
usually occurs much before other damage mechanisms such as delamina-
tions and fiber fracture. The influence of manufacturing-induced defects
such as voids and fiber clusters may further complicate the analysis. These
interactions may become important for failure analysis and have been
considered in some studies [18–24].
The above list of complex issues concerning cracked laminates, even for simple
configurations, makes the task of stress and failure analysis daunting. The
accuracy of stress analysis by itself may not always be the answer to the
engineering problem at hand. Often approximations and simplifications need to
be made in a judicious manner guided by the application. In the following the
efforts made at analyzing damage and its effects are treated in increasing order of
complexity.

4.3.2 Assumptions
Consider a cross-ply laminate loaded in tension (Figure 4.7(a)). At a certain value
of the load, it develops transverse cracks in the 90 -plies. As described in Chapter
3 these cracks instantaneously traverse the thickness of the 90 -plies and span the
composite specimen width, i.e., the extent of the uniformly distributed load in the
width direction. Thus, regarding the width direction to be infinite in extent,
the boundary value problem can be reduced to a generalized plane strain problem,
as indicated in Figure 4.7(b). The crack spacing 2l shown in the figure is assumed
to be periodic. Experimental observations indicate that although the crack spacing
is nonuniform in the early stage of the cracking process, it attains uniformity soon
and maintains it until crack saturation.
In addition to the assumptions concerning the crack geometry and spacing, the
ply material is assumed to be homogeneous, linearly elastic, with symmetry prop-
erties dictated by the fiber direction. Assuming all fibers to be parallel, the cross-
sectional plane becomes a plane of symmetry. The two mutually orthogonal planes
that are orthogonal to the cross-sectional plane, namely the mid-plane and the
through-thickness plane, can both be assumed as planes of symmetry, rendering the
ply an orthotropic symmetry. Alternatively, one can assume the cross-sectional
plane to be the plane of isotropy, which would make the ply transversely isotropic.

Boundary value problem for cracked cross-ply laminates


As stated above, the cracked cross-ply laminate can be treated as a two-
dimensional BVP shown in Figure 4.7(b). The laminate coordinate system “x–y–z”
is shown in the figure. For the two-dimensional geometry shown in Figure 4.7(b) the
“x–z” axes are placed with the origin located midway between the cracks, as indi-
cated. The local material coordinate system in a given lamina is “x1–x2–x3,” denoted
simply as “1–2–3.”
72 Micro-damage mechanics

(a)

Nxx Nxx

z
y

x
O
(b) z

0º t0
h
t 90
σc 90º x


2l

Figure 4.7. Construction of a unit cell for stress analysis of a cracked cross-ply laminate:
(a) cracked laminate in tension; (b) equivalent 2-D unit cell.

The following are specified for the BVP:

Ply material – transversely isotropic:


E1 ¼ Longitudinal Young’s modulus
E2 ¼ Transverse Young’s modulus
n12 ¼ Major Poisson’s ratio
G12 ¼ In-plane shear modulus
E2
n21 ¼ n12 ¼ Minor Poisson’s ratio
E1
Ex ¼ Longitudinal Young’s modulus for the laminate
Ex0 or Ec ¼ Longitudinal Young’s modulus for the virgin (undamaged) laminate
E0x0 ¼ Longitudinal Young’s modulus for the 0 -ply (undamaged laminate)

E90
x0 ¼ Longitudinal Young’s modulus for the 90 -ply (undamaged laminate)
ð4:48Þ
Geometry:
A ¼ Cross-sectional area
2t0 ¼ Thickness of 0 -ply
2t90 ¼ Thickness of 90o -ply ð4:49Þ
2h ¼ 2ðt0 þ t90 Þ ¼ Total laminate thickness
2l ¼ Average spacing between two adjacent cracks
4.4 One-dimensional models: shear lag analysis 73

Loading:

Nxx ¼ Applied tensile load per unit width in x-direction


ð4:50Þ
(in-plane stress resultant)

Stresses and strains:


sc ¼ Total applied stress for laminate along x-direction
s0xx ¼ Total x-direction stress in 0 -ply

s90
xx ¼ Total x-direction stress in 90 -ply ð4:51Þ

s0xx0 ¼ Initial (virgin-laminate) x-direction stress in 0 -ply
s90
xx0 ¼ Initial (virgin-laminate) x-direction stress in 90 -ply

Other notations will be defined as and when needed. The boundary value problem
can be stated as:

For a cracked cross-ply laminate loaded in tension determine the displacement and stress
fields which satisfy equilibrium and boundary conditions, and further determine its effective
stiffness properties for a fixed state of damage (given crack spacing).

That is, determine sij which satisfy the following conditions:


1. Force balance:

Nxx ¼ sc A: ð4:52Þ

2. Equilibrium conditions:
sij; j ¼ 0: ð4:53Þ
3. Boundary and continuity conditions:
Laminate mid-plane symmetry: s90
xz ðx; 0Þ ¼ 0
Traction continuity across interface: s90 0
xz ðx; t90 Þ ¼ sxz ðx; t90 Þ
s90 0
zz ðx; t90 Þ ¼ szz ðx; t90 Þ
Traction-free boundary: s0xz ðx; hÞ ¼0 ð4:54Þ
s0zz ðx; hÞ ¼0
Traction-free crack surfaces: s90
xz ð l; zÞ ¼ 0; t90 z t90
s90
xx ð l; zÞ ¼ 0; t90 z t90 :

4.4 One-dimensional models: shear lag analysis

A class of one-dimensional models that has been useful in the analysis of


multiple cracking is the so-called shear lag models. Although the stress analysis
is inherently inaccurate in these models, their ability to capture the stress
74 Micro-damage mechanics

transfer at the interface by shear stress makes them useful. Historically, the
shear lag analysis was first used by Cox [25] to describe the stress transfer
between a fiber and a matrix for discontinuous fiber composites. He considered
an axisymmetrical model of a single fiber embedded in the matrix. Later,
Aveston and Kelly [6] modified the model for predicting strain to initiate
multiple matrix cracking in a unidirectional fiber composite with brittle matrix.
A recent work considering axisymmetrical model has shown that Cox’s original
result can be derived by a series of approximations to the elasticity theory [26].
Garrett, Bailey, and Parvizi [27, 28] and Manders et al. [29] adopted Cox’s
approach for the particular case of transverse cracking. Most of the early
studies on transverse cracking used shear lag analysis for the unit cell shown
in Figure 4.7(b). There have been many modifications and extensions of the
same analysis. For a more detailed discussion on shear lag analysis, the reader
is referred to [30–58].
All shear lag analyses are based on the following basic concept:
In the plane of a transverse crack the transverse ply does not carry the axial load,
while away from the crack a part of this load is transferred back to the transverse ply
by axial shear at the interface between the cracked transverse ply and the adjacent
uncracked ply.

The shear lag analysis is essentially a 1-D analysis and it uses the following basic
assumptions:
1. The axial shear stress, txy a dn/dx, where v is the displacement in the
axial, y-direction. This violates the relationship of linear elasticity,
 
dn du
txy ¼ Gxy þ . Hence, this is equivalent to assuming that du/dy = 0 or
dx dy
du/dy  dn/dx.
2. The axial normal stress remains constant over the thickness of the transverse
ply after cracking. In other words, concentration of local stress near cracks is
neglected.
3. Cracks remain sufficiently far apart so that their mutual interactions can be
neglected.
The shear lag analyses will be presented here with consistent notations and
expressions that may differ in form from those given in the original articles.

4.4.1 Initial shear lag analysis


Here we describe the analysis presented by Garrett, Bailey, and Parvizi [27, 43]
and modified by Manders et al. [29] to account for the presence of neighboring
cracks. For convenience, we follow a treatment similar to the one described in a
review paper by Berthelot [41].
The objective is to determine the variation of the axial (x-direction) stress
in the transverse plies on crack formation in these plies. To begin, one
4.4 One-dimensional models: shear lag analysis 75

0º τ

90
90º σ xx 90 90
σ xx + dσ xx

τ

x dx

Figure 4.8. Stresses acting on an element of 90 -ply.

assumes that the x-displacement in the 0 -ply is constant through its thick-
ness while the corresponding displacement of the transverse plies increases
linearly from the 0/90 ply interface towards the laminate mid-plane. In
accordance with the first shear lag assumption stated above, the interfacial
shear stress is given by
 
u90  u0
t ¼ G90
xz ; ð4:55Þ
t90
where u0 is the x-displacement in the 0 -ply, u90 is the x-displacement at mid-plane,

and G90
xz is the transverse shear modulus of the 90 -ply. The equilibrium of axial

forces in an element of the 90 -ply (see Figure 4.8) gives

ds90
xx
t ¼ t90 ; ð4:56Þ
dx

where s90
xz is the axial normal stress, assumed to be constant in the 90 -ply
across the thickness (z-direction). Substituting Eq. (4.56) into Eq. (4.55), one
obtains
 
ds90
xx 90 u90  u0
¼ Gxz : ð4:57Þ
dx t290
The axial stresses in the 0 - and 90 -plies are related to the applied stress sc as

ls0xx þ s90
xx ¼ ð1 þ lÞsc ; ð4:58Þ
where the ply thickness ratio l is defined as
t0
l¼ : ð4:59Þ
t90
Lastly, the stress–strain relations in the 0 - and 90 -plies are:
du0
s0xx ¼ E0x0 e0xx ; with e0xx ¼ ;
dx ð4:60Þ
du90
s90 90 90
xx ¼ Ex0 exx ; with e90
xx ¼ ;
dx
76 Micro-damage mechanics


where E0x0 ¼ E1 and E90x0 ¼ E2 are the initial axial Young’s moduli of the 0 - and

90 -plies, respectively. Differentiating Eq. (4.57) with respect to x and using
Eqs. (4.58)–(4.60), one obtains the differential equation for axial stress in the
transverse ply as

d2 s90
xx b2 90 b2 E90x0
 s ¼  sc ; ð4:61Þ
dx2 t290 xx t290 Ex0
with
 
1 1
b2 ¼ G90
xz0 þ ; ð4:62Þ
E90
x0 lE0x0

where G90
xz0 is the initial in-plane shear modulus of the 90 -ply and E x0 = E c
is the axial modulus of the undamaged laminate given by the rule of
mixtures as
lE0x0 þ E90
x0
Ex0 ¼ : ð4:63Þ
1þl
This value can be estimated accurately by using the classical laminates plate
theory. Since the crack surfaces are traction free, the axial stress in the transverse
ply vanishes on the crack planes (x = l). Hence, the solution of Eq. (4.61) is
given by
0 1
x
cosh b
E90 B
x0 B t90 C
C:
s90
xx ¼ sc @ 1 ð4:64Þ
Ex0 lA
cosh b
t90
Thus b appears as a load transfer parameter, and is sometimes known as the shear
lag parameter.
A quite similar shear lag analysis was conducted by Dvorak et al. [44], except
that they included residual thermal stresses and in place of Eq. (4.55) assumed that
the shear stress is given by

t ¼ K ðu90  u0 Þ; ð4:65Þ

where K is a shear parameter, which is to be determined from experimental data.


Dvorak et al. [44] suggested determining b from the experimentally meas-
ured stress at first ply failure. The differential equation for the axial stress in this
case is
 
d2 s90
xx b2 90 b2 90 E90
x0
 s ¼  s þ s c ; ð4:66Þ
dx2 t290 xx t290 xxR
Ex0

where the shear lag parameter is now given by


   
2 Kt90 t0 E0x0 þ t90 E90
x0 1 1
b ¼ ¼ Kt90 90 þ 0 ; ð4:67Þ
t0 E0x0 E90
x0 Ex0 lEx0
4.4 One-dimensional models: shear lag analysis 77

ts t0

h
t 90
σc 90º x

2l

Figure 4.9. Unit cell for interlaminar shear lag analysis.

and the solution obtained for axial stress in the transverse ply is
0 1
x
  cosh b
E90
x0 B
B t90 C
C;
s90 90
xx ¼ sxxR þ sc @ 1 ð4:68Þ
Ex0 lA
cosh b
t90

where s90
xxR is the axial residual thermal stress in the 90 -ply.

4.4.2 Interlaminar shear lag analysis


Based on extensive experimental observations of cracks, Highsmith and Reifsnider
[30] found that shear deformations in a given ply were restricted to a thin region
near interfaces between adjacent plies. This region is resin rich and thus is less stiff in
its response to shear stress than the central portion of the ply. Transverse cracks
were observed to extend up to the region, but not into it. Based on these observa-
tions, interlaminar shear lag models were developed [30, 45], in which shear stresses
were assumed to develop only within this thin region, whose thickness and shear
modulus are unknown. The unit cell for such models is shown in Figure 4.9.
We present here the interlaminar shear lag analysis reported by Fukunaga et al.
[45], who incorporated the thermal residual stresses and Poisson’s effect into the
analysis. The displacements in the x- and y-directions of each ply are still assumed
to be constant across thickness, as above, but are expressed as
A12
u0 ¼ ec x þ U0 ð xÞ; v0 ¼  ec y þ V0 ð xÞ;
A22
ð4:69Þ
A12
u90 ¼ ec x þ U90 ð xÞ; v90 ¼ ec y þ V90 ð xÞ;
A22
where the coefficients Aij are the in-plane stiffness components of half the cross-ply
laminate. These are the same as the components of the [A] matrix in classical
laminates plate theory, and can be expressed in terms of the reduced stiffness
components Qij of the 0 - and 90 -plies as
A11 ¼ Q11 t0 þ Q22 t90 ; A12 ¼ Q12 h; A22 ¼ Q22 t0 þ Q11 t90 : ð4:70Þ
78 Micro-damage mechanics

These relations suppose that the 0 - and 90 -plies have the same elastic properties.
The strain ec is the average strain of the laminate and is given by
sc
ec ¼ ; ð4:71Þ
Ex
where the axial Young’s modulus of the laminate Ex is given by

A11 A22  A212


Ex ¼ : ð4:72Þ
hA22
The equilibrium of the axial forces yields

d2 u0 G
Q11 þ ðu90  u0 Þ ¼ 0;
dx2 t0 ts
ð4:73Þ
d2 u90 G
Q22  ðu90  u0 Þ ¼ 0;
dx2 t90 ts
where G and ts represent the shear modulus and the thickness of the “resin-rich”
layers, respectively. Next, the balance of the in-plane shear loads gives

d2 v0 G
Q66 þ ðv90  v0 Þ ¼ 0;
dx2 t0 ts
ð4:74Þ
d2 v90 G
Q66  ðv90  v0 Þ ¼ 0:
dx2 t90 ts
The solution of the equilibrium equations, Eqs. (4.73) and (4.74), determines the
displacement field, from which the strains are determined using the strain–dis-
placement relations. Finally, the stress components are found by using the consti-
tutive relations. The axial stress in the 90 -ply is obtained as
0 1
x
   cosh b
Q22 Q12 A12 B B1  t90 C
C;
s90 90
xx ¼ sxxR þ sc 1 @ ð4:75Þ
Ex0 Q22 A22 lA
cosh b
t90
where the shear lag parameter b is given by
 
2 t90 1 1
b ¼G þ : ð4:76Þ
ts Q22 lQ11
The interlaminar shear lag models by Highsmith and Reifsnider [30] differ from
Fukunaga’s analysis [45] in their definition of the shear-lag parameter. For the
Highsmith–Reifsnider model,
 
t90 1 1
b2 ¼ G þ : ð4:77Þ
ts E90
x0 lE0x0
It can be seen that the expressions for the axial stress in the 90 -ply, Eqs. (4.64),
(4.68), and (4.75), are very similar in different shear lag models. Usually, the
definition of the shear lag parameter is different. The biggest limitation of the
4.4 One-dimensional models: shear lag analysis 79

interlaminar shear lag analysis is that the thickness of the boundary layer must be
assumed in a somewhat arbitrary fashion.

4.4.3 Extended shear lag analysis


Lim and Hong [38] modified the interlaminar shear lag model of Fukunaga et al. [45]
to include the effect of crack interaction. The governing differential equations are the
the same as in Fukunaga’s analysis (Eqs. (4.73), (4.74)). However, the boundary
conditions are used such that they are valid for any crack spacing. With this
modification, the axial stress in the 90 -ply is given by
h
i
bx=t90
s90
xx ¼ s 90
xx0 1  a 1 e bx=t90
þ a 2 e ; ð4:78Þ
where
1  e2bl=t90 e2bl=t90  1
a1 ¼ ; a2 ¼ ; ð4:79Þ
e2bl=t90  e2bl=t90 e2bl=t90  e2bl=t90
and the shear lag parameter b is given by Eq. (4.76). The solution in Eq. (4.78) is
different than in Eq. (4.75) due to differences in the assumed boundary conditions.
It is noted that s90
xx0 in Eq. (4.78) is the total stress in the transverse ply before
cracking, and thus includes thermal residual stress, if any.
To account for the progressive shear in the 90 ply, Steif [46] developed a shear
lag theory, where he used an x-displacement that was constant through the 0 layer
thickness but varied parabolically through the 90 layer thickness. This analysis
leads to an expression of the axial stress in the transverse ply similar to that given
in Eq. (4.64), with a modification in the shear lag parameter:
 
2 90 1 1
b ¼ 3Gxz0 90 þ 0 ; ð4:80Þ
Ex0 lEx0
which is three times the value of b2 as given by Eq. (4.62). Later, the same
approach was applied by Ogin et al. [47, 48] to study the stiffness reduction of
glass-fiber cross-ply laminates subjected to quasi-static or fatigue loading. The
transverse shear effects in both 0 - and 90 -plies have been considered by Nuismer
and Tan [49] and Lee, Daniel and Yaniv [31, 56]. These authors consider stress
approaches based on the assumption of linear shear stress variation in each layer
through the thickness, which extends the initial shear lag analysis. Lee and Daniel
[31] relate these linear variations to parabolic variations of the x-displacements
through the thicknesses of the 0 and 90 -plies. These approaches also lead to
modified expressions for the load transfer parameter. For example, the analysis by
Lee and Daniel yields the following expression for the shear lag parameter
 
3G90 1 1
b2 ¼ xz0
þ : ð4:81Þ
G90 90
xz0 Ex0 lE0x0
1þl 0
Gxz0
It can be seen that Eq. (4.81) reduces to Eq. (4.80) when G0xz0  G90
xz0 .
80 Micro-damage mechanics

4.4.4 2-D shear lag models


Flaggs [32], and Nuismer and Tan [33, 49] developed “two-dimensional” shear lag
models. Contrary to the authors’ claims, these models are essentially equivalent to
1-D models with a minor modification to correct for the contraction due to the
Poisson’s effect [42].
The analysis by Flaggs [32] accounts for both applied normal and shear load-
ings and results into a system of two coupled second-order differential equations.
For only normal applied loading, it reduces to a single ODE identical to general
1-D analysis equations, except that the shear lag parameter is given by
0
1
90 90 0 0
B 1 Qyy0  Q Q
xy0 xy0 =Q xx0 C
2@ 0 þ
2 A
lQxx0
Q90 90
xx0 Qyy0  Qxy0
90

b2 ¼   ; ð4:82Þ
1 1 1 l
 þ 0
k2 2 G90 xz0 2Gxz0
where k is a transverse shear correction factor and the Q’s are stiffness coefficients
for the virgin laminate. The Flaggs analysis therefore uses a minor correction for
the Poisson’s contraction introduced by approximate inclusion of the transverse
coordinate (y).
Another 2-D elasticity analysis was performed by Nuismer and Tan [33, 49].
This analysis uses the shear lag parameter defined as
1 1
90
þ 0
Q lQ
b2 ¼ xx0 xx0
: ð4:83Þ
1 l
þ
3G90
xz0 3G0xz0
The analysis also yields a nonzero o (see Eq. (4.85) described in the next subsec-
tion), given by
1 0 
o¼ t90 ti  b2 s90
x0 ; ð4:84Þ
l
0
where ti is the slope of the interfacial shear stress at the crack location. This
analysis is also equivalent to other 1-D models with a minor correction for the
Poisson’s ratio.
More recently, there have been efforts to modify and extend the 1-D and 2-D
shear lag models to enable predictions for laminates of other than cross-ply
layups, such as the equivalent constraint models [13, 51–53] and 2-D displacement
analysis [54, 55]. These models also suffer from deficiencies that are common in the
shear lag models described above.

4.4.5 Summary of shear lag models


It can be shown [42] that all one-dimensional stress analyses discussed above can
be reduced to a generalized form of Garrett and Bailey’s equation [27],
4.4 One-dimensional models: shear lag analysis 81

d2 s
þ b2 s ¼ oðPÞ; ð4:85Þ
dx2
where Ds represents the total stress transferred from the 90 -layer to the 0 -layer,
and is given by

s ¼ s0xx  s0xx0 ; ð4:86Þ

and x = x/t90 is a dimensionless x-coordinate, b is the shear lag parameter, and


o(P) is a function which may depend on the laminate structure, crack spacing, and
the applied load (P). The corresponding boundary conditions, equivalent to
traction-free crack surfaces, are

s90
xx0
sðx ¼ rÞ ¼ ; ð4:87Þ
l
where r = l/t90 denotes the crack spacing normalized by the cracked ply thickness.
Other than in the one-dimensional model by Nuismer and Tan [33, 49], the
function o is chosen to be zero. It can be clearly observed that all the one-
dimensional shear lag models are essentially the same; the difference being in the
choice of the shear lag parameter. The different shear lag models are summarized
in Table 4.1.
The solution of the differential equation (4.85) provides
 90 
o sxx0 o cosh bx
s ¼  2 þ þ 2 : ð4:88Þ
b l b cosh br
With this, the stress in the 0 -layer is given by
 90 
0 0 o sxx0 o cosh bx
sxx ¼ sxx0  2 þ þ 2 ; ð4:89Þ
b l b cosh br
and in the 90 -layer, it is given by
  
lo cosh bx
s90
xx ¼ s 90
xx0 þ 1  : ð4:90Þ
b2 cosh br

Now, the rate of load transfer from the 90 -layer to the 0 -layer, analogous to Eq.
(4.56), can be rewritten as
ds t
¼ : ð4:91Þ
dx l
It is noted here that the difference in Eqs. (4.56) and (4.91) is due to the definition
of Ds (see Eq. (4.86)). Differentiating Eq. (4.88) with respect to x, and comparing
the result with Eq. (4.91), gives the interfacial shear stress as
 
0 lo cosh bx
ti ¼ sxx0 þ 2 : ð4:92Þ
b cosh br
82 Micro-damage mechanics

Table 4.1 A summary of different shear lag models to analyze cracked cross-ply laminates

Shear lag model Feature Shear lag parameter (b2) o(P)


 
Garret and Bailey [27], Simplest shear 1 1 0
G90
xz0 þ
Manders et al. [29] lag model E90 lE0x0
 x0 
Laws and Use first ply 1 1 0
Kt90 90 þ 0
Dvorak [36] failure data to Ex0 lEx0
determine b  
Steif [46], Parabolic 1 1 0
3G90
xz0 þ 0
Ogin et al. [47, 48] displacement 90
Ex0 lEx0
profile  
Highsmith and Use an effective t90 1 1 0
G þ
Reifsnider [30] shear transfer ts E90
x0 lE0x0
layer  
Fukunaga et al. [45] Use an effective t90 1 1 0
G þ
shear transfer ts Q22 lQ11
layer
 
Lim and Hong [38] Use an effective t90 1 1 0
G þ
shear transfer ts Q22 lQ11
layer, account
for crack
interaction
Nuismer and Tan Account for 1 1 1 0 
þ 0 t90 t i  b2 s90
x0
[33, 49] Poisson’s effect Q90
xx0 lQ xx0 l
1 l
þ 0
3G90
xz0 3Gxz0
0 1
Flaggs [32] 2-D shear lag
90
Q90 0
xy0 Qxy0
0
analysis to B 1 Q yy0  0 C
B Qxx0 C
account for 2B 0 þ
2 C
@lQxx0 90 90 90 A
both transverse Qxx0 Qyy0  Qxy0
and shear  
loading 1 1 1 l
 þ 0
k2 2 G90 xz0 2Gxz0
 
Lee and Daniel [31] Parabolic 3G90
xz0 1 1 0
displacement þ 0
G90 E90 lE
variation, 1 þ l 0xz0 x0 x0
Gxz0
account for
Poisson’s effect

The stress predictions from various shear lag models are compared in Figure 4.10 for
[0/902]s carbon/epoxy (Hercules AS4/3501–6) laminate with lamina properties [31,
56]: tply = 0.154 mm, E1 = 130 GPa, E2 = 9.7 GPa, G12 = 5.0 GPa, G13 = 3.6 GPa,
n12 = 0.3, and n13 = 0.5. The axial normal stress in the 90 -plies increases from zero on
the crack to a maximum value midway between two cracks, as the load is transferred
from the 0 -plies back into the 90 -plies. The traction-free boundary condition at the
crack surfaces is thus fulfilled accurately for the normal stress in all models. However,
none of the shear lag models represents interfacial shear stress accurately. The shear
stress has a maximum on the crack surface and decays towards zero as we move away
4.4 One-dimensional models: shear lag analysis 83

60

40

20
σ xz (MPa)

0
90
90
xx⬘
σ

–20
Manders et al.
Laws & Dvorak
Fukunaga et al.
–40
Lee & Daniel
Steif
–60
–4 –3 –2 –1 0 1 2 3 4
ξ

Figure 4.10. Variation of the axial normal and interfacial stresses between two cracks in a
[0/902]s laminate from shear lag analysis assuming normalized crack spacing r = 4 and
applied stress of 50 MPa.

from the crack. The nonzero shear stress on the crack surfaces is a clear violation of
the boundary condition. This makes all 1-D analyses inherently inaccurate.
Let us now turn our attention to the stiffness degradation. The treatment
provided here follows the approach covered in a book chapter by Nairn and
Hu [42]. Before cracking, both the 0 - and 90 -plies share the applied load. But
after cracking, the 90 -layers at the crack planes do not carry any load. The total
applied load at these planes is carried by the 0 -plies. Away from cracks, the
90 -plies take up load again through the shear transfer mechanism. To determine
the modulus of the cracked laminate, let us first determine the total x-displacement
of the load bearing layer in the region between two cracks in the transverse ply.
It can be found by integrating the axial strain as
ðl ð þr !
0 s0xx n0xz s0zz n0xy s0yy
uð P Þ ¼ exx dx ¼ t90  0  dx; ð4:93Þ
l r E0x Ex E0x

where P is the applied load. It is noted that the thermal strain is ignored because for
linear thermoelastic materials the modulus is independent of the residual stresses.
The compliance of a unit cell of the laminate with cracks is defined as
u ð PÞ  uð 0Þ
C¼ : ð4:94Þ
P
The effective modulus is then given by
2l
E¼ : ð4:95Þ
hWC
84 Micro-damage mechanics

Assuming a plane stress state,

s0yy ¼ 0: ð4:96Þ

Moreover there is no load applied in the thickness (z) direction, i.e.,


ð þr
s0zz dx ¼ 0: ð4:97Þ
r

Thus, for both one-dimensional analysis and two-dimensional plane stress analy-
sis the total displacement reduces to
ð þr
s0xx
uðPÞ ¼ t90 0
dx: ð4:98Þ
r Ex

Substituting stress from Eq. (4.89) for 1-D shear lag analyses, we obtain
   
1 1 E90 tanh br o ð P Þ  o ð 0Þ tanh br
¼ 0 1 þ x00  1  : ð4:99Þ
E x Ec lEx0 br E0x0 b2 s0 br

Assuming o(P)o(0) = 0, as is the case for most one-dimensional analyses, the


effective axial modulus is then
 
1 1 E90
x0 tanh br
¼ 1þ 0 : ð4:100Þ
Ex E0c lEx0 br

Thus, the axial modulus for the cracked laminate normalized with its virgin value
is given by
 1
Ex E90
x0 tanh br
¼ 1þ 0 : ð4:101Þ
E0c lEx0 br

The predictions for stiffness reduction using various shear lag models as a function of
crack density for [0/903]s glass/epoxy (Scotch Ply 1003) laminate are shown in Figure
4.11. It should be noted, however, that these predictions may vary depending upon
the parameters used in the models. It is important to note that the original shear lag
model by Garrett, Bailey, and Parvizi [27, 43], or equivalently by Manders et al. [29],
is reasonably accurate. More complicated shear lag models not only involve more
complicated analysis, and adjustable parameters, but also do not yield more accurate
predictions. The basic deficiency of shear lag models lies in their one-dimensionality
of stress field, and no significant improvement is essentially possible.

4.5 Self-consistent scheme

The self-consistent method is widely employed for estimating properties of elastic


solids containing entities such as inclusions, voids, and cracks; see for example
Nemat-Nasser and Hori [2]. Using this method generally requires assuming
4.5 Self-consistent scheme 85

1.0

[0/903]s
0.9 Glass/Epoxy
Ogin et al.
Relative Stiffness

0.8
Nuismer & Tan
Reifsnider
0.7 Flaggs
Garrett & Bailey
0.6
Hashin

0.5
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Microcrack Density (1/mm)

Figure 4.11. Variation of normalized Young’s modulus as a function of crack density for a
[0/903]s glass/epoxy (Scotch Ply 1003) laminate. The experimental data are from [30].
Reprinted, with kind permission, from Damage Mechanics of Composite Materials,
J. A. Nairn and S. Hu, Micromechanics of damage: a case study of matrix microcracking,
pp. 187–243, copyright Elsevier (1994).

an infinite solid. For composite laminates containing ply cracks, Laws and
Dvorak [59] estimated elastic properties using this method. They first assumed that
a unidirectional ply (within a laminate) with cracks along its fibers can be modeled
as an infinite solid with the same fiber volume fraction and crack density. Taking
then the estimated average properties of the infinite solid as those of the cracked ply,
they replaced the cracked ply in a composite laminate by a homogeneous ply of the
degraded properties of the infinite solid. Figure 4.12 illustrates the assumed scheme.
Elastic properties of the cracked composite laminate were then calculated by the
classical laminate plate theory. This procedure in principle allows calculating
stiffness degradation of any composite laminate with given crack density in any
ply, i.e., for any orientation of cracks. It is doubtful, however, that the assumption
of an infinite solid will give accurate elastic property estimates for plies that are
typically of a fraction of a millimeter in thickness. The predicted properties of
various laminates with cracks in different plies and with different crack density
values were reported in Laws and Dvorak [59] but these properties were not
compared with experimental data or independent numerical computations.
Readers interested in details of the self-consistent method for composite mater-
ials with cracks are referred to Laws et al. [60]. We provide here a brief account
of the concepts involved. Essentially, Hill [61] and Budiansky [62] developed the
self-consistent method for composite materials, i.e., homogeneous solids with
inclusions. The idea in this method is that a single inclusion is embedded in a
homogeneous elastic solid which has the yet-unknown overall properties of the
heterogeneous solid and the local fields thus estimated are then used to obtain the
overall properties. Laws et al. [60] use a key result obtained for ellipsoidal
inclusions by Eshelby [63, 64] that the elastic field within the inclusion is uniform.
This result was used for two models: a three-phase system and a two-phase system.
86 Micro-damage mechanics

(a) (b)

Figure 4.12. Self-consistent scheme: (a) cracked laminate containing fibers of small
diameter and periodic cracks (b) homogenized cracked lamina.

In the former the fibers and cracks are considered as two phases (special cases of
inclusions) embedded in an elastic matrix, while in the latter the fibers and matrix are
smeared into a homogeneous solid in which cracks are embedded. Cracks in both
models are introduced as slits that are regarded as limiting cases of elliptical voids.
When the fiber diameter is much smaller than the crack size, the two-phase
model is more appropriate. Thus if Q is the estimated stiffness matrix of the
cracked ply, and Q0 is its value without the cracks present, then the two matrices
are related by [60]
1
Q ¼ Q0  bQ0 Q; ð4:102Þ
4
where b is a crack density parameter and L is a function of the aspect ratio of the
elliptical crack and the stiffness coefficient matrix Q.
It is seen from this equation that the estimation procedure for stiffness of a
cracked ply is iterative and it requires quantities L that are estimated by assuming
the homogenized composite ply to be of infinite extent.
Variations of the self-consistent method described above have also appeared in
the literature. Noteworthy work is that of Hoiseth and Qu [65, 66]. They followed
the differential self-consistent method and derived an incremental differential
equation describing the effective axial modulus for the cracked laminate by
representing the change in strain energy due to an increase in terms of the number
of cracks in a lamina. If r = t90/l denotes the normalized crack density, then the
axial modulus is given by
4.6 2-D stress analysis: variational methods 87

  
dEc d  E1
¼ 2Ec ðrÞ  ; ð4:103Þ
dr 4t90 1  n12 n21
subject to the initial condition
Ec ð0Þ ¼ Ec0 ; ð4:104Þ
where d is the average crack opening displacement (COD), defined as
ð
 1 t90
d¼ dðzÞ dz; ð4:105Þ
t90 0
with d(z) being the COD variation across the z-direction (thickness). The authors
obtained it numerically using finite element calculations on a unit cell. For cross-
ply laminates, their predictions compared well with the independent FE simula-
tions that they carried out themselves. For multidirectional laminates, these
approaches become complex and do not yield accurate estimates.

4.6 2-D stress analysis: variational methods

4.6.1 Hashin’s variational analysis


An improved 2-D stress analysis can be obtained using the principle of minimum
complementary energy applied on a cracked laminate volume. Hashin [67] con-
sidered cross-ply laminates and used this principle to solve the boundary value
problem. He constructed an admissible stress field assuming that the normal
stresses in the loading direction are constant over the ply thickness. The admissible
stress field satisfies equilibrium and boundary and interface conditions. The
boundary volume problem thus solved provides the stresses and the reduced
stiffness coefficients for the cracked cross-ply laminate.
We follow the general notations for geometry, material, and stress components as
given in Section 4.3. Consider a symmetric cross-ply laminate containing a parallel
array of transverse cracks and subjected to a uniform axial load (Figure 4.13)
Nxx per unit laminate width. Let us first investigate the stress states before and after
cracking. Before cracking, all shear stresses and the transverse normal stress in the
entire laminate are zero, i.e.,
s0xx0 ; s90
xx0 6¼ 0; s0yy0 ; s90
yy0 6¼ 0;
szz0 ¼ s90
0
zz0 ¼ 0; syz0 ¼ s90
0
yz0 ¼ 0; ð4:106Þ
s0xz0 ¼ s90xz0 ¼ 0; s0xy0 ¼ s90xy0 ¼ 0;

where superscripts denote the layer orientations. The nonzero axial stresses in the
0 - and 90 -layers for virgin laminate are constant throughout individual layers
and can be obtained using the laminate theory as

E0x0 E90
s0xx0 ¼ sc ; s90
xx ¼
x0
sc ; ð4:107Þ
Ex0 Ex0
where sc = Nxx/2h is the normal tensile stress on the laminate.
88 Micro-damage mechanics

Nxx

z
O

t0 t90 t90 t0
h h

Figure 4.13. A cross-ply laminate loaded in axial tension.

Now on sufficient tensile load, the 90 -ply develops continuous intralaminar


cracks in the fiber direction that extend through the whole laminate width. These
cracks will cause perturbation in the stress fields in both layers. In a homogeneous
elastic material, there is theoretically a crack-tip singularity in the stress field.
However, for transverse matrix cracking in composite laminates, the stresses at a
crack tip are finite because the fibers have a finite size and will blunt the crack tip.
In Hashin’s analysis, he assumes the following:
1. The perturbations in axial stresses s90 0
xx and sxx are constant through the
thickness of layers. Thus,
s90 90 0 0
xx ¼ sxx ð xÞ; sxx ¼ sxx ð xÞ: ð4:108Þ
2. There are no perturbations in syy, sxy, and syz, i.e.,
s90 90 90 0 0 0
yy ¼ sxy ¼ syz ¼ syy ¼ sxy ¼ syz ¼ 0: ð4:109Þ

The stress components for layer m (m = 0, and 90 for 0 -ply and 90 -ply,
respectively) of the cracked laminate can be expressed as
sm m m
ij ¼ sij0 þ sij ð4:110Þ
where the subscript 0 indicates the uncracked laminate. The stress perturbations
can be expressed as
s90 90
xx ¼ sxx0 f90 ð xÞ;
ð4:111Þ
s0xx ¼ s0xx0 f0 ð xÞ;
where f90 and f0 are unknown perturbation functions.
4.6 2-D stress analysis: variational methods 89

The equilibrium equations for the plies are given by


m
@sm
xx
@sxy @sm
þ þ xz ¼ 0;
@x @y @z
@sm
yx @s m
yy @s m
yz
þ þ ¼ 0; ð4:112Þ
@x @y @z
@sm
zx
@sm
zy @sm
þ þ zz ¼ 0:
@x @y @z
Using Eqs. (4.108)–(4.110), one obtains
@smxx @smxz
þ ¼ 0;
@x @z ð4:113Þ
@smzx @smzz
þ ¼ 0:
@x @z
Substituting Eq. (4.111) into Eq. (4.113) and integrating, one gets
0
s90 90
xz ¼ sxx0 ½f90 ð xÞz þ f90 ð xÞ;
 
90 1 00 0
ð4:114Þ
s90 ¼ s f ð x Þz 2
þ f ð x Þz þ g 90 ð x Þ ;
zz xx0
2 90 90

for m = 90 (90 -ply), and

s0xz ¼ s0xx0 ½f00 ð xÞz þ f0 ð xÞ;



1 ð4:115Þ
szz ¼ sxx0 f000 ð xÞz2 þ f 00 ð xÞz þ g0 ð xÞ;
0 0
2
for m = 0 (0 -ply), where f0(x), f90(x), g0(x), and g90(x) are unknown functions
and primes denote derivatives with respect to x.
To obtain the relation between f90 and f0, consider equilibrium in the
x-direction of the undamaged laminate
ðh
 
Nxx ¼ sxx0 dz ¼ 2 s90 0
xx0 t90 þ sxx0 t0 : ð4:116Þ
h

If the same membrane force is applied to the cracked laminate, the equilibrium in
the x-direction will give
   90 
Nxx ¼ 2 s90 0 0
xx0 t90 þ sxx0 t0  2 sxx0 t90 f90 ð xÞ þ sxx0 t0 f0 ð xÞ : ð4:117Þ

From Eqs. (4.116) and (4.117), we get

s90 0
xx0 t90 f90 ð xÞ þ sxx0 t0 f0 ð xÞ ¼ 0; ð4:118Þ
i.e.,

s90
xx0 1
f0 ð xÞ ¼  f ð xÞ; ð4:119Þ
s0xx0 l 90
where l = t0/t90 is the ply thickness ratio (Eq. (4.59)).
90 Micro-damage mechanics

t90 t0

Figure 4.14. A repeating unit cell for a cracked cross-ply laminate.

The boundary value problem for solving the three stress components sxx, sxz,
and szz is defined on the unit cell shown in Figure 4.14, in which the origin of the
coordinate system is placed at the mid-plane of the laminate, and midway between
two cracks.
Clearly z = 0 is the plane of symmetry. Hence, the shear stress sxz at all points
along this plane is zero. At the interface between the cracked 90 -ply and the
uncracked 0 -ply (z = t90), the shear stress sxz and normal stress szz must be continu-
ous. Moreover, the surface at z = h is free from any external loading, i.e., the shear
stress sxz and normal stress szz are zero on that surface. Finally, the crack surfaces
x = l are traction free, i.e., s90 0 90 90 90
xz ¼ sxz ¼ 0 and sxx ¼ sxx0 þ sxx ¼ 0 at x = l.
Thus, all the applicable boundary conditions for the total stresses are written as

Symmetry : s90
xz ðx; 0Þ ¼ 0; ðaÞ
Interface : s90
xz ðx; t90 Þ ¼ s0xz ðx; t90 Þ; ðbÞ
90
szz ðx; t90 Þ ¼ s0zz ðx; t90 Þ; ðcÞ
Free boundary : s0xz ðx; hÞ ¼ 0; ðdÞ ð4:120Þ
s0zz ðx; hÞ ¼ 0; ðeÞ
Traction free : s90
xz ð l; zÞ ¼ 0; t90 z t90 ; ðfÞ
90
sxx ð l; zÞ ¼ 0; t90 z t90 ; ðgÞ

where 2l is the distance between any two adjacent cracks. These conditions are the
same as those given earlier without derivation in Eq. (4.54)
Denoting f90(x) by f(x), putting boundary conditions, Eq. (4.120), into Eqs.
(4.114)–(4.115), and using Eq. (4.119), after some mathematical manipulations,
the resulting stress field in the cracked laminate is given by
4.6 2-D stress analysis: variational methods 91

s90 90
xx ¼ sxx0 ½1  fð xÞ;
90 0
s90
xz ¼ sxx0 f ð xÞz; ð4:121Þ
90 00 1 
s90
zz ¼ sxx0 f ð xÞ ð1 þ lÞt290  z2 ;
2
in the 90 -ply, and
1
s0xx ¼ s0xx0 þ s90
xx0 fð xÞ;
l
0 1
s0xz ¼ s90
xx0 f ð xÞ ½ð1 þ lÞt90  z;
ð4:122Þ
l
1
s0zz ¼ s90 00
xx0 f ð xÞ ½ð1 þ lÞt90  z2 ;
2l
in the 0 -ply. Using the crack surface boundary conditions, Eq. (4.120f–g), we get
additional conditions for the unknown perturbation function f(x)

At x ¼ l : s90 90
xx ¼ 0 ¼ sxx0 ½1  fð xÞ;
0
ð4:123Þ
s0xz ¼ 0 ¼ s90
xx0 f ð xÞz :

Thus,
0
fð lÞ ¼ l; f ð lÞ ¼ 0; t90 z t90 : ð4:124Þ
This completes the description of the resulting boundary value problem. The object-
ive is to estimate f(x) in order to determine the stress field in the cracked laminate.
The stress field in Eqs. (4.121)–(4.122) represents an admissible stress field as it
satisfies all equilibrium and boundary conditions. Thus, the principle of minimum
complementary energy (see Section 2.1.5) can be applied to obtain the unknown
function f(x). The procedure for doing this, given in Hashin [67], is described next.
For the problem at hand, we can define an admissible stress field seij within the
volume V of the cracked body with associated tractions Tei which satisfy only
equilibrium and the traction boundary conditions. Since the tractions are not
altered during the cracking process,

T~i ¼ Ti0 on S ¼ ST ; ð4:125Þ


where Ti0 are the tractions applied on the uncracked body. Also, the crack surfaces
Sc are traction free, i.e.,

T~i ¼ 0 on S ¼ Sc : ð4:126Þ
If we write

~ij ¼ s0ij þ s0ij ;


s T~i ¼ Ti0 þ Ti0 ; ð4:127Þ
0
where s0ij are stresses in the uncracked laminate, sij are the perturbation stresses
due to cracking, and T0 is the additional traction due to cracking. Then from the
traction boundary conditions in Eqs. (4.125) and (4.126), we obtain
92 Micro-damage mechanics

T 0 i ¼ 0 on S ¼ ST ;
ð4:128Þ
T 0 i ¼ Ti0 on S ¼ Sc :

The complementary energy functional is given by (see Eq. (2.36), Section 2.1.5)
ð ð
1
0 ¼ Sijkl s0ij s0kl dV  Ti0 u^i dS ð4:129Þ
2
V Su

for the uncracked body, where Sijkl is the compliance tensor, and
ð ð
~  ¼ 1 Sijkl s
 ~kl dV  T~i u^i dS
~ij s ð4:130Þ
2
V Su

for the cracked body. Since only traction boundary conditions are applied, the second
term is zero (Su = 0; ST = St). However, for the sake of completeness, we will follow
the proof given in Appendix 1 of [67], which applies to mixed boundary conditions.
Substituting the stress field given in Eq. (4.127) into Eq. (4.130) and using Eq. (4.129),
the complementary energy for the cracked solid can be expressed as
ð ð ð
~  ¼  þ Sijkl s0 skl0 dV þ 1 Sijkl s0ij s0kl dV  T 0i u^i dS :
 ð4:131Þ
0 ij
2
V V Su
0
Consider the first integral in Eq. (4.131). The stresses sij satisfy equilibrium since
s0ij and seij do. Also, the strain field in the uncracked body is given by

e0ij ¼ Sijkl s0kl : ð4:132Þ

Therefore, by virtual work (see Eq. (2.33), Section 2.1.5)


ð ð ð
J ¼ Sijkl skl sij dV ¼ T i ui dS þ T 0i u0i dS0 ;
0 0 0 0
ð4:133Þ
V Su Sc

where the tractions in the second surface integral are defined with respect to the
inward normal to Sc. Using the boundary condition in Eq. (4.128), we have
ð ð
0
J ¼ T i u^i dS  Ti0 u0i dS : ð4:134Þ
Su Sc

The second integral is taken over the two adjacent surfaces of each crack. Since the
normal on one surface is the negative of the normal on the other surface and since
Ti0 and u0i are continuous across the crack surface, the integrals on the two crack
surfaces cancel one another and therefore the total integral vanishes. Introducing
the remainder of Eq. (4.134) into Eq. (4.131) we obtain the complementary energy
for a cracked solid as
ð
~   1
 ¼ 0 þ Sijkl s0ij s0kl dV ¼ 0 þ 0 ð4:135Þ
2
V
4.6 2-D stress analysis: variational methods 93

0
where П0 is the complementary energy due to perturbation stresses. The effective
elastic compliances of the undamaged and damaged laminates can be expressed in
terms of the complementary energy functionals. To obtain the relation, consider
homogeneous boundary conditions on the cracked body
Ti ¼ s ij ¼ constant;
ij nj ; s ð4:136Þ
and let
ð
1
ij ¼
s sij dV: ð4:137Þ
V
V

Then, from Eq. (4.130) we have


1
 ¼ Sijkl s
ij s
kl V; ð4:138Þ
2
where Sijkl is the effective elastic compliance tensor for the homogenized medium [68].
The laminate under consideration is loaded by a membrane force Nxx on two
horizontal edges. Thus, volume and the average stress for this laminate are given by
Nxx
V ¼ 2Ah sxx ¼ ¼ sc ; ð4:139Þ
2h
where A is the area of the plane normal to the x-axis. The complementary energy
for cracked and uncracked laminates takes the form

1 s2c 1 s2c
0 ¼
2Ah  ¼
2Ah; ð4:140Þ
2 Ex0 2 Ex
where Ex0, Ex are the longitudinal moduli of undamaged and damaged laminates,
respectively. Now, from the principle of minimum complementary energy, it
follows that if the admissible stress system, Eqs. (4.121)–(4.122), with boundary
conditions, Eq. (4.124), is introduced into Eq. (4.135) then

~    :
 ð4:141Þ

Using Eqs. (4.135) and (4.140) into Eq. (4.141), we obtain

1 s2c 1 s2c

2Ah þ 0 
2Ah; ð4:142Þ
2 Ex0 2 Ex

where П0 represents the perturbation in complementary energy due to cracking. Hence,

1 1 0
þ 2 : ð4:143Þ
Ex Ex0 sc Ah

Therefore, the variational boundary value problem results in the following calcu-
lus of variations problem:
Find f(x) that minimizes П0 .
94 Micro-damage mechanics

The complementary energy change П0 is the sum of energies brought about by


perturbation stresses in the individual laminae, i.e.,
ð l ð t90 ðl ðh
0
 ¼2 W90 dz dx þ 2 W0 dz dx; ð4:144Þ
l 0 l t90

where W90 and W0 are stress energy densities due to perturbation stresses in the
90 - and 0 -layers, respectively. The stress energy density in a transversely iso-
tropic unidirectional fiber composite is given by
1
W ¼ sij eij
2  2   2 

1 s211 s22 þ s233 2n12 s11 ðs22 þ s33 Þ 2n23 s22 s33 s223 s12 þ s213
¼ þ   þ þ ;
2 E1 E2 E1 E2 G23 G12
ð4:145Þ
where 1 is in the fiber direction and 2, 3 are transverse directions, and the elastic
moduli are defined in Eq. (4.48). To determine the perturbation stress energy, we
can express stress energy densities in the 90 - and 0 -plies in terms of the perturb-
ation stresses in the corresponding plies. The stresses in each lamina are given in
terms of perturbation stresses by

s22 ¼ s90
xx ; s23 ¼ s90
xz ; s33 ¼ s90
zz ; s11 ¼ s12 ¼ s13 ¼ 0 ð4:146Þ
for the 90 -ply, and

s11 ¼ s0xx ; s13 ¼ s0xz ; s33 ¼ s0zz ; s12 ¼ s22 ¼ s23 ¼ 0 ð4:147Þ
for the 0 -ply. Thus, the energy densities in the two sets of laminae are
" 2  90 2  90 2 #
1 s90 xx szz 2n23 s90xx szz
90 sxz
W90 ¼ þ  þ ;
2 E2 E2 E2 G23
" 2  0 2  0 2 # ð4:148Þ
1 s0xx szz 2n12 s0xx s0zz sxz
W0 ¼ þ  þ :
2 E1 E2 E1 G12

Consider again the laminate region bounded by two transverse planes through
adjacent cracks, Figure 4.14. Due to symmetry about z = 0, it is sufficient to take
one half of the laminate l x l, 0 z h with unit width in y direction.
Further, the symmetry across x results into
fð xÞ ¼ fðxÞ; ð4:149Þ
Substituting Eqs. (4.121), (4.122), and (4.148) into Eq. (4.144), and carrying out
the integration along z, П0 is given by
ð
 90 2 l h 2 2
i
0
 ¼ sxx0 t90 C00 f2 þ t390 C02 ff00 þ t390 C11 f0 þ t590 C22 f00 dx; ð4:150Þ
l
4.6 2-D stress analysis: variational methods 95

where
 
1 1 2 u23 l u12
C00 ¼ þ ; C02 ¼ l þ  ;
E2 lE1 3 E2 3 E 1
  ð4:151Þ
1 1 l   1
C11 ¼ þ ; C22 ¼ ðl þ 1Þ 3l2 þ 12l þ 8 :
3 G23 G12 60E2
Introducing a nondimensional geometry parameter x = x/t90, Eq. (4.150) can be
rewritten as
ðr " 2  2  2 2 #
 2 d f df df
0 ¼ s90 2
xx0 t90 C00 f2 þ C02 f 2 þ C11 þ C22 dx; ð4:152Þ
r dx dx dx2

where r = l/t90 denotes the crack spacing normalized with the cracked ply thickness.
The Euler–Lagrange equation for f(x) is
d4 f d2 f
þ p þ qf ¼ 0; ð4:153Þ
dx4 dx2
where the coefficients p and q are given by
C02  C11 C00
p¼ ; q¼ : ð4:154Þ
C22 C22
This is a fourth-order ODE whose characteristic equation is

r 4 þ pr 2 þ q ¼ 0 : ð4:155Þ

The roots of the characteristic equation are


pffiffiffiffiffiffiffi
r ¼ ða1 þ ia2 Þ; i ¼ 1;
y y
a1 ¼ q1=4 cos ; a2 ¼ q1=4 sin ;
s2ffiffiffiffiffiffiffiffiffiffiffiffiffiffi 2 ð4:156Þ
4q
y ¼ tan1  1;
p2

provided that 4q/p2 > 1. The general solution to Eq. (4.153) will involve terms with
e a1 x cos a2 x and e a1 x sin a2 x. However, it is more convenient to use hyperbolic
functions instead of exponentials, because of symmetry condition, Eq. (4.149).
Since only even product functions are admissible solutions, the solution of the
fourth-order ODE, Eq. (4.153), leads to the following expression for f(x)

f ¼ A1 cosh a1 x cos a2 x þ A2 sinh a1 x sin a2 x; ð4:157Þ

where A1 and A2 are constants determined from the boundary conditions as

2ða1 cosh a1 r sin a2 r þ a2 sinh a1 r cos a2 rÞ


A1 ¼ ;
a1 sin 2a1 r þ a2 sinh 2a2 r
ð4:158Þ
2ða2 cosh a1 r sin a2 r  a1 sinh a1 r cos a2 rÞ
A2 ¼ ;
a1 sin 2a1 r þ a2 sinh 2a2 r
96 Micro-damage mechanics

when 4q/p2 < 1, f(x) is given by [68]


0 0 0 0
a2 cosh a1 x a1 cosh a2 x
f¼ þ ;
sinh a1 rða2 coth a1 r  a1 coth a2 rÞ sinh a2 rða1 coth a20 r  a20 coth a10 rÞ
0 0 0 0 0 0 0

ð4:159Þ
0 0
where a1 ; a2 are defined as
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
rffiffiffiffiffiffiffiffiffiffiffiffiffiffi
0 0 p p2
a1 ; a2 ¼  q; ð4:160Þ
2 4

assuming that p < 0, which usually holds good for typical glass/epoxy and
graphite epoxy materials [69]. Once f is known, the stress field can be calculated
from Eqs. (4.121)–(4.122). The effective stiffness coefficients for the cracked
laminate are then determined by calculating average stresses and strains.
It is important to note here that variational analysis gives lower bounds of stiffness
properties for the cracked laminates, as can be seen from the inequality in Eq. (4.143).

4.6.2 Effect of residual stresses


As a laminate is cooled from the cure temperature, thermal residual stresses are
generated due to differential thermal expansion of the 0 and 90 -plies. These
residual stresses can be easily incorporated by laminate theory for uncracked
laminate. The effect of thermal stresses on crack initiation was investigated in
detail by [70]. The variational stress formulation for cracked cross-ply laminate
including thermal residual stresses is covered in [69, 71]. The complementary
energy analogous to Eq. (4.152) is then given by [69]
ðr
0 ¼ 00 þ t290
r
"  2  2 2 #
2 d2 c dc dc d2 c
C00 c þ C02 c 2 þ C11 þ C22  2a Tc þ C2T 2 dx;
dx dx dx2 dx
ð4:161Þ
where
 
a T a T
c¼ s90
xx0  fþ ; ðaÞ
C00 C00
a ¼ a22  a11 ; T ¼ T0  Tref ; ðbÞ ð4:162Þ
     2
n23 sc 2 n12 sc l
C2T ¼ a22 T  þ l þ a22 T  ; ðcÞ
Ec 3 Ec 3
where a11 and a22 are the coefficients of thermal expansion in the longitudinal and
transverse directions, respectively, T0 is the service temperature, and Tref is the
stress-free (reference/curing) temperature of the laminate. The corresponding
Euler–Lagrange equation for this case is
4.6 2-D stress analysis: variational methods 97

d4 c d2 c a T
4
þ p 2
þ qc ¼ : ð4:163Þ
dx dx C22
The solution for f remains the same as in Eqs. (4.158)–(4.160), from which c can be
obtained using Eq. (4.162a). The stresses can then be calculated using Eqs. (4.121)–(4.122).

4.6.3 [0m/90n]s vs. [90n/0m]s laminates


The stiffness changes due to cracking in [0m/90n]s laminates are different than in [90n/
0m]s laminates because in the former case we have cracks in an internal 90 -layer,
whereas the cracks in the latter are exposed to the free boundary. For the same area
of crack surfaces, the tractions on the crack surfaces (see Eq. (4.128)) do more work
in closing (or opening) the crack surfaces in the latter case, and, therefore, the
perturbation complementary energy has a higher value. This results in a lower
effective axial stiffness as compared to laminates with cracks in internal plies.
Nairn [69] carried out a variation of the solution for transverse cracking in
[90n/0m]s laminates. The procedure is the same as for [0m/90n]s laminates. The
obtained solution for f is also the same, except that some of the constants
appearing in Eqs. (4.150) and (4.162) are different. The new constants are as follows
 
2 n12 n23
C02 ¼ 1 þ l  ;
3 E1 3E2
  1
C22 ¼ ðl þ 1Þ 3 þ 12l þ 8l2 ; ð4:164Þ
60E
   2  
1 n23 sc n12 sc 2
C2T ¼  a22 T þ þ a22 T l þ l2 :
3 Ec Ec 3
A comparison of the stress profiles in [0/902]s and [902/0]s laminates is illustrated in
Figure 4.15(b)–(c). A discussion follows later.

4.6.4 Improved variational analysis


Some major improvements to Hashin’s variational analysis were suggested by Varna and
Berglund [71–73] with the most refined model described in [73]. In Hashin’s approach,
the axial stress variation across the thickness is assumed constant, whereas the Varna–
Berglund approach determines these by further application of the principle of minimum
complementary energy. The stress functions in the 90 - and 0 -plies are chosen as
 2 2  
90 90 
z 
z 
 ¼ sxx0  cðxÞ þA þ c1 ðxÞ’2 ðzÞ t290 ;
2 2
   ð4:165Þ
0 0 z2 90 90 2
 ¼ sxx0  sxx0 ’1 ð zÞcðxÞ þ sxx0 c1 ðxÞ’3 ðzÞ t90 ;
2

where s0xx0 ands90
xx0 are the axial stresses in the 0 - and 90 -plies of uncracked
laminate, respectively, x ¼ x=t90 and z ¼ z=t90 are nondimensional coordinates,
98 Micro-damage mechanics

(a) 60 (MPa)
90
s xx 50
40
30
20
Crack 10 Crack
x
–3 –2 –1 –10 1 2 3
–4 4
–20
90
s xz (interfacial) –30
–40

(b) 60 (MPa)
90 50
σ xx
40
30
20
Crack Crack
10
x
–3 –2
–1 –10 1 2 3
–4 4
–20
–30 σ 90
zz (interfacial)
90
σ xz (interfacial)
–40

(c)
70 (MPa)
60 s 90
zz (interfacial)
90
s xx 50
40
30
20
Crack 10 Crack
x
–3 –2 –1 1 2 3
–10
–4 4
–20
–30
s 90
xz (interfacial)

Figure 4.15. Profile of stresses between two adjacent ply cracks in a carbon/epoxy cross-ply
laminate: (a) using 1-D shear lag analysis for [0/902]s; (b) using 2-D variational analysis for
[0/902]s; and (c) using 2-D variational analysis for a [902/0]s laminate. The normal stress s90
xx
is same for the entire 90 -ply group, whereas shear stress s90 90
xz and transverse stress szz are
plotted at the 0/90 ply interface x = 1. Reprinted, with kind permission, from Damage
Mechanics of Composite Materials, J. A. Nairn and S. Hu, Micromechanics of damage: a
case study of matrix microcracking, pp. 187–243, copyright Elsevier (1994).

and A* is a constant. In the above expressions, the arbitrary function ’2 ðzÞ repre-
sents nonuniformity in the x-axis stress distribution in the 90 -layer close to the
cracks along z, whereas c1 ðxÞ characterizes this stress variation along the x-direc-
tion. For the 0 -layer, the redistribution in the stress field is described by c1 ðxÞ and
’3 ðzÞ along the x- and z-directions, respectively. If we neglect ’2 ðzÞ; ’3 ðzÞ; c1 ðxÞ and
set A ¼ h=2t90 and ’1 ðzÞ ¼ ðh  zÞ2=2t0 t90 we revert back to Hashin’s model.
For the current model, the stress components in layer m are given by
4.6 2-D stress analysis: variational methods 99

@ 2 m @ 2 m @ 2 m
sm
xx ¼ sm
zz ¼ sm
xz ¼ 
: ð4:166Þ
@z2 @x2 @x@z
Thus,
(a) in the 90 -layer

s90 90
Þ þ c1 ðxÞ’002 ðzÞ;
xx ¼ sxx0 ½1  cðx
0
s90 90
xz ¼ sxx0 ½c ðx Þz  c0 1 ðxÞ’02 ðzÞ;
 2   ð4:167Þ
00 
z  00
s90
zz ¼ s 90
xx0 c ð 
x Þ þ A þ c 1 ð 
xÞ’ 2 ð 
z Þ ;
2

(b) in the 0 -layer

s0xx ¼ s0xx0  s90 Þ’001 ðzÞ  c1 ðxÞ’003 ðzÞ;


xx0 ½cðx
0
s0xz ¼ s90 Þ’01 ðzÞ  c01 ðxÞ’03 ðzÞ;
xx0 ½c ðx ð4:168Þ
00
s0zz ¼ s90
xx0 ½c ðxÞ’1 ðzÞ þ c001 ðxÞ’3 ðzÞ :

The boundary and interface conditions remain the same as given in Eq. (4.120),
which, after using Eqs. (4.167) and (4.168), results in

1
’01 ð1Þ ¼ 0; ’1 ð1Þ ¼ A þ ; ’01 ðhÞ ¼ ’1 ðhÞ ¼ 0;
2
’03 ðhÞ ¼ ’3 ðhÞ ¼ 0; ’02 ð1Þ ¼ ’03 ð1Þ ¼ 1; ’2 ð1Þ ¼ ’3 ð1Þ; ð4:169Þ
cð rÞ ¼ 1; c0 ð rÞ ¼ c01 ð rÞ ¼ c1 ð rÞ ¼ 0;

where h ¼ h=t90 . Clearly, this model is complex and requires determining the
constant A*, and the functions cðxÞ; c1 ðxÞ; and ’1 ðzÞ; ’2 ðzÞ’3 ðzÞ. However, the
authors found that the following choice of functions ’1 ðzÞ; ’2 ðzÞ; ’3 ðzÞ is useful
1  cosh 1 ðz  hÞ
’ 1 ð
zÞ ¼ ;
1 sinh 1 l
z2n
’ 2 ð
zÞ ¼ A þ ; ð4:170Þ
2n
1  cosh 3 ðz  hÞ
’ 3 ð
zÞ ¼ ;
3 sinh 3 l
with A a constant and D1, D3, and n arbitrary shape parameters, where n is an
integer. Using the boundary conditions described in Eq. (4.169), we obtain
1 1  cosh 1 l 1 1  cosh 3 l
A ¼  þ ; A¼ þ : ð4:171Þ
2 1 sinh 1 l 2n 3 sinh 3 l
Now, the total complementary energy of the cracked laminate system is given by
ðr
0 ¼ 00 þ vðc; c;0 c00; c1 ; c01 ; c001 Þ d
x; ð4:172Þ
r
100 Micro-damage mechanics

0
where П0 is the complementary energy of the virgin laminate, and v is the comple-
mentary energy density due to perturbation stresses. Using the minimization
procedure as described above, we finally obtain the following system of ordinary
differential equations with constant coefficients
 
C022 c000 þ C002  C011 c00 þ C000 c  C01 000 00 01
22 c 1  CR c 1  C00 c1 ¼ 0;
000 00 000
  00
ð4:173Þ
 C01 01 1 1 1
22 c  CR c  C00 c þ C22 c 1 þ C02  C11 c 1 þ C00 c1 ¼ 0;
1

with the boundary conditions as in Eq. (4.169) and constants given by


E2 1 A 2
C000 ¼ 1 þ I1 ; C022 ¼ þ þ A þ I 2 ;
E1 20 3
 
E2 E2 1 E2
C011 ¼ þ I3 ; C002 ¼ 2n23 A þ  2n12 I4 ;
3G23 G12 6 E1
E2
C100 ¼ I1T þ I1L ; C122 ¼ I2T þ I2L ;
E1
E2 T E2 L E2 L
C111 ¼ I þ I ; C102 ¼ 2n23 I4T  2n12 I ; ð4:174Þ
G23 3 G12 3 E1 4
E2
C01
00 ¼ 1 þ I1C ; C01 C
22 ¼ F3 þ I2 ;
E1
E2 E2 C E2 C2
C01
11 ¼ F1 þ I ; C01
02 ¼ 2n23 F2  2n12 I ;
G23 G12 3 E1 4
E2 1  01 
C01
20 ¼ 2n23 F4  2n12 I4C1 ; CR ¼ C þ C01 01
02  C11 ;
E1 2 20
where
ð h ð h ð h
2 2
I1 ¼ ½’00 1 ð
zÞ d
z; I2 ¼ ½’1 ðzÞ2 dz; I3 ¼ ½’01 ðzÞ dz;
1 1 1
ð h ð1 ð1
2
I4 ¼ zÞ’001 ð
’ 1 ð zÞd
z; I1T ¼ ½’002 ðzÞ dz; I2T ¼ ½’2 ðzÞ2 dz;
1 0 0
ð1 ð1 ð1
2
I3T ¼ ½’02 ð
zÞ d
z; I4T ¼ ’2 ðzÞ’002 ðzÞdz; F1 ¼ z’2 ðzÞdz;
0 0 0
ð1 ð1  
z2 1
F2 ¼ ’2 ð
zÞd
z; F3 ¼ ’2 ðzÞ þ A dz; F4 ¼ þ A  ’ 2 ð 1 Þ þ F2 ;
0 0 2 2
ð h ð h ð h
2 2
I1L ¼ ½’003 ð
zÞ d
z; I2L ¼ ½’3 ðzÞ2 dz; I3L ¼ ½’03 ðzÞ dz;
1 1 1
ð h ð h ð h
I4L ¼ zÞ’003 ðzÞd
’ 3 ð z; I1C ¼ ’1 00 ðzÞ’003 ðzÞdz; I2C ¼ ’1 ðzÞ’3 ðzÞdz;
1 1 1
ð h ð h ð h
I3C ¼ 0
’1 ðzÞ’03 ð
zÞd
z; I4C1 ¼ ’1 ðzÞ’003 ðzÞdz; I4C2 ¼ ’001 ðzÞ’3 ðzÞdz:
1 1 1
ð4:175Þ
4.6 2-D stress analysis: variational methods 101

The minimum value of the complementary energy corresponding to the obtained


solution is given by [73] as
 90 2
s t90  01 000
0min ¼ 00 þ xx0 C22 1 ðrÞ  C022 000 ðrÞ : ð4:176Þ
2E2

The stresses resulting from cracking can be calculated by solving the system of
coupled ODEs in Eq. (4.173) to get the perturbation functions, and then putting
them into Eqs. (4.167)–(4.168). The average of the stresses will yield the average
stiffness properties of the cracked laminate. For this model, the longitudinal
Young’s modulus normalized with its virgin state value is given by [73]

Ex 1
¼ ; ð4:177Þ
Ex0 E2 f ðrÞ
1 þ lE 1 r

where
ð þr
1
f ðrÞ ¼ ½ ðxÞ  1 ðxÞ d
x: ð4:178Þ
2 r

The normalized effective Poisson’s ratio for the cracked cross-ply laminate is
given by

t f ðrÞ
90
1  1  EE21
nxy h r
¼ : ð4:179Þ
n0xy E2 f ðrÞ
1 þ lE 1 r

4.6.5 Related works


There have been some further developments based on the variational analysis
procedure. The analysis of transverse cracks in cross-ply laminates undergoing
shear loading was covered by Hashin in his original paper [67]. Later he utilized
the variational approach to predict the thermal expansion coefficients for the
cracked cross-ply laminate [74]; this is covered later in Section 4.11. In another
paper [75], he also analyzed the case of orthogonally cracked cross-ply lamin-
ates. Another significant analysis of cross-ply laminates was carried out by
Kuriakose and Talreja [76] where they analyzed cross-ply laminates subjected
to bending moments.

4.6.6 Comparison between 1-D and 2-D stress-based models


Comparison of stress predictions between 1-D shear lag [27] and 2-D
variational analysis [67] for a [0/902]s carbon epoxy laminate are shown in
102 Micro-damage mechanics

Figure 4.15(a)–(b). The laminate is subjected to an applied mechanical stress


of 100 MPa, and a temperature change of DT = –125 C. The normalized
crack spacing is r = 4.
As seen in Figure 4.15(b), s90 xx ¼ 0 at x¼ r as the crack surfaces are traction
free and the transverse plies carry no axial load at the crack planes. As we move
away from the crack surfaces, s90 xx increases as stress is transferred back into the
90 -layer from the 0 -layer. This stress reaches a maximum midway between two
transverse cracks and its value depends on the crack spacing. When the cracks are
sufficiently far apart (i.e., no interaction between adjacent cracks), s90 xx will reach
90 90
sxx0 at this position (x = 0). The distributions of sxx are qualitatively similar for
shear lag and variational models. However, the maximum value of s90 xx is different
in the two cases. Also, variational predictions correctly show that this stress has
zero slope at crack surfaces (x = r), while the shear lag models display a
nonzero slope.
Shear lag analysis predicts a maximum value for the interfacial shear stress
ti ¼ s90
xz ðz ¼ t90 Þ at the crack surfaces, in violation of the boundary conditions that
require ti = 0 at crack planes (Figure 4.15(a)). In addition, it cannot provide the
transverse normal stress s90 zz . The one-dimensional nature of the shear lag models
results in no distinction between [0/90]s and [90/0]s laminates, while experimental
observations show differences in the cracking behavior and resultant stiffness
degradation in the two cases. These differences are appropriately predicted by
the variational solution shown for a [902/0]s laminate in Figure 4.15(c), where the
stress variations shown are different from the [0/902]s case. The interfacial trans-
verse stress s90zz near crack surfaces is compressive in [0/902]s while it is tensile in
[902/0]s laminate. This suggests that cracks in a [902/0]s laminate will tend to
promote mode I delamination.
In an alternative way, Varna and coauthors [77–79] investigated shear lag and
variational methods from the viewpoint of average crack opening displacement
(COD). They noted that the basic difference between different models is how they
model the average COD, which is defined as
ð t90
1
u ¼ uðzÞ dz : ð4:180Þ
t90 0

Consider transverse cracking in a symmetric and balanced [S/90n]s laminate,


consisting of two balanced sublaminates as outer layers and 90 layers in the
middle. The longitudinal modulus and Poisson’s ratio for the laminate are
given by

s0 eSyy
Exx ¼ ; n xy ¼  ; ð4:181Þ
eSxx eSxx

where s0 is the applied stress, the overbars denote volume averages, and
superscript S denotes sublaminate. The average COD for this laminate is given by
4.6 2-D stress analysis: variational methods 103

 
u ¼ l eSxx  e90
xx : ð4:182Þ

Following [79], the stiffness properties for the cracked laminate can be shown to
be given by

Exx 1 uxy 1  crR l
¼  ; ¼  ; ð4:183Þ
Exx0 1 þ arR l uxy0 1  arR l

where r = 1/2l is the crack density; a, c, and g are known functions of laminate
material and thickness; RðlÞ is the average stress perturbation function;
and l ¼ l=t90 is the half-crack spacing normalized with the 90 -layer thickness.
Different approaches model RðlÞ differently:
1. Shear lag models:

2
RðlÞ ¼ tanhðblÞ; ð4:184Þ
b

where b is the shear lag parameter. Different shear lag models use different
definitions for this parameter [78].
2. Variational methods:

4a1 a2 coshð2a1 lÞ  cosð2a2 lÞ


RðlÞ ¼ 2 ; ð4:185Þ
a1 þ a22 a2 sinhð2a1 lÞ þ a1 sinð2a2 lÞ

where the constants a1, a2 are defined as per model.


The comparisons of different models for [S/904]s laminates, as performed
by Joffe and Varna [79] are shown in Figures 4.16–4.19 for different values
of y. For the 2-D-0 model, the expression for the stress perturbation function
is the same as for Hashin’s variational analysis, except that the coefficients are
now given as [79]

1 1
C00 ¼ þ I1 ;
E2 ESx
 
n32 1  nS
C02 ¼ 2 þ A  2 xzS I4 ; ð4:186Þ
E2 6 Ex
1 1
C11 ¼ þ I3 :
2G23 GSxz

This comparison confirms that variational analyses consistently provide a


lower bound to stiffness properties and are significantly better than shear lag
models. The variational analysis predictions are also very close to FE
predictions.
104 Micro-damage mechanics

(a) [02/904]s
1.0

0.9

0.8
Ex/Ex0

0.7
Sh.L.-2
0.6 Hashin
FE
0.5 2D-0
Experiment
0.4
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
Crack density (cr/mm)

(b) [02/904]s
1.0
Sh.L.-2
Hashin
0.9 FE
2D-0
0.8 Experiment
Vxy/Vxy0

0.7

0.6

0.5

0.4
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
Crack density (cr/mm)

Figure 4.16. Variation of elastic stiffness moduli with crack density for a [02/904]s glass/epoxy
laminate: (a) normalized Young’s modulus; (b) normalized Poisson’s ratio versus crack
density. Reprinted, with kind permission, from Compos Sci Technol, Vol. 59, R. Joffe and
J. Varna, Analytical modeling of stiffness reduction in symmetric and balanced laminates
due to cracks in 90 degrees layers, pp. 1641–52, copyright Elsevier (1999).

4.7 Generalized plain strain analysis – McCartney’s model

These models may be seen as a further development of stress analysis-based


damage models for cracked laminates. The cracked cross-ply laminates, symmetric
about the laminate mid-plane, were considered in a generalized plane strain
condition, thereby reducing the directional dependence of essentially a 3-D stress
field in the cracked laminate. Assuming a regular array of fully grown parallel
cracks in 90 -ply, McCartney [80] developed a theory of stress transfer between
0 - and 90 -plies while retaining all relevant stress and strain components.
The equilibrium equations, as in Hashin’s formulation, are given by Eq. (4.112).
Assuming lamina properties to be transversely isotropic (or orthotropic), the
4.7 Generalized plain strain analysis – McCartney’s model 105

(a) [+_15/904]s
1.0

0.9

0.8
Ex /Ex0

0.7
Sh.L.-2
0.6 Hashin
FE
0.5 2D-0
Experiment
0.4
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
Crack density (cr/mm)

(b) [±15/904]s
1.0
Sh.L.-2
Hashin
0.9
FE
2D-0
0.8 Experiment
Vxy /Vxy0

0.7

0.6

0.5

0.4
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
Crack density (cr/mm)

Figure 4.17. Variation of elastic stiffness moduli with crack density for a [ 15/904]s glass/
epoxy laminate: (a) normalized Young’s modulus; (b) normalized Poisson’s ratio versus
crack density. Reprinted, with kind permission, from Compos Sci Technol, Vol. 59, R. Joffe
and J. Varna, Analytical modeling of stiffness reduction in symmetric and balanced
laminates due to cracks in 90 degrees layers, pp. 1641–52, copyright Elsevier (1999).

thermomechanical constitutive relations for 0 - and 90 -plies in material (lamina)


coordinate system can be written as
sm nm m
21 s22 nm m
31 s33
em
11 ¼
11
m  m 
m
m þ a11 T;
E11 E22 E33
ð4:187Þ
nm sm sm nm m
32 s33
em
22 ¼  12 m 11 þ 22 m 
m
m þ a22 T;
E11 E22 E33

nm m
13 s11 nm m
23 s22 sm33
em
33 ¼   þ m
m þ a33 T;
Em11 E m
22 E 33
sm sm sm
em
12 ¼ 12
; e m
13 ¼ 13
; em
23 ¼ 23
;
2Gm12 2Gm 13 2Gm 23
106 Micro-damage mechanics

(a) Shear lag model: [±30/904]s


1.0
Experiment
0.9
1
0.8
Ex /Ex0

3
0.7
2
0.6

0.5

0.4
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
Crack density (cr/mm)

(b) [±30/904]s
1.0

0.9

0.8
Vxy /Vxy0

0.7

0.6 Sh.L.-2
Hashin
0.5 FE
2D-0
Experiment
0.4
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
Crack density (cr/mm)

Figure 4.18. Variation of elastic stiffness moduli with crack density for a [ 30/904]s glass/
epoxy laminate: (a) normalized Young’s modulus; (b) normalized Poisson’s ratio
versus crack density. Reprinted, with kind permission, from Compos Sci Technol, Vol. 59,
R. Joffe and J. Varna, Analytical modeling of stiffness reduction in symmetric and
balanced laminates due to cracks in 90 degrees layers, pp. 1641–52, copyright Elsevier
(1999).

where m = 0, 90 for 0 -ply and 90 -ply, respectively, and Eii, Gij, nij; i, j = 1, 2, 3
represent the Young’s modulus, shear modulus, and Poisson’s ratio, respectively,
in corresponding directions and planes. Following the plane strain assumption,
the displacement field for the damaged laminate has the following form

um ¼ um ðx; zÞ; nm ¼ ecT y; wm ¼ wm ðx; zÞ; ð4:188Þ


where ecT is the uniform strain in the cracked laminate along the transverse (y)
direction. This representation of the displacement field is valid if ply cracks form
in planes normal to the x-direction, and are well away from the laminate edges.
4.7 Generalized plain strain analysis – McCartney’s model 107

(a) [±40/904]s
1.0
Sh.L.-2
Hashin
0.9
FE
2D-0
0.8 Experiment
Ex /Ex0

0.7

0.6

0.5

0.4
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
Crack density (cr/mm)

(b) [±40/904]s
1.0

0.9

0.8
Vxy / Vxy0

0.7

0.6 Sh.L.-2
Hashin
0.5 FE
2D-0
Experiment
0.4
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
Crack density (cr/mm)

Figure 4.19. Variation of elastic stiffness moduli with crack density for a [ 40/904]s glass/
epoxy laminate: (a) normalized Young’s modulus; (b) normalized Poisson’s ratio versus
crack density. Reprinted, with kind permission, from Compos Sci Technol, Vol. 59, R. Joffe
and J. Varna, Analytical modeling of stiffness reduction in symmetric and balanced
laminates due to cracks in 90 degrees layers, pp. 1641–52, copyright Elsevier (1999).

Now, if we additionally assume that the longitudinal stress components in both


the 0 - and 90 -plies are independent of the transverse coordinate z, the relevant
stress components can be written in the following functional form

s0xx ¼ s0xx0 þ Cð xÞ; s90 90


xx ¼ sxx0  lCð xÞ; ð4:189Þ
where
 

n012 E022 0 0
s0xx0 ¼ Q011 ec þ 0  a11 T ; s90
0 90 0 c 90
xx0 ¼ Q11 ec þ n12 eT  a22 T ; ð4:190Þ
E11
where ec is the average longitudinal strain applied to the laminate, and Q011 and Q90
11
are the reduced stiffnesses for the 0 - and 90 -plies, respectively, given by
108 Micro-damage mechanics

Em
Qm
11 ¼
11
m ; m ¼ 0; 90 ð4:191Þ
1  nm12 n21
00 0 90
and a 11 and a 22 are given as

0
0 E022 0 0
a11 ¼ a011 þ n012 a ; a2290
¼ a90 0 90
11 þ n12 a11 : ð4:192Þ
E011 22

To fully represent the stress state in the cracked laminate, the basic task is to
determine the unknown function C(x) appearing in Eq. (4.189). The boundary
conditions for the stress field are: symmetry about the mid-plane, symmetry about
the plane parallel to cracks lying midway between two adjacent cracks, stress
continuity at the interface between the 0 - and 90 -plies, and zero traction at crack
surfaces and the external surfaces of the outer plies (similar to Hashin’s analysis,
see Eq. (4.120)). Additionally, by force balance we have
ð t90 ðh
sc h ¼ s90
xx dz þ s0xx dz : ð4:193Þ
0 t90

Also, since there is no load applied on the laminate in the transverse direction, the
average transverse stress must add to zero, i.e.,
ð t90 ð l ðh ðl
s90
yy dxdz þ s90
yy dxdz ¼ 0 : ð4:194Þ
0 0 t90 0

The stress field in the cracked laminate can be obtained by substituting Eq. (4.189)
into the equilibrium equations, Eq. (4.112), and integrating and applying bound-
ary conditions. Thus, the transverse normal and shear stresses in the laminate are
given by
s0xz ¼ C0 ð xÞ½ð1 þ lÞt90  z; s90 0
xz ¼ lC ð xÞz;
1 1 00   ð4:195Þ
s0zz ¼ C00 ð xÞ½ð1 þ lÞt90  z2 ; s90 2 2
zz ¼ C ð xÞ ð1 þ lÞt90  z :
2 2
It can be observed that these relations are quite similar in form to those obtained
by Hashin (Eqs. (4.121) and (4.122)). In fact, they are identical when
CðxÞ ¼ ð1=lÞs90 xx fðxÞ. The complete stress field is given by combining Eqs.
(4.189) and (4.195).
The displacement field, given by Eq. (4.188), can be obtained by using the
stress field into constitutive relations, Eq. (4.187), and integrating. However, the
stress–strain relations must be satisfied in an average sense. Hence, the stress
and strain fields are averaged over each ply thickness. These averaged relations
lead to the following fourth-order differential equation for the unknown func-
tion C(x)
FCIV ð xÞ  GC00 ð xÞ þ HCð xÞ ¼ 0; ð4:196Þ
where the coefficients F, G, H are given by
4.7 Generalized plain strain analysis – McCartney’s model 109

"  #
1 2 t90 t90 2 5 t90 15
F¼ þ þ þ ;
20E022 15E9022 t0 t0 2 t0 8
  
1 1 1 t90 n012 n90 23 t90 ð4:197Þ
G¼ þ þ  2 þ3 ;
3 G012 G023 t0 E011 E90 22 t0
1 t90 1
H¼ 0 þ :
E11 t0 E022
When the inner and outer plies are made of the same material the differential
equation, Eq. (4.196), has the same form as the Euler–Lagrange equation derived
by Hashin. It is worth noting that the Reissner energy functional has a stationary
value when the constitutive relations and equilibrium equations are satisfied in
an average sense. Thus, McCartney’s approach is quite analogous to Hashin’s
approach: the former uses a displacement formulation (equivalent to minimiza-
tion of Reissner’s energy functional), while the latter uses a stress formulation
(equivalent to minimization of complementary energy functional).
For a cracked cross-ply laminate, McCartney’s 2-D approach yields the
following relations for the effective longitudinal modulus
Exx0
Exx ¼ ; ð4:198Þ
t90 E022
1þ 
l E011
where
 
2pq 2 pl 2 ql
¼ 2 þ cosh  cosh ;
p þ q2 t0 t0
1 pl pl ql ql
¼ q sinh cosh þ p sin cos ;
 t0 t0 t0 t0
rffiffiffiffiffiffiffiffiffiffi rffiffiffiffiffiffiffiffiffiffiffiffiffi ð4:199Þ
rþs jr  sj
p¼ ; q¼ ;
2 2
rffiffiffiffi
G H
r¼ > 0; s¼ > 0:
2F F
If the plies are thick, the averaging of stresses and strains is not advisable as it can result
in inaccurate predictions. To improve the approach, McCartney [81–86] used the so-
called “ply-refinement” procedure, in which each ply is divided thickness-wise into N
segments, and the resulting stress–strain relations are satisfied, in an average sense,
over each ply segment. Consequently, the unknown functions Ci(x) for the ith ply
segment (i = 1,2, . . . , N), satisfy the following N homogeneous differential equations

X
N X
N
00
X
N
Fij CIV
i ð xÞ Gij Ci ð xÞ þ Hij Ci ð xÞ ¼ 0; j ¼ 1; 2; : : : ; N; ð4:200Þ
i¼1 i¼1 i¼1

where the coefficients Fij, Gij, and Hij are calculated numerically using a suitable
ODE solver. McCartney [82, 84, 86] has later generalized this approach for multi-
layered cross-ply laminates, and also for triaxial loading.
110 Micro-damage mechanics

McCartney [82, 84, 86] has also shown that for cracked cross-ply laminates, the
moduli are interrelated. Similar interrelationships among the Poisson’s ratio,
transverse modulus, and longitudinal modulus have also been independently
derived by Nairn [87]. If we define a damage parameter or normalized stiffness
parameter as
1 1
D¼  ; ð4:201Þ
Exx Exx0
then other moduli for the cracked cross-ply laminate are interrelated as
nxy0 nxy 1
 ¼ k1 D; Eyy  E1 ¼ k12 D;
Exx0 Exx yy0
nxz0 nxz 1 1
 ¼ k2 D; Ezz  Ezz0 ¼ k22 D; ð4:202Þ
Exx0 Exx
nyz0 nyz
 ¼ k1 k2 D;
Eyy0 Eyy

and
axx  axx0 ¼ k3 D; ayy  ayy0 ¼ k1 k3 D; azz  azz0 ¼ k2 k3 D ; ð4:203Þ

where
 
Exx0 B  nxy0 Eyy0 =Exx0
k1 ¼ ;
Eyy0 1  nxy0 B  
Exx0 A  nxz0  nxz0 Exx0 =Eyy0 B
k2 ¼ ; ð4:204Þ
 1  nxy0 B 
Exx0 sxx0 þ Bayy0  C
k3 ¼ ;
1  nxy0 B
with
n13 n23 n12
A¼ þ ; B ¼ n12 ; C ¼ a22 þ n12 a11 : ð4:205Þ
E22 E11
These interrelationships can in principle reduce the burden of evaluating change in
each modulus subsequent to matrix cracking. Instead, the evaluation of degrad-
ation in longitudinal modulus is sufficient for predicting other moduli. However,
this result has yet to be verified experimentally.

4.8 COD-based methods

One way to view the elastic response changes caused by the presence of cracks in a
medium is to consider the additional overall (global) strain of the RVE contrib-
uted by the crack surface displacements of the individual cracks within the RVE.
Contrarily, if none of the cracks in the RVE conducts surface displacements, the
overall elastic response of the RVE will not change. This observation motivates
the focus on crack surface displacements. Although generally these displacements
4.8 COD-based methods 111

tk lk

k+1
k

lk

Figure 4.20. A general 3-D laminate with cracks in some off-axis layers.

can be described as crack opening displacement (COD) and crack sliding displace-
ment (CSD), one commonly refers only to COD. We will in the following discuss
stiffness relations in terms of average COD that can be calculated either analytic-
ally [88–96] or numerically [97–101].

4.8.1 3-D laminate theory: Gudmundson’s model


Consider a general three-dimensional laminate consisting of N plies (see
Figure 4.20). Each ply is defined by its material properties, layup angle, and
thickness. For uncracked laminate the global average stresses sij and strains eij
are defined as [102]

X
N X
N
ij ¼
s Vk skij ; eij ¼ Vk ekij ; ð4:206Þ
k¼1 k¼1

where skij ; ekij are the average stresses and strains, respectively, in the kth ply (k =
1, 2, . . ., N), and Vk stands for the volume fraction of the kth ply such that
PN
k¼1 Vk ¼ 1. Note that for uncracked laminates the ply average stresses and
strains are the same as the individual (constant) ply stresses. For the sake of
simplicity, we will denote the tensors with a . Now, let us partition the stresses,
strains, and thermal expansion coefficients into in-plane and out-of-plane parts
     
s I e
I
sI
s ¼ ; e ¼ ;s ¼  ; ð4:207Þ
 s
 O e
 O  s
 O

where
0 1 0 1 0 1
sxx exx axx
s
I
¼ @ syy A; e I ¼ @ eyy A; 
a I ¼ @ ayy A; ð4:208Þ
sxy 2exy 2axy

denote the in-plane stresses, strains, and thermal expansion coefficients, and
0 1 0 1 0 1
szz ezz azz
s
O
¼ @ sxz A; e O ¼ @ 2exz A; 
aO ¼ @ 2axz A; ð4:209Þ
syz 2eyz 2ayz
112 Micro-damage mechanics

denote the out-of-plane stresses, strains, and thermal expansion coefficients. The consti-
tutive law between global average stress tensor and average strain tensor is given by

e ¼ Ss
 
þ
aT

2 3
  SII 
SIO
     ð4:210Þ
e 6 7 sI aI
) I ¼ 4  T 5  þ  T;
e O SIO SOO  O
s a O
 

where S is the compliance tensor, and the superscript T represents the trans-

pose. Similarly, the relation between ply stress tensor and ply strain tensor can
be written as
 
k k
e k ¼ S s 
 s

kðrÞ
þ
a T


0 1 2 3
ek Sk
I
S k 0s k
IO s kðrÞ
1 0 1
sIk ð4:211Þ
I 6 7 I I
A þ @ AT;
) @ k A ¼ 4  T 5@ k k
eO Sk Sk s
O
s
O
IO  OO

kðrÞ
where s
denote residual stresses present due to reasons other than thermal
mismatch, such as chemical shrinkage during the manufacturing process. The
global average in-plane and out-of-plane residual stresses vanish due to equilib-
rium. The compatibility and equilibrium conditions in the laminate give us

ek
I
¼ e I ; s
O
k
¼s
O
: ð4:212Þ

Substituting Eq. (4.212) into Eq. (4.210), one obtains


 1  
k k k k kðrÞ
s
I
¼ S II e  S IO s
I Oa I T þ s
I
: ð4:213Þ
  

Using Eq. (4.206), and using the fact that the volume average of residual stresses
vanish, the above equation can be rewritten as

e ¼
I
SII s
I
þ
SIO s
O
þ
aI T; ð4:214Þ

where
"  1 #1
P
N
S ¼
II
Vk SIIk ;

k¼1
"  1 #
PN
S ¼
IO
SII Vk S IIk S IOk ; ð4:215Þ
 
" k¼1   #
1
PN
k k
a ¼
I
SII Vk S II a :
I

k¼1
4.8 COD-based methods 113

In a similar way, substituting Eq. (4.212) into Eq. (4.211), and applying Eq.
(4.206), we get
 T
e ¼ 
O
SIO sI
þSOO s
O
þaO T; ð4:216Þ

where
 T  1 "
T  1 # 
k
P
N
SOO ¼  SIO SII SIO þ Vk S OO  S IOk Sk Sk ;
      II  IO
k¼1
 T  1 "  T   ð4:217Þ
P
N
k k k 1 k
a
O
¼ SIO SII a
I
þ V a
k O  S S Þ a :
   IO  II I
k¼1

Eqs. (4.215) and (4.217) relate the effective laminate properties to the local
ply level properties. The above equations form the basis of a 3-D laminate
theory. The corresponding relations for a 2-D laminate theory are obtained
by considering only the in-plane tensors. To obtain the effective laminate
properties for a cracked laminate, define the nondimensional crack density in
the kth ply as
tk
rk ¼ ; ð4:218Þ
lk
where tk and lk denote the thickness and average crack spacing, respectively, in the
kth ply. The process of transverse cracking reduces the elastic energy of the
laminate and the change in elastic energy is associated with the release of tractions
on the crack surfaces. It was shown by Gudmundson and Ostlund [92] that the
coupling terms between the energy of the uncracked laminate and the change in
elastic energy due to matrix cracking vanish (see also the proof by Hashin [67],
which is described earlier in the chapter, Eq. (4.135)).
It is noted that for a cracked laminate, the effective strains are different from the
average strains, whereas there is no distinction between global effective and
average stresses. The effective strains are the strains measured on a global scale,
whereas average strains come from averaging strains in individual plies over the
RVE. The difference between the effective strains and average strains is equal to
strain increment caused by crack opening displacements. Thus, the average
stresses are defined as (see Eq. (4.206))

ðaÞ
X
N
kðaÞ
ij ¼
s Vk sij ; ð4:219Þ
k¼1

where the superscript (a) denotes average variables. The global effective strains are
defined as
ð
1  
eðeÞ
ij ¼ ui nj þ uj ni d ; ð4:220Þ
2V out
114 Micro-damage mechanics

where ui, i = 1, 2, 3 indicates the displacement vector, ni is the unit normal vector
on the outer boundary surface Gout of the representative volume V, and the
superscript (e) denotes the effective variables. In the same way the effective ply
strains can be defined as [102]
ð

kðeÞ 1
eij ¼ k uki nkj þ ukj nki d ; ð4:221Þ
2V kout

where Vk is the volume of the ply k and the surface integral is performed on the
outer boundary of ply k. Obviously,
ð X
N  ð

1   1 k k k k
ui n j þ u j n i d ¼ Vk u i nj þ u j n i d : ð4:222Þ
2V out
k¼1
2V k kout

Hence, from Eqs. (4.220)–(4.222), one gets

X
N
eðeÞ
ij ¼
kðeÞ
Vk eij : ð4:223Þ
k¼1

Applying the divergence theorem on Eq. (4.221), we get


ð
ð

kðeÞ 1 k k 1
eij ¼ k ui;j þ uj;i d  k uki nkj þ ukj nki d ; ð4:224Þ
2V V k 2V kc

where the first integral is equal to the average ply strain, given by
ð

ðaÞ 1
eij ¼ k uki;j þ ukj;i d ; ð4:225Þ
2V V k

and the second integral is the strain increment due to matrix cracks ekij is
given by
ð

k 1 rk k k

eij ¼  k uki nkj þ ukj nki d ¼ uki nkj ;


ui nj þ 
 ð4:226Þ
2V kc 2tk
where
ð tk
ð
1 kðþÞ kðÞ 1 tk
uki
 ¼ ui  ui dtk ¼ uki dtk
 ð4:227Þ
tk 0 tk 0
is the average crack opening displacement for ply k. Thus, we can write effective
ply strains as

kðeÞ kðaÞ
eij ¼ eij þ ekij : ð4:228Þ

Finally, after applying the proper boundary conditions, Gudmundson and


Zang [89] arrive at the following expressions for effective properties of cracked
laminate
4.8 COD-based methods 115

" 1  T N #1


P
N P
S ¼ SII  nk rk Ak bki Ak ;
IIðcÞ   
k¼1 i¼1
" 1  T N #1
P
N P
S ¼S
IOðcÞ IIðcÞ 
SII SIO þ nk rk A
 
k
bki B

i
;
k¼1 i¼1
 T  1  T  1  T N
PN
k k
k P ki i
S ¼ SIOðcÞ S S  SIO SII SIO þSOO þ n r B b B;
OOðcÞ  IIðcÞ IOðcÞ     
k¼1 i¼1
  T N
PN P ki i
a ¼a þSIIðcÞ nk rk Ak
IðcÞ I 
b C;
  
 k¼1 T  i¼1
1   N   N
P k k k TP
a ¼a
OðcÞ O
þ S S a a
IðcÞ I
þ n r B bki Ci ;
IOðcÞ IIðcÞ  
k¼1 i¼1
ð4:229Þ

where matrices SII ; SIO ; SOO ; SII ; 


aI ; and 
aO are given in Eqs. (4.215) and (4.217),
   
b is the matrix containing average crack opening displacements (COD), and

matrices A; B; C are given by
  
 1   
A
k
¼ N kI 
S kII ; B k ¼ N Ok  N kI SkII Þ1 
S kIO ; C k ¼ Ak  ak ;
aI  ð4:230Þ
      

where the matrices N kI and N kO represent the unit normal vector on the crack
 
surfaces in ply k, i.e.,
2 k 3 2 3
n1 0 nk2 0 0 0
k 4
N I
¼ 0 nk2 nk1 5 N Ok ¼ 4 0 0 0 5: ð4:231Þ
 
0 0 0 0 nk1 nk2

The expressions in Eq. (4.229) for stiffness of the cracked laminates are exact
assuming the homogenization procedure. However, the main problem is the
determination of average COD for the cracked laminate. The heterogeneity in
the composite laminates and constraint effects on crack surfaces from the sur-
rounding uncracked plies make it impossible to derive exact analytical solutions.
Gudmundson and coworkers, however, made the following assumptions in order
to evaluate the average crack opening displacements:
1. The surface displacements of a ply crack in a finite-thickness laminate are
equal to those of a crack in an infinite, homogeneous transversely isotropic
medium. The stress intensity factors for an infinite row of equidistant cracks
in an infinite homogeneous isotropic medium under the action of uniform
tractions on crack surfaces, given by Benthem and Koiter [104] and Tada et al.
[105], are assumed to hold for the current case of transversely isotropic
medium.
2. There is no effect of orientation of a cracked ply. This means that the matrices bki

in Eq. (4.229) may not be accurate for off-axis cracks in laminates.
116 Micro-damage mechanics

Table 4.2 Numerical parameters used in calculation of COD matrix coefficients bk1 ; bk2 ; and bk3

j a b c
1 0.63666 0.63666 0.25256
2 0.51806 –0.08945 0.27079
3 0.51695 0.15653 –0.49814
4 –1.04897 0.13964 8.62962
5 8.95572 0.16463 –51.2466
6 –33.0944 0.06661 180.9631
7 74.32002 0.54819 –374.298
8 –103.064 –1.07983 449.5947
9 73.60337 0.45704 –286.51
10 –20.3433 – 73.84223

3. The crack density is low (r  1).


4. There is no coupling between crack opening displacements of different plies.
Hence the matrices bki are diagonal.

With these assumptions, the matrices bki are given by

bki ¼ 0; for all k 6¼ i;
 
2 3
bk1 0 0 ð4:232Þ
or bkk ¼ 4 0 bk2 0 5;

0 0 bk3
where bk1 ; bk2 ; and bk3 are determined by Gudmundson using a numerical integra-
tion and are given by
h k
i
4 ln cosh pr2
bk1 ¼ g1 ;
p ð rk Þ 2
p X 10
aj
bk2 ¼ g2 ; ð4:233Þ
2 j¼1 ð1 þ rk Þj

p X 9
bj
bk3 ¼ g3 ;
2 j¼1 ð1 þ rk Þj2
for cracks in an internal ply, where
1 1  u12 u21
g1 ¼ ; g2 ¼ g3 ¼ ; ð4:234Þ
2G12 E2
and aj and bj are numerical parameters given in Table 4.2. For surface cracks (or
cracks in external plies), we have
  
kðsÞ 8 ln cosh prk
b1 ¼ g1 ;
p ð2rk Þ2
" # ð4:235Þ
kðsÞ 2 p
X 10
cj
b2 ¼ 2ð1:12Þ g2 ;
2 j¼1 ð1 þ rk Þj
4.8 COD-based methods 117

where cj are another set of numerical parameters also given in Table 4.2.
It is clear despite the assumptions used in COD calculation, the above approach
is quite complex and difficult to implement in a numerical scheme. Similar
relations are also available for the combined extension and bending loading
scenario [94, 95].

4.8.2 Lundmark–Varna model


Lundmark and Varna [98, 99, 106, 107] have recently derived effective properties
using a homogenization very similar to that in Gudmundson’s approach. How-
ever, their relationships are simpler and they use the crack surface displacements
numerically obtained from FEM simulations, which makes their model much
more accurate. However, their model has been checked against experimental
data only for cross-ply laminates. Here, we provide the main relations derived
by them. Accordingly, the average strains in a ply, analogous to (4.228), are
defined as
a LAM
eij k ¼ eij þ b  ; ð4:236Þ
ij k

where the superscript a means “average” and


8 9a 8 9LAM 8 9
 =
a < e11 = LAM < e11 = < b 11
eij k ¼ ¼  
e ; eij e ; b ij k ¼ b ; ð4:237Þ
: 22 ; : 22 ; : 22 ;
2e12 k 2e12 2b12 k

where b 
ij k represents the Vakulenko–Kachanov tensor, defined by
ð

1

bij k ¼ k uki nkj þ ukj nki d ; ð4:238Þ
2V kc


which is same as ekij in Eq. (4.226). In terms of crack surface displacements, b
ij k
for a cracked ply is derived as
 h i
 ¼ rkn ½ A ½U  ½ A Q fe0 gLAM  fa0 g T ;
b ð4:239Þ
k k k k k k
E2
where [A]k is the transformation matrix for ply k and [U]k is the displacement
matrix given by
2 3
0 0 0
½U k ¼ 2 ¼ 4 0 uk2an 0 5; ð4:240Þ
E2 k
0 0 G12 u1an

where uk1an and uk2an are the normalized average crack face sliding and opening
displacements, respectively, and are given by

G12 E2
uk1an ¼ uk1a ; uk2an ¼ uk2a k ; ð4:241Þ
tk sk120 tk s20
118 Micro-damage mechanics

with uk1a and uk2a being the average crack surface displacements, defined as
tk tk
ð2 ð2
1 1
uk1a ¼ u1 ðzÞ dz; uk2a ¼ u2 ðzÞ dz; ð4:242Þ
2tk 2tk
t t
 2k  2k

where Du1 and Du2 are the relative separation of the two crack faces along and
normal to the crack surface. Now, the average stress–strain relationships for the
kth ply in the global coordinate system are
  
gak ¼ Q k fegak fagk T :
fs ð4:243Þ

Since, the average stresses for the laminate remain equal to the applied stresses,
we have
X
N
tk
fsgLAM ¼ fs
 ga ¼ gak
fs ; ð4:244Þ
k¼1
H

where H is the total thickness of the laminate. Substituting Eqs. (4.236) and (4.239)
into Eq. (4.244), we obtain
h i 1X N  
fsgLAM ¼ ½Q0 LAM fegLAM  fe0 gLAM
th þ  ;
tk Q k b k
ð4:245Þ
H k¼1

where [Q0]LAM is the stiffness matrix for the undamaged laminate, and
P
fe0 gLAM
th ¼ ð1=H Þ Nk¼1 tk fagk T are the thermal strains in the undamaged
laminate. Substituting Eq. (4.239) into Eq. (4.245), the average thermomechanical
response of the damaged laminate is given by
h i
fsgLAM ¼ ½Q0 LAM fegLAM  fe0 gLAM
th

1 X N    h i ð4:246Þ
 rkn tk Q k ½ Ak ½U k ½ Ak Q k fe0 gLAM  fa0 gk T :
HE2 k¼1

For purely mechanical response, the stiffness matrix of the damaged laminate is
given by
!1
LAM 1 X N    
½Q ¼ ½I þ rkn tk Q k ½ Ak ½Uk ½ Ak Q k ½S0 LAM ½Q0 LAM ; ð4:247Þ
HE2 k¼1

where [S0]LAM = ([Q0]LAM)1 is the compliance tensor for the undamaged


laminate.
The only remaining unknown in this model is [U]k, Eq. (4.240). To determine the
crack surface displacements, Lundmark and Varna [98] suggest using actual FE
calculations on a unit cell of cracked laminate. For cross-ply laminates, they
carried out a set of such FE calculations for varying values of ply thickness and
4.9 Computational methods 119

(a) (b)
[30/–30/904]s [30/–30/904]s
1.0 1.0
+ experimental + experimental
model model
0.9 0.9

nxy /nxy0
Ex /Ex0

0.8 0.8

0.7 0.7

0.6 0.6
0 0.2 0.4 0.6 0 0.2 0.4 0.6
Crack density (cracks/mm) Crack density (cracks/mm)

Figure 4.21. Reductions in longitudinal modulus (a) and Poisson’s ratio (b) for a graphite/
epoxy [ 30/904]s laminate using the Lundmark–Varna model [98]. Reprinted, with kind
permission, from P. Lundmark and J. Varna, Int J Damage Mech, Vol. 14, pp. 235–59,
copyright # 2005 by Sage Publications.

stiffness, and by curve fitting they provided the following power law-type expres-
sions for the normalized average crack opening displacements
 p
E2
u2an ¼ A þ B s ; ð4:248Þ
Ex
where A, B, and p are constants that depend on the thickness ratio of cracked to
uncracked plies, and the type of crack (in an internal or external ply). For
example, for an internal crack in a GFRP [Sn/90m]s laminate,
 
t90  2ts
A ¼ 0:52; B ¼ 0:3075 þ 0:1652 ;
2ts
 2   ð4:249Þ
t90 t90
p ¼ 0:0307  0:0626 þ 0:7037;
2ts 2ts

where t90 and ts are the thicknesses of cracked 90 and each uncracked sublaminate
layer. Figure 4.21 shows comparison of the model with experimental data for a
[ 30/904]s laminate with cracks in the central 90 layer.

4.9 Computational methods

For simulating the effect of damage in composite laminates, in general,


many computational methods have been utilized, such as the finite element
method (FEM), the finite difference method (FDM), and the boundary element
method (BEM). For the particular problem of transverse cracking, the most
common numerical tool is FEM. Some simpler numerical tools have also been
devised, e.g., the finite strip method by Li et al. [108], and the layerwise theory of
Reddy [109]. In this subsection, we will briefly describe some of these
developments.
120 Micro-damage mechanics

4.9.1 Finite element method (FEM)


FEM is the most widely used numerical approach to analyze damage in composite
laminates. Its advantages are that it can model highly complex situations, and
provides quite accurate results. As can be seen from the preceding discussion, the
analytical models of cracking are severely limited in scope with respect to laminate
layup and loading scenario. Nonetheless, all numerical methods have a unique
limitation: every time the laminate geometry, loading, or material changes, simu-
lations need to be carried out afresh, which might involve new mesh generation,
and computation. Hence, FE modeling of cracked laminates is time consuming
and does not by itself provide much insight into the damage mechanisms. Even
with these limitations it can be successfully used for calibration/verification of
analytical models as well as for computation of parameters or constants useful
in analytical methods (for example, it can be used to evaluate constants used in
Talreja’s continuum damage model, see the next chapter for details). FE modeling
can also be utilized to simulate complex experimental situations, a task equivalent
to carrying out “numerical experiments,” and eliminating the need for cumber-
some experimentation whenever possible.
The first task in FE modeling is to define a geometrical model of the cracked
laminate. Mostly, a representative unit cell is developed assuming periodic array
of self-similar cracks (see Figure 4.7(a)). For cross-ply laminates, a 3-D repeating
unit can be reduced to a 2-D plane stress/strain model (see Figure 4.7(b)).
Furthermore, the symmetry of the resulting boundary value problem can
reduce the size of the modeled unit cell. For example, cracking in a [0/90]s
laminate can be modeled using a quarter of a representative unit cell as illus-
trated in Figure 4.22.
For multidirectional laminates with cracks in more than one orientation, how-
ever, a repeating unit cannot be defined uniquely because of differences in direc-
tionality and mutual spacing of cracks in different plies. A repeating unit cell can
be defined for up to two off-axis cracking modes with cracked surfaces represented
on nonorthogonal boundaries [110]. The RVE in such cases is therefore 3-D and
skewed (nonorthogonal in the x–y plane; z being the thickness direction).
A reduction to two dimensions is often not possible in such cases.
As is common in micromechanics, the displacement, strain, and stress fields
must be periodic across repeating unit cells. Thus the following periodic boundary
conditions [111] are applied on the FE model
 
@ui
ui ðxa þ xa Þ ¼ ui ðxa Þ þ xb ;
@xb
ð4:250Þ
eij ðxa þ xa Þ ¼ eij ðxa Þ;
sij ðxa þ xa Þ ¼ sij ðxa Þ;

where ui, h@ui/@ubi, and Dxb, i, b = 1,2,3 represent the displacements, the volume
averaged displacement gradients, and the vector of periodicity, respectively. For a
4.9 Computational methods 121

(a) z (b)

0˚ t0
h sc
t90
sc 90˚ x


2l

Figure 4.22. Building FE model for a cracked cross-ply laminate: (a) 2-D representative unit
cell; (b) FE model (1/4 cell).

skewed RVE, periodic BC might be complicated to enforce. For a detailed


discussion about skewed RVE, periodic BCs, and symmetries in composites, the
reader is referred to [110, 112–114].
Once the FE results are available, the overall stiffness properties of the damaged
laminate can be calculated by volume averaging of stresses and strains inside the
RVE. For example, for a cracked laminate undergoing in-plane loading, the in-
plane elasticity properties can be calculated by
     
hsxx i syy sxy eyy
Ex ¼ ; Ey ¼   ; Gxy ¼   ; nxy ¼  ; ð4:251Þ
hexx i eyy gxy hexx i

where <g> denotes the volume average of a field g. Important numerical


studies related to analysis of cracked composite laminates can be found in [110–
113, 115, 116].

4.9.2 Finite strip method


To enable predictions for laminated composites with arbitrary layups [110,
117, 118], Li et al. [108] devised an approximate numerical approach based on
the generalized plain strain formulation. Although there can be more than
one cracked ply, all the cracked plies have to be of the same orientation.
Hence, this method cannot be used for laminates with cracks in multiple
orientations.
Consider first an uncracked laminate. Given the generalized strains {e},
 T
feg ¼ exx0 ; eyy0 ; 2exy0 ; kxx ; kyy ; 2kxy ; ð4:252Þ

the displacement field based on the classical laminate theory without rigid body
displacements can be written as
u0 ¼ xexx0 þ yexy0 þ xzkxx þ yzkxy ;
v0 ¼ yeyy0 þ xexy0 þ yzkyy þ xzkxy ;
ð4:253Þ
x2 y2
w0 ¼  kxx  kyy  xykxy þ oðzÞ;
2 2
122 Micro-damage mechanics

where o is an arbitrary integration function which depends on boundary


conditions.
When transverse cracks appear, perturbations to this displacement field are
induced. These perturbations can simply be superimposed on the displacements due
to the linearity of the problem. Assuming sufficiently long cracks, the perturbations
are independent of y. Hence, the displacement field for a cracked laminate reads
u ¼ u0 þ U ðx; zÞ;
v ¼ v0 þ Vðx; zÞ; ð4:254Þ
w ¼ w0 þ Wðx; zÞ;
where U, V, and W denote changes due to the presence of cracks. The resulting
strains are calculated using infinitesimal deformation theory. Our objective then is
to solve for U, V, and W. The approach in the finite strip method is to divide the
planar region of a typical segment of the cracked laminate into a finite number of
strip elements parallel to the x-axis. In each element, M nodal lines are introduced
along which displacements are functions of x only and the displacements in the
strip elements are then interpolated by polynomials in the z-direction. For a
typical element, the displacement field can be expressed as
8 9e 8 9e
< U = X3 < Ui =
V ¼ Ni ðzÞ Vi ; ð4:255Þ
: ; : ;
oþW i¼1 Wi

where the superscript e denotes the element number, and Ni are shape functions,
the same as those used in the one-dimensional finite element analysis [119], which
are given by
1 1
N1 ¼ zð1  zÞ; N2 ¼ 1  z2 ; N1 ¼ zð1 þ zÞ; ð4:256Þ
2 2
with 1 z 1 being the nondimensional coordinate along the z-axis

X
3
z¼ Ni ðzÞzi ; ð4:257Þ
i¼1

where zi is the z-coordinate of the ith nodal line in the element. Going through the
usual FEM formulation, Li et al. arrived at the following set of differential
equations for the resulting variational problem
n o
n o
 ½K2  €
y þ ½K1   ½K1 T y_ þ ½K0 fyg ¼ fF0 g; ð4:258Þ

and a set of boundary conditions, where [K2], [K1], and [K0] are the matrices
defined in Li et al. [108] and
8 9
<U=
fyg ¼ V : ð4:259Þ
: ;
W
4.9 Computational methods 123

No direct solution to the differential Eq. (4.258) exists. An approximate solution


can be obtained by taking the nodal displacements as a Fourier series of unknown
constants, as

x XKn
kpx
Un ¼ Unh þ Unk sin ;
l k¼1 l
x XKn
kpx
Vn ¼ Vnh þ V k sin ; ð4:260Þ
l k¼1 n l
x
2 XKn
kpx
Wn ¼ Wnh þ Wn0 þ Unk cos ;
l k¼1
l

for n = 1,2, . . ., M, and the coefficients Unh ; Vnh ; Wnh ; Unk ; Vnk ; Vnk ; and Wn0 are
unknown constants to be determined, Kn is the order where the Fourier series is
truncated for the approximate solution, and M is the number of nodal lines in the
laminate. Substituting Eq. (4.260) into Eq. (4.258), and rearranging, the following
algebraic equations similar to FEM are obtained
½ Afg ¼ fg : ð4:261Þ
The solution of these algebraic equations along with the appropriate boundary
conditions gives the displacement field, from which the strain field can be
obtained.

4.9.3 Layerwise theory


The layerwise theory of Reddy [109] is motivated by a desire to develop a
computational model that is more efficient than the conventional 3-D finite
element models [119] and can incorporate damage effects such as transverse cracks
and delaminations in a layered medium. It is based on 3-D kinematics where the
displacement field within each layer is expanded using the Lagrange family of
finite elements. For a comparable mesh, the layerwise theory typically takes less
computational time as compared to the conventional 3-D FEM while providing
the same level of accuracy.
In the theory, the whole laminate is divided into a number of subdivisions
across its thickness. The displacement field in the laminate is written as
X
N
ui ðx; y; zÞ ¼ UiJ ðx; yÞJ ðzÞ; i ¼ 1; 2; 3; ð4:262Þ
J¼1

where N is the number of subdivisions through the thickness of the laminate and FJ
are global interpolation functions defined in terms of the Lagrange interpolation
functions associated with the layers connected to the Jth interface through the
laminate thickness, and UiJ are the nodal displacements. Independent interpolation
functions for u1, u2, and u3 can also be used whenever necessary. The strain field is
determined using the von Karman nonlinear theory. Then the governing equations
124 Micro-damage mechanics

for the nodal displacements are derived using the principle of virtual displacements
[120, 121]. The resulting system of partial differential equations can be converted to
a system of equations analogous to the FEM in a procedure similar to that followed
in previous subsection. Solution of this system gives us the nodal displacement,
from which strain and stress fields can be determined. Further details on the
approach and its implementation can be found in [122–127]. Recent papers by Na
and Reddy [128, 129] provide direct implementation of the layerwise theory for
transverse cracking and delaminations, respectively, in cross-ply laminates.

4.10 Other methods

There have been other developments in analyzing cracked laminates. The detailed
treatment is not covered here and we will highlight the important aspects of these
approaches. The readers are referred to the cited articles for further details. It
should be pointed out, however, that these approaches are mostly extensions of
the previously developed ideas, and do not really improve the predictions. More-
over, many of these approaches are complex, semi-analytical and thus difficult to
implement, and may sometimes need high computational times.
Close to the development of the variational analysis by Hashin, Aboudi and
coworkers [130, 131] developed a three-dimensional semi-analytical method. In
this approach, an approximate analytical solution for the displacement field is
sought using a series expansion in the form of Legendre polynomials.
In a somewhat similar way, Lee and Hong [132] developed a refinement of
the shear lag approach using a series polynomial expansion. The approach
accounted for the crack opening displacement, thermal stresses, and Poisson’s
effects. However, the improvement over the traditional shear lag methods was
insignificant.
Another similar effort by Gamby and Rebiere [133] used the transverse shear
stresses in the 0 - and 90 -layers as double Fourier series in x and z (longitudinal
and transverse coordinates). Vanishing tractions at the crack surfaces were used as
boundary conditions. To obtain the stresses and strains in the cracked laminate,
the use of constitutive equations and the equilibrium equations was made. Finally,
the coefficients of the Fourier series were derived using the principle of minimum
complementary energy.
Quite recently, Zhang et al. [134, 135] have developed a “state space method” to
analyze cracked off-axis laminates. In this approach they use displacement fields
and the stresses to be given by the Fourier series expansions. The coefficient terms
in expansions are obtained numerically using the equilibrium equations, boundary
conditions, and ply refinement.
Following on the lines of modeling of thin shells, Shoeppner and Pagano [136–
138] developed a method to model the thermoelastic response of flat laminated
composites. This approach employs the idea that the stress field in an axisym-
metric cylinder approaches that in a long flat coupon as the radius to thickness
4.11 Changes in thermal expansion coefficients 125

ratio approaches infinity. Thus, the stress field in a flat laminate is nearly the same
as in a large radius axisymmetric hollow layered cylinder model. To obtain the
stress fields, they used Reissner’s variational principle along with the equilibrium
equations. The predictions matched quite well with FE simulations. However, the
governing equations are very complex, need numerical tools to solve them,
and this limits the usability of the approach. The models of McCartney and
Shoeppner–Pagano are compared in [139].

4.11 Changes in thermal expansion coefficients

Hashin [74] and Nairn [140] have developed variational models to predict changes
in thermal expansion coefficients due to ply cracking. They can be derived in a
relatively straightforward way following Hashin’s treatment (refer to [74] for
details). The longitudinal expansion coefficient for a cracked cross-ply laminate
is given by
 
a11  a22 B0 90 
a90
xx ¼ a c þ 1 þ k f; ð4:263Þ
1þl C0 x
where ac is the longitudinal thermal expansion coefficient for an uncracked
composite laminate, and
ð
 1 l
f¼ fð xÞ dx;
2l l ð4:264Þ
n12
B 0 ¼  ð 1 þ lÞ ;
E1
Lundmark and Varna [98] also developed a model for thermal expansion coeffi-
cients. Following the analysis covered in Section 4.8.2 and noting that only
thermal loading is present, the strain field in the laminate is given by (setting
mechanical loads to zero in Eq. (4.245))

1X N  
 :
fegLAM ¼ fe0 gLAM
th ½S0 LAM tk Q k b k
ð4:265Þ
H k¼1

Substituting b from Eq. (4.239) and dividing by DT yields
k
!
LAM
XN
1X N
fag ¼ ½I þ tk rkn ½ Dk fagLAM  tk r ½ D fagk ; ð4:266Þ
k¼1
0
H k¼1 kn k

where
1    
½ Dk ¼ ½S0 LAM Q k ½T Tk ½U k ½T k Q k : ð4:267Þ
E2
The model predictions for longitudinal thermal expansion coefficient are plotted
in Figure 4.23 for a [0/90]s carbon/epoxy laminate. For comparison, experimental
and model data from Kim et al. [141] are also shown.
126 Micro-damage mechanics

[0/90]s
1.0
experimental, Kim et al. (2000)
model
0.8 model, Kim et al. (2000)
ax / ax0

0.6

0.4

0.2
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Crack density (cracks / mm)

Figure 4.23. Variation in longitudinal thermal expansion coefficient for a carbon/epoxy


[02/902]s laminate. The Lundmark–Varna model is compared with model and
experiment from Kim et al. [141]. Reprinted, with kind permission, from P. Lundmark and
J. Varna, Int J Damage Mech, Vol. 14, pp. 235–59, copyright # 2005 by Sage Publications.

4.12 Summary

This chapter has provided an exposition of the main concepts and methods related to
evaluating the effects of multiple cracking in composite materials on their deform-
ational response. Beginning with the early approaches, known as shear lag methods,
where one-dimensional stress analysis is used, and progressing to computational
methods that accurately determine the local stress fields, the range is covered as much
as possible. While the field is still evolving and more methods are appearing in the
literature, it is hoped that the treatments and discussions provided here are useful for
researchers in the field to gain an appreciation of this class of approaches. A key
consideration in selecting an approach is the purpose at hand. For material selection
purposes a quick assessment of the elastic modulus may be needed, in which case the
shear lag approach may be adequate, while for structural analysis purposes an
evaluation of all properties and the suitability of the approach as an integral part
of a structural analysis scheme would be the factors of consideration. In any case, the
micro-damage mechanics (MIDM) treated here is one part of the total damage
mechanics picture. The next chapter on macro-damage mechanics (MADM) pro-
vides another perspective on the complex problem of damage in composite materials.

References

1. Z. Hashin, Analysis of damage in composite materials. In Yielding, Damage, and Failure


of Anisotropic Solids, ed. J. P. Boehler. (London: Mechanical Engineering Publications,
1990), pp. 3–31.
2. S. Nemat-Nasser and M. Hori, Micromechanics: Overall Properties of Heterogeneous
Materials, 2nd edn. (Amsterdam: North Holland, 1999).
References 127

3. J. Aveston, G. A. Cooper, and A. Kelly, Single and multiple fracture. In The Properties
of Fiber Composites. (Surrey, UK: IPC Science and Technology Press, National Phys-
ical Laboratory, 1971), pp. 15–26.
4. A. Kelly, The 1995 Bakerian Lecture: composite materials. Phil Trans R Soc London
A, 354:1714 (1996), 1841–74.
5. G. A. Cooper and J. M. Sillwood, Multiple fracture in a steel reinforced epoxy resin
composite. J Mater Sci, 7:3 (1972), 325–33.
6. J. Aveston and A. Kelly, Theory of multiple fracture of fibrous composites. J Mater Sci,
8:3 (1973), 352–62.
7. S. W. Wang and A. Parvizi-Majidi, Experimental characterization of the tensile behav-
iour of Nicalon fibre-reinforced calcium aluminosilicate composites. J Mater Sci, 27:20
(1992), 5483–96.
8. J. E. Masters and K. L. Reifsnider, An investigation of cumulative damage develop-
ment in quasi-isotropic graphite/epoxy laminates. In Damage in Composite Materials,
ASTM STP 775, ed. K. L. Reifsnider. (Philadelphia, PA: ASTM, 1982), pp. 40–62.
9. J. Tong, F. J. Guild, S. L. Ogin, and P. A. Smith, On matrix crack growth in quasi-isotropic
laminates – I. Experimental investigation. Compos Sci Technol, 57:11 (1997), 1527–35.
10. D. T. G. Katerelos, P. Lundmark, J. Varna, and C. Galiotis, Analysis of matrix
cracking in GFRP laminates using Raman spectroscopy. Compos Sci Technol, 67:9
(2007), 1946–54.
11. T. Yokozeki, T. Aoki, and T. Ishikawa, Consecutive matrix cracking in contiguous
plies of composite laminates. Int J Solids Struct, 42:9–10 (2005), 2785–802.
12. T. Yokozeki, T. Aoki, T. Ogasawara, and T. Ishikawa, Effects of layup angle and ply
thickness on matrix crack interaction in contiguous plies of composite laminates.
Compos A, 36:9 (2005), 1229–35.
13. M. Kashtalyan and C. Soutis, Stiffness and fracture analysis of laminated composites
with off-axis ply matrix cracking. Compos A, 38:4 (2007), 1262–9.
14. M. Kashtalyan and C. Soutis, Modelling off-axis ply matrix cracking in
continuous fibre-reinforced polymer matrix composite laminates. J Mater Sci, 41:20
(2006), 6789–99.
15. L. N. McCartney and G. A. Schoeppner, Predicting the effect of non-uniform ply
cracking on the thermoelastic properties of cross-ply laminates. Compos Sci Technol,
62:14 (2002), 1841–56.
16. V. V. Silberschmidt, Effect of micro-randomness on macroscopic properties and frac-
ture of laminates. J Mater Sci, 41:20 (2006), 6768–76.
17. V. V. Silberschmidt, Matrix cracking in cross-ply laminates: effect of randomness.
Compos A, 36 (2005), 129–35.
18. L. E. Asp, L. A. Berglund, and R. Talreja, A criterion for crack initiation in glassy
polymers subjected to a composite-like stress state. Compos Sci Technol, 56:11 (1996),
1291–301.
19. L. E. Asp, L. A. Berglund, and R. Talreja, Effects of fiber and interphase on matrix-initiated
transverse failure in polymer composites. Compos Sci Technol, 56:6 (1996), 657–65.
20. L. E. Asp, L. A. Berglund, and R. Talreja, Prediction of matrix-initiated transverse
failure in polymer composites. Compos Sci Technol, 56:9 (1996), 1089–97.
21. H. S. Huang and R. Talreja, Numerical simulation of matrix micro-cracking in short
fiber reinforced polymer composites: initiation and propagation. Compos Sci Technol,
66:15 (2006), 2743–57.
128 Micro-damage mechanics

22. H. Huang and R. Talreja, Effects of void geometry on elastic properties of


unidirectional fiber reinforced composites. Compos Sci Technol, 65:13 (2005),
1964–81.
23. K. A. Chowdhury, R. Talreja, and A. A. Benzerga, Effects of manufacturing-induced
voids on local failure in polymer-based composites. J Eng Mater Tech, Trans ASME,
130:2 (2008), 021010.
24. K. A. Chowdhury, Damage initiation, progression and failure of polymer based
composites due to manufacturing induced defects. Ph.D. thesis, Texas A&M Univer-
sity, College Station, TX (2007).
25. H. L. Cox, The elasticity and strength of paper and other fibrous materials. Br J Appl
Phys, 3 (1952), 72–9.
26. J. A. Nairn, On the use of shear-lag methods for analysis of stress transfer unidirec-
tional composites. Mech Mater, 26:2 (1997), 63–80.
27. K. W. Garrett and J. E. Bailey, Multiple transverse fracture in 90 degrees cross-ply
laminates of a glass fiber-reinforced polyester. J Mater Sci, 12:1 (1977), 157–68.
28. A. Parvizi, K. W. Garrett, and J. E. Bailey, Constrained cracking in glass fiber-
reinforced epoxy cross-ply laminates. J Mater Sci, 13:1 (1978), 195–201.
29. P. W. Manders, T. W. Chou, F. R. Jones, and J. W. Rock, Statistical analysis of
multiple fracture in 0 /90 /0 glass fibre/epoxy resin laminates. J Mater Sci, 18:10
(1983), 2876–89.
30. A. L. Highsmith and K. L. Reifsnider, Stiffness-reduction mechanisms in composite
laminates. In Damage in Composite Materials, ASTM STP 775, ed. K. L. Reifsnider.
(Philadelphia, PA: ASTM, 1982), pp. 103–17.
31. J. W. Lee and I. M. Daniel, Progressive transverse cracking of crossply composite
laminates. J Compos Mater, 24:11 (1990), 1225–43.
32. D. L. Flaggs, Prediction of tensile matrix failure in composite laminates. J Compos
Mater, 19:1 (1985), 29–50.
33. S. C. Tan and R. J. Nuismer, A theory for progressive matrix cracking in composite
laminates. J Compos Mater, 23:10 (1989), 1029–47.
34. Y. M. Han and H. T. Hahn, Ply cracking and property degradations of symmetric balanced
laminates under general in-plane loading. Compos Sci Technol, 35:4 (1989), 377–97.
35. Y. M. Han, H. T. Hahn, and R. B. Croman, A simplified analysis of transverse ply
cracking in cross-ply laminates. Compos Sci Technol, 31:3 (1988), 165–77.
36. N. Laws and G. J. Dvorak, Progressive transverse cracking in composite laminates.
J Compos Mater, 22:10 (1988), 900–16.
37. H. L. McManus and J. R. Maddocks, On microcracking in composite laminates
under thermal and mechanical loading. Polymers & Polymer Composites, 4:5 (1996),
305–14.
38. S. G. Lim and C. S. Hong, Prediction of transverse cracking and stiffness reduction in
cross-ply laminate composites. J Compos Mater, 23 (1989), 695–713.
39. M. Caslini, C. Zanotti, and T. K. O’Brien, Fracture mechanics of matrix cracking and
delamination in glass/epoxy laminates. J Compos Tech Res, 9:4 (1987), 121–30.
40. J. A. Nairn and D. A. Mendels, On the use of planar shear-lag methods for stress-
transfer analysis of multilayered composites. Mech Mater, 33:6 (2001), 335–62.
41. J. M. Berthelot, Transverse cracking and delamination in cross-ply glass-fiber and
carbon-fiber reinforced plastic laminates: static and fatigue loading. Appl Mech Rev,
56:1 (2003), 111–47.
References 129

42. J. A. Nairn and S. Hu, Micromechanics of damage: A case study of matrix micro-
cracking. In Damage Mechanics of Composite Materials, ed. R. Talreja. (Amsterdam:
Elsevier, 1994), pp. 187–243.
43. A. Parvizi and J. E. Bailey, Multiple transverse cracking in glass-fiber epoxy cross-ply
laminates. J Mater Sci, 13:10 (1978), 2131–6.
44. G. J. Dvorak, N. Laws, and M. Heiazi, Analysis of progressive matrix cracking in
composite laminates – I. Thermoelastic properties of a ply with cracks. J Compos Mater,
19 (1985), 216–34.
45. H. Fukunaga, T.-W. Chou, P. W. M. Peters, and K. Schulte, Probabilistic failure strength
analysis of graphite/epoxy cross-ply laminates. J Compos Mater, 18:4 (1984), 339–56.
46. P. S. Steif, Parabolic shear lag analysis of a [0/90]s laminate. Transverse ply crack growth
and associated stiffness reduction during the fatigue of a simple cross-ply laminate. In S.
L. Ogin, P. A. Smith, and P. W. R. Beaumont (eds.), Report CUED/C/MATS/TR 105,
Cambridge University, Engineering Department, UK (September 1984).
47. S. L. Ogin, P. A. Smith, and P. W. R. Beaumont, Matrix cracking and stiffness
reduction during the fatigue of [0/90]s GFRP laminate. Compos Sci Technol, 22:1
(1985), 23–31.
48. S. L. Ogin, P. A. Smith, and P. W. R. Beaumont, Stress intensity factor approach to the
fatigue growth of transverse ply cracks. Compos Sci Technol, 24:1 (1985), 47–59.
49. R. J. Nuismer and S. C. Tan, Constitutive relations of a cracked composite lamina.
J Compos Mater, 22:4 (1988), 306–21.
50. J. Fan and J. Zhang, In-situ damage evolution and micro/macro transition for lamin-
ated composites. Compos Sci Technol, 47:2 (1993), 107–18.
51. M. Kashtalyan and C. Soutis, Stiffness degradation in cross-ply laminates damaged by
transverse cracking and splitting. Compos A, 31:4 (2000), 335–51.
52. J. Q. Zhang, K. P. Herrmann, and J. H. Fan, A theoretical model of matrix cracking in
composite laminates under thermomechanical loading. Acta Mech Solida Sin, 14:4
(2001), 299–305.
53. M. Kashtalyan and C. Soutis, Analysis of composite laminates with intra- and inter-
laminar damage. Prog Aerosp Sci, 41:2 (2005), 152–73.
54. J. M. Berthelot, Analysis of the transverse cracking of cross-ply laminates: a general-
ized approach. J Compos Mater, 31:18 (1997), 1780–805.
55. J. M. Berthelot, P. Leblond, A. El Mahi, and J. F. Le Corre, Transverse cracking of
cross-ply laminates: part 1. Analysis. Compos A, 27:10 (1996), 989–1001.
56. J. W. Lee, I. M. Daniel, and G. Yaniv, Fatigue life prediction of cross-ply composite
laminates. Compos Mater: Fatigue and Fracture, 2 (1989), 19–28.
57. J. A. Nairn and S. F. Hu, The formation and effect of outer-ply microcracks in cross-
ply laminates – a variational approach. Eng Fract Mech, 41:2 (1992), 203–21.
58. P. W. Manders, T. W. Chou, F. R. Jones, and J. W. Rock, Statistical analysis of
multiple fracture in [0/90/0] glass fiber/epoxy resin laminates. J Mater Sci, 18:10 (1983),
2876–89.
59. N. Laws and G. J. Dvorak, The loss of stiffness of cracked laminates. In Proceedings of the
IUTAM Eshelby Memorial Symposium. (Cambridge: Cambridge University Press, 1985).
60. N. Laws, G. J. Dvorak, and M. Hejazi, Stiffness changes in unidirectional composites
caused by crack systems. Mech Mater, 2:2 (1983), 123–37.
61. R. Hill, A self-consistent mechanics of composite materials. J Mech Phys Solids, 13:4
(1965), 213–22.
130 Micro-damage mechanics

62. B. Budiansky, On elastic moduli of some heterogeneous materials. J Mech Phys Solids,
13:4 (1965), 223–7.
63. J. D. Eshelby, The determination of the elastic field of an ellipsoidal inclusion, and
related problems. Proc R Soc London A, 241:1226 (1957), 376–96.
64. J. D. Eshelby, Elastic inclusion and inhomogeneities. In Progress in Solid Mechanics.
(Amsterdam: North Holland, 1961), pp. 89–140.
65. K. Hoiseth, A micromechanics study of transverse matrix cracking in cross-ply
composites. M.Sc. thesis, Georgia Institute of Technology, Atlanta, GA (1995).
66. J. Qu and K. Hoiseth, Evolution of transverse matrix cracking in cross-ply laminates.
Fatigue Fract Eng Mater Struc, 21:4 (1998), 451–64.
67. Z. Hashin, Analysis of cracked laminates: A variational approach. Mech Mater, 4:2
(1985), 121–36.
68. Z. Hashin, Analysis of composite materials – a survey. J Appl Mech, Trans ASME, 50:3
(1983), 481–505.
69. J. A. Nairn, The strain energy release rate of composite microcracking: A variational
approach. J Compos Mater, 23:11 (1989), 1106–29.
70. J. E. Bailey, P. T. Curtis, and A. Parvizi, On the transverse cracking and longitudinal
splitting behavior of glass and carbon-fiber reinforced epoxy cross-ply laminates and
the effect of Poisson and thermally generated strain. Proc R Soc London A, 366:1727
(1979), 599–623.
71. J. Varna and L. A. Berglund, Multiple transverse cracking and stiffness reduction in
cross-ply laminates. J Compos Tech Res, 13:2 (1991), 97–106.
72. J. Varna and L. Berglund, Two-dimensional transverse cracking in [0(m)/90(n)](s)
cross-ply laminates. Eur J Mech A – Solids, 12:5 (1993), 699–723.
73. J. Varna and L. A. Berglund, Thermoelastic properties of composite laminates with
transverse cracks. J Compos Tech Res, 16:1 (1994), 77–87.
74. Z. Hashin, Thermal-expansion coefficients of cracked laminates. Compos Sci Technol,
31:4 (1988), 247–60.
75. Z. Hashin, Analysis of orthogonally cracked laminates under tension. J Appl Mech,
Trans ASME, 54:4 (1987), 872–9.
76. S. Kuriakose and R. Talreja, Variational solutions to stresses in cracked cross-ply
laminates under bending. Int J Solids Struct, 41:9–10 (2004), 2331–47.
77. J. Varna and A. Krasnikovs, Transverse cracks in cross-ply laminates 2. Stiffness
degradation. Mech Compos Mater, 34:2 (1998), 153–70.
78. J. Varna, L. Berglund, A. Krasnikovs, and A. Chihalenko, Crack opening geometry in
cracked composite laminates. Int J Damage Mech, 6 (1997), 96–118.
79. R. Joffe and J. Varna, Analytical modeling of stiffness reduction in symmetric and balanced
laminates due to cracks in 90 degrees layers. Compos Sci Technol, 59:11 (1999), 1641–52.
80. L. N. McCartney, Theory of stress transfer in a 0-degrees-90-degrees-0-degrees
cross-ply laminate containing a parallel array of transverse cracks. J Mech Phys Solids,
40:1 (1992), 27–68.
81. L. N. McCartney, Energy-based prediction of failure in general symmetric laminates.
Eng Fract Mech, 72:6 (2005), 909–30.
82. L. N. McCartney, Physically based damage models for laminated composites. Proc Inst
Mech Engineers L – J Mater Design App, 217:L3 (2003), 163–99.
83. L. N. McCartney, Prediction of ply crack formation and failure in laminates. Compos
Sci Technol, 62:12–13 (2002), 1619–31.
References 131

84. L. N. McCartney, Model to predict effects of triaxial loading on ply cracking in


general symmetric laminates. Compos Sci Technol, 60:12–13 (2000), 2255–79.
85. L. N. McCartney, Predicting transverse crack formation in cross-ply laminates.
Compos Sci Technol, 58:7 (1998), 1069–81.
86. Errata to “Model to predict effects of triaxial loading on ply cracking in general
symmetric laminates.” [Compos Sci Technol 2000; 60(12–13):2255–2279], Compos
Sci Technol, 62:9 (2002), 1273–4.
87. J. A. Nairn, Fracture mechanics of composites with residual thermal stresses. J Appl
Mech, Trans ASME, 64 (1997), 804–10.
88. S. Ostlund and P. Gudmundson, Numerical analysis of matrix-crack-induced delami-
nations in [+/55-degrees] GFRP laminates. Compos Eng, 2:3 (1992), 161–75.
89. P. Gudmundson and W. L. Zang, An analytic model for thermoelastic properties of
composite laminates containing transverse matrix cracks. Int J Solids Struct, 30:23
(1993), 3211–31.
90. P. Gudmundson and W. Zang, Thermoelastic properties of microcracked composite
laminates. Mech Compos Mater Struct, 29:2 (1993), 107–14.
91. P. Gudmundson and S. Ostlund, Numerical verification of a procedure for calculation
of elastic-constants in microcracking composite laminates. J Compos Mater, 26:17
(1992), 2480–92.
92. P. Gudmundson and S. Ostlund, 1st order analysis of stiffness reduction due to matrix
cracking. J Compos Mater, 26:7 (1992), 1009–30.
93. P. Gudmundson and S. Ostlund, Prediction of thermoelastic properties of composite
laminates with matrix cracks. Compos Sci Technol, 44:2 (1992), 95–105.
94. E. Adolfsson and P. Gudmundson, Matrix crack initiation and progression in
composite laminates subjected to bending and extension. Int J Solids Struct, 36:21
(1999), 3131–69.
95. E. Adolfsson and P. Gudmundson, Thermoelastic properties in combined bending
and extension of thin composite laminates with transverse matrix cracks. Int J Solids
Struct, 34:16 (1997), 2035–60.
96. E. Adolfsson and P. Gudmundson, Matrix crack induced stiffness reductions in
[(0_m/90_n/+theta_p/theta_q)(s)](m) composite laminates. Compos Eng, 5:1
(1995), 107–23.
97. J. Varna, N. V. Akshantala, and R. Talreja, Crack opening displacement and the associ-
ated response of laminates with varying constraints. Int J Damage Mech, 8 (1999), 174–93.
98. P. Lundmark and J. Varna, Constitutive relationships for laminates with ply cracks in
in-plane loading. Int J Damage Mech, 14:3 (2005), 235–59.
99. P. Lundmark, Damage mechanics analysis of inelastic behaviour of fiber composites.
Ph.D. thesis, Luleå University of Technology, Luleå, Sweden (2005), p. 175.
100. A. Krasnikovs and J. Varna, Transverse cracks in cross-ply laminates. I. Stress
analysis. Mech Compos Mater Struct, 33:6 (1997), 565–82.
101. R. Joffe, A. Krasnikovs, and J. Varna, COD-based simulation of transverse cracking and
stiffness reduction in [S/90n]s laminates. Compos Sci Technol, 61:5 (2001), 637–56.
102. R. Hill, Elastic properties of reinforced solids: Some theoretical principles. J Mech Phy
Solids, 11:5 (1963), 357–72.
103. O. Rubenis, E. Sp arninš, J. Andersons, and R. Joffe, The effect of crack spacing
distribution on stiffness reduction of cross-ply laminates. Appl Compos Mater, 14:1
(2007), 59–66.
132 Micro-damage mechanics

104. J. P. Benthem and W. T. Koiter, Asymptotic approximations to crack problems. In


Mechanics of Fracture I: Methods of Analysis and Solutions of Crack Problems, ed. G. C.
Sih. (Leyden: Noordhoff, 1972), pp. 131–78.
105. H. Tada, P. Paris, and G. Irwin, The Stress Analysis of Cracks Handbook. (St. Louis,
MO: Del Research Corporation, 1973).
106. P. Lundmark and J. Varna, Crack face sliding effect on stiffness of laminates with ply
cracks. Compos Sci Technol, 66:10 (2006), 1444–54.
107. J. Varna, Physical interpretation of parameters in synergistic continuum damage
mechanics model for laminates. Compos Sci Technol, 68:13 (2008), 2592–600.
108. S. Li, S. R. Reid, and P. D. Soden, A finite strip analysis of cracked laminates. Mech
Mater, 18:4 (1994), 289–311.
109. J. N. Reddy, A generalization of two-dimensional theories of laminated composite
plates. Commun Appl Nume Mthds, 3:3 (1987), 173–80.
110. S. Li, C. V. Singh, and R. Talreja, A representative volume element based on transla-
tional symmetries for FE analysis of cracked laminates with two arrays of cracks. Int
J Solids Struct, 46:7–8 (2009), 1793–804.
111. J. Noh and J. Whitcomb, Effect of various parameters on the effective properties of a
cracked ply. J Compos Mater, 35 (2001), 689–712.
112. S. Li, General unit cells for micromechanical analyses of unidirectional composites.
Compos A, 32:6 (2001), 815–26.
113. S. Li, On the unit cell for micromechanical analysis of fibre-reinforced composites.
Proc R Soc London A, 455:1983 (1999), 815–38.
114. S. Li, Boundary conditions for unit cells from periodic microstructures and their
implications. Compos Sci Technol, 68:9 (2008), 1962–74.
115. S. G. Li and A. Wongsto, Unit cells for micromechanical analyses of particle-
reinforced composites. Mech Mater, 36:7 (2004), 543–72.
116. K. Srirengan and J. D. Whitcomb, Finite element based degradation model for
composites with transverse matrix cracks. J Thermoplast Compos Mater, 11:2 (1998),
113–23.
117. J. Tong, F. J. Guild, S. L. Ogin, and P. A. Smith, On matrix crack growth in quasi-isotropic
laminates – II. Finite element analysis. Compos Sci Technol, 57:11 (1997), 1537–45.
118. F. G. Yuan and M. C. Selek, Transverse cracking and stiffness reduction in composite
laminates. J Reinf Plas Compos, 12:9 (1993), 987–1015.
119. R. Cook, D. Malkus, M. Plesha, and R. Witt, Concepts and Applications of Finite
Element Analysis. (John Wiley & Sons, 2002).
120. J. Reddy, Energy Principles and Variational Methods in Applied Mechanics, 4th edn.
(Hoboken, New Jersey: John Wiley & Sons, Inc., 2002).
121. J. N. Reddy, Mechanics of Laminated Composite Plates and Shells: Theory and Analysis,
2nd edn. (Boca Raton: CRC Press, 2004).
122. D. H. Robbins and J. N. Reddy, Analysis of piezoelectrically actuated beams using a
layer-wise displacement theory. Computers & Structures, 41:2 (1991), 265–79.
123. D. H. Robbins, J. N. Reddy, and F. Rostam-Abadi, An efficient continuum damage
model and its application to shear deformable laminated plates. Mech Adv Mater Struc,
12:6 (2005), 391–412.
124. J. E. S. Garcao, C. M. M. Soares, C. A. M. Soares, and J. N. Reddy, Analysis of
laminated adaptive plate structures using layerwise finite element models. Computers
& Structures, 82:23–26 (2004), 1939–59.
References 133

125. J. N. Reddy, An evaluation of equivalent-single-layer and layer-wise theories of


composite laminates. Compos Struct, 25:1–4 (1993), 21–35.
126. Y. S. N. Reddy, C. M. D. Moorthy, and J. N. Reddy, Nonlinear progressive failure
analysis of laminated composite plates. Int J Non Linear Mech, 30:5 (1995), 629–49.
127. D. H. Robbins and J. N. Reddy, Modeling of thick composites using a layerwise
laminate theory. Int J Numer Methods Eng, 36:4 (1993), 655–77.
128. W. J. Na and J. N. Reddy, Multiscale analysis of transverse cracking in cross-ply
laminates beams using the layerwise theory. J Solid Mech, 2:1 (2010), 1–18.
129. W. J. Na and J. N. Reddy, Delamination in cross-ply laminated beams using the
layerwise theory. Asian J Civil Eng, 10:4 (2009), 451–80.
130. J. Aboudi, S. W. Lee, and C. T. Herakovich, Three-dimensional analysis of laminates
with cross cracks. J Appl Mech, Trans ASME, 55:2 (1988), 389–97.
131. S. W. Lee and J. Aboudi, Analysis of Composites Laminates with Matrix Cracks, 3rd edn.
Report CCMS-88–03, College of Engineering, Virginia Polytechnic Institute and State
University, Blacksburg, Virginia (1988).
132. J. H. Lee and C. S. Hong, Refined two-dimensional analysis of cross-ply laminates
with transverse cracks based on the assumed crack opening deformation. Compos Sci
Technol, 46:2 (1993), 157–66.
133. D. Gamby and J. L. Rebiere, A two-dimensional analysis of multiple matrix cracking
in a laminated composite close to its characteristic damage state. Compos Struc, 25:1–4
(1993), 325–37.
134. D. Zhang, J. Ye, and D. Lam, Properties degradation induced by transverse cracks in
general symmetric laminates. Int J Solids Struct, 44:17 (2007), 5499–517.
135. D. Zhang, J. Ye, and D. Lam, Ply cracking and stiffness degradation in cross-ply
laminates under biaxial extension, bending and thermal loading. Compos Struct,
75:1–4 (2006), 121–31.
136. N. J. Pagano, On the micromechanical failure modes in a class of ideal brittle matrix
composites. Part 1. Coated-fiber composites. Compos B, 29:2 (1998), 93–119.
137. N. J. Pagano, G. A. Schoeppner, R. Kim, and F. L. Abrams, Steady-state cracking
and edge effects in thermo-mechanical transverse cracking of cross-ply laminates.
Compos Sci Technol, 58:11 (1998), 1811–25.
138. G. A. Schoeppner and N. J. Pagano, Stress fields and energy release rates in cross-ply
laminates. Int J Solids Struc, 35:11 (1998), 1025–55.
139. L. N. McCartney, G. A. Schoeppner, and W. Becker, Comparison of models for transverse
ply cracks in composite laminates. Compos Sci Technol, 60:12–13 (2000), 2347–59.
140. J. A. Nairn, The strain-energy release rate of composite microcracking – a variational
approach. J Compos Mater, 23:11 (1989), 1106–29.
141. R. Y. Kim, A. S. Crasto, and G. A. Schoeppner, Dimensional stability of composite in
a space thermal environment. Compos Sci Technol, 60:12–13 (2000), 2601–8.
5 Macro-damage mechanics

5.1 Introduction

As described in the previous chapter, the two main approaches to treating damage
in composite materials are micro-damage mechanics (MIDM) and macro-damage
mechanics (MADM). We shall now discuss MADM, which is also commonly
known as continuum damage mechanics (CDM).
Consider an undamaged continuum (Figure 5.1, left) representing a hetero-
geneous solid undergoing deformation upon prescribed traction t and displacement
u on parts of the bounding surfaces St and Su, respectively. Assuming the average
response of the solid to be linear elastic, the stress–strain relationships can be
written as

sij ¼ Cijkl ekl ; ð5:1Þ

where sij ; ekl ; and Cijkl denote the tensors of stress, strain, and stiffness, respectively,
at a point of the continuum body in its undamaged state. At some loading state, the
body may initiate and evolve damage in the form of voids, microcracks, cavities, etc.
(Figure 5.1, middle). These damage entities will cause perturbations in the local
stress and strain states of the continuum body. The stress–strain response averaged
over a representative volume element (RVE) at a fixed damage state is still given by
Eq. (5.1) but now Cijkl denotes the stiffness tensor of the homogenized continuum
body containing damage. The main objective of the CDM approach is to character-
ize Cijkl for the continuum body in which damage entities have been homogenized.
A concept central to CDM is the “internal” state representing the damage entities
that have been homogenized in the continuum (Figure 5.1, right). Historically, this
concept is credited to Kachanov [1, 2] who introduced the so-called “continuity” as a
varying state of a metal in an attempt to describe its degradation due to creep.
Accordingly, a single scalar variable, denoted as f, was used to describe the continu-
ity with values equal to 1 in the virgin state, and 0 at rupture. The loss of “continuity”
signifies material deterioration, and was assumed to occur at a rate given by
 m
df s
¼ A ; ð5:2Þ
dt f

where A and m are material constants and s is the maximum principal tensile stress.
5.1 Introduction 135

ti ti ti
St St
St

Damage Homogenization

Initiation

ui ui ui
Su Su Su
Undamaged Continuum Homogenized
homogenized with damage damaged
continuum continuum

Figure 5.1. The basic concept of internal variable damage mechanics.

Force, F = s.S Force, F = s∗.S∗

Area S*

Area S

Damaged state Equivalent fictitious undamaged state


(0<ϕ<1) (ϕ = 0)

Figure 5.2. The concept of effective stress for isotropic damage in uniaxial loading.

The variable f can be viewed as representing surface density of discontinuities


in the material, and can then be used to describe an “effective” stress. Thus in
Eq. (5.2), the effective stress is equal to s/f, which represents the current level
of stress causing further material degradation. Later, Robotnov [3] interpreted
“discontinuity,” defined by the parameter o ¼ 1  f, as a measure of net area
reduction (see Figure 5.2). Accordingly, o is defined as
SD S  S
o¼ ¼ ; ð5:3Þ
S S
where the “damaged” cross section SD ¼ S – S*, in which S is the cross-sectional
area in the undamaged state and S* is the net area of the damaged specimen which
excludes the area held by the damage entities (discontinuities). Thus the net area of
the specimen effectively carries the applied load. The damage parameter o ¼ 0
signifies an undamaged state, while o ¼ oc represents a critical state corresponding
136 Macro-damage mechanics

to the rupture of the specimen. Following this approach, Robotnov [3] postulated
the creep strain rate as

de s
n
¼B ; ð5:4Þ
dt 1o
where B and n are constants. It should be noted here that Kachanov’s
effective stress, as defined, increases to infinity at failure, while Robotnov’s effect-
ive stress, as interpreted, would increase only by a few percent since the volume
fraction of voids or discontinuities at failure is found by microstructural studies
for metals to be small.
Robotvov’s concept of effective stress was also used later by Murakami and
Ohno [4] for creep damage of polycrystalline metals and by Lemaitre and co-
workers [5, 6] for effective elastic properties. Denoting the quantity in parenthesis
in Eq. (5.4) by
s

s ; ð5:5Þ
1o
Hooke’s law can be expressed in two equivalent forms as

~ e; s
s ¼ Ee ~ ¼ Eee ; ð5:6Þ
where ee is the elastic strain and E~ is the effective Young’s modulus, which is
related to the damage parameter o as follows

s E E~

s ¼ s)o¼1 : ð5:7Þ
1  o E~ E
In a somewhat similar way, for brittle creep damage, Lemaitre and Chaboche [7]
defined the damage variable in terms of the strain rate. Using a power law, the
strain rate during secondary creep (undamaged state), e_ s , can be described by
s
N
e_ s ¼ ; ð5:8Þ
l

where l is a material constant and the exponent N is found from experimental


tests. The damage variable follows easily from the measurement of the strain rate
during tertiary creep, e_ , and the use of effective stress concept as
 1=N
e_
o¼1 : ð5:9Þ
e_ s
Similar to the effective stress concept, the “strain equivalence principle” has also
been proposed by Lemaitre and coworkers [6, 8], which states that “any strain
constitutive equation for a damaged material may be derived in the same way as
for a virgin material except that the usual stress is replaced by the effective stress.”
In creep of metals as well as in brittle cracking of ceramics, rocks, concrete, etc. the
voids and cracks usually form along some preferred orientations, e.g., on grain
5.1 Introduction 137

boundaries and in weak planes. It is obvious that a scalar damage variable cannot
account for the directional dependence of the effects of voids and cracks.
To characterize this effect, Murakami and Ohno [4] considered an arbitrarily oriented
plane in a solid containing arbitrarily distributed voids and defined a second-order
tensor to describe the anisotropic net area reduction. This tensor was assumed to
represent the directional nature of discontinuities and was called the damage tensor.
The Murakami–Ohno damage tensor V relates the elemental area dA of an
arbitrarily oriented plane with the unit normal n lying in the damaged configur-
ation to the elemental area dA* of the same plane with the unit normal n* lying in
a fictitious configuration with reduced net load carrying area. Thus,

n dA ¼ ðI  VÞn dA; ð5:10Þ

where I is the identity tensor of second order.


The effective stress tensor s* then is the magnified (net area) stress in the fictitious
configuration which is related to the stress in the damaged configuration by

s ¼ðI  VÞ1 s: ð5:11Þ

Following (5.7), Chaboche [9–11] generalized the relation between elastic con-
stants and the damage variable to the three-dimensional case, replacing o by a
damage tensor D. This damage tensor, in general, is a nonsymmetric fourth-order
tensor, and is defined by the following transformation

E ¼ ðI  DÞ : E; ð5:12Þ

where I is the identity tensor of fourth order, E and E are the elasticity tensor for
the undamaged and damaged materials, respectively, and (:) stands for the second-
order contraction. Thus, the damage tensor is given by

D ¼ I  E
E1 ; ð5:13Þ

and the effective stress tensor is given by

s ¼ ðI  DÞ1 : s: ð5:14Þ

It is worth noting that the second-order damage tensor defined by Murakami and
Ohno [4], Eq. (5.10), is based on the notion of effective net area reduction, while
the fourth-order damage tensor defined by Chaboche [9–11], Eq. (5.13), uses the
notion of the effective stress defined by Eq. (5.6).
The damage tensor D defined by (5.13) has been the focus of much attention in
several works on continuum damage mechanics. Ju [12], for instance, has dis-
cussed the notion of isotropic and anisotropic damage variables and has demon-
strated that isotropic damage does not necessarily imply a scalar damage variable.
From a mechanics study of changes in the elasticity compliance tensor due to
microcracks he found that the fourth-order tensor D is isotropic for the case of
isotropic damage, e.g., when the cracks are perfectly randomly distributed in all
138 Macro-damage mechanics

Figure 5.3. Schematic illustration of anisotropic constraint effect in composites. An


ellipsoid entity is shown surrounded by fibers or plies in a composite. The axes of the
ellipsoid are inclined with respect to the axes of the constraining elements. Reprinted, with
kind permission, from Damage Mechanics of Composite Materials, R. Talreja, Damage
characterization by internal variables, pp. 53–78, copyright Elsevier (1994).

directions. However, for preferred directions in crack orientations the damage


tensor will, in general, be anisotropic.

5.2 Continuum damage mechanics (CDM) of composite materials

The damage variables defined on the basis of the effective stress concept, or
the associated notion of the reduced net area, do not contain any specific details
of the damage entities. In the Lemaitre–Chaboche damage tensor, for instance,
the components of the damage tensor are determined from the elastic compli-
ance changes. Thus, e.g., two sets of damage entities of different characteristic
sizes and concentrations leading to the same elastic compliance changes will be
represented by the same variables. Such equivalency of different sets of damage
entities cannot, in general, be expected to hold for all effects of damage.
Furthermore, in media with, say, two levels of microstructure, e.g., in compos-
ites, the damage entities forming at the lower size-scale may be affected by the
higher size-scale microstructure as well as by the difference in the symmetry
properties of the two microstructures. To illustrate this aspect, consider an
ellipsoidal particle embedded in a compliant matrix and surrounded by a
parallel array of stiff fibers (Figure 5.3). Let the particle debond from the
matrix under appropriate loading such that an ellipsoidal damage entity is
formed. The local stress perturbation induced by the surface displacements of
this damage entity, and the resulting overall elastic compliance, will depend on
the local microstructure details such as inclination of the fibers with respect to
the principal planes of the damage entity and the fiber stiffness. Thus the same
microstructure arranged differently can produce a different effect on the overall
elastic response. This example serves to illustrate that a characterization of
damage in terms of the elastic compliance changes will not necessarily be
unique.
5.2 Continuum damage mechanics (CDM) of composite materials 139

In the alternative characterization of damage, adopted here, the damage


entities are represented by appropriate specific variables, and averages of these
variables over an RVE then give the damage variables. The degree of charac-
terization, i.e., to what extent the details of damage entities are represented,
depends on the type of variables chosen. Thus, a scalar variable will only
represent the size but not the shape or orientation characteristics of the damage
entities. A vector variable can account for the size and orientation, but the
positive and negative senses of a vector introduce ambiguities in the damage
characterization that cause difficulties in the description of damage evolution.
With two vectors simultaneously associated with a damage entity these difficul-
ties are overcome. At the same time, the resulting second-order tensor gives
mathematical convenience of formulating response functions, which are com-
monly expressed in terms of the second-order stress and strain tensors, as
discussed below.
Talreja’s original paper on CDM used a vectorial description of damage [13].
Although the ambiguities concerning the sense of damage vector could be remed-
ied by imposing an additional condition, it was found that using a second-order
tensor instead avoided this step [14–17]. Here we will follow the tensorial
description.

5.2.1 RVE for damage characterization


Any continuum description of a solid entails homogenization since materials
are inherently heterogeneous. For polycrystalline metals, for instance, the scale
of heterogeneity (e.g., grain size) is often small compared to the scale at which
material response characteristics (e.g., the elastic constants) are measured,
allowing the stress and strain states to be defined as continuous fields. For
commonly used fiber-reinforced solids, such as glass/epoxy and carbon/epoxy,
the fiber diameter of approximately 10 micrometers allows treating these
materials as a homogeneous continuum with good accuracy. When internal
surfaces in composite materials form, their characteristic dimensions and
mutual distances between them can be orders of magnitude larger than the
scale of heterogeneities underlying the homogenized pristine composite.
Furthermore, on application of external loads the internal surfaces are subject
to evolution (enlargement and multiplication), as discussed above. This war-
rants a separate (and different) homogenization of the composite with internal
surfaces (collectively called damage).
Figure 5.4 depicts a homogenization procedure for a composite solid contain-
ing damage. The heterogeneities in the pristine (undamaged) composite are
referred to as “stationary microstructure” and are homogenized first. This may
also be called “classical” homogenization. Textbooks on mechanics of composite
materials usually begin with this homogenization. In fact, the classical laminate
theory goes one step further by developing homogeneous constitutive relations
for laminates consisting of stacked layers of homogenized unidirectionally
140 Macro-damage mechanics

Stationary microstructure
RVE for damage characterization

Evolving microstructure

Step 1 Homogenization of P
stationary microstructure

V
Step 2

Homogenization of P
evolving microstructure
ai
nj

Continuum after homogenizing Fully homogenized continuum


the stationary microstructures
Characterization of a
damage entity

Figure 5.4. Homogenization of a continuum body with heterogeneous stationary


structure and evolving damage entities. A tensorial characterization of a damage entity
is depicted on the right.

reinforced composite (ply or lamina). Returning to Figure 5.4, the second hom-
ogenization pertains to the internal surfaces, collectively named as damage or
“evolving microstructure” to highlight their ability to permanently change by
processes of energy dissipation. Homogenization of the evolving microstructure
necessitates employing the notion of a representative volume element (RVE),
which will be discussed next.
A general and thorough exposition of the RVE notion in the context of micro-
mechanics is given in [18]. Here, we shall apply this notion to the particular case
of composite materials with damage. With reference to Figure 5.4 again, a generic
point P in the homogenized composite with damage has associated with it a
damage state (in addition to stress and strain states), which is given by an
appropriate volume-averaged measure of the presence of internal surfaces that
affect the constitutive behavior (stress–strain relations) at the point P. The volume
over which the averaging is performed must be representative of the neighborhood
of point P that can be associated with P. This neighborhood is the RVE, whose
volume is not fixed but depends on the geometrical configuration (size, spacing,
etc.) of the internal surfaces around P. As this configuration changes under
applied loading, the RVE size changes.
With the notion of RVE at hand, the damage state at P can be defined by a
set of variables obtained by averaging over the RVE. The choice of the
variables is guided by the type of internal surfaces formed, and in this respect
5.2 Continuum damage mechanics (CDM) of composite materials 141

the knowledge of damage mechanisms discussed above is useful. In general,


the variables can be scalars, vectors, or tensors of second order or higher.
Settling on which variables to employ is a matter of finding a balance between
capturing sufficient physics of the damage process and usefulness of the ensu-
ing formulation of constitutive relations. In the following the second-order
characterization of damage in composite materials adopted by Talreja [14, 15,
17] is described.

5.2.2 Characterization of damage


As shown in Figure 5.4, a single internal surface within a RVE, called a damage
entity from now on, can be characterized by two vectors: a unit outward normal n
at a point on the surface, and an “influence” vector a at the same point. A dyadic
product of the two vectors, integrated over the surface S, is denoted the damage
entity tensor, and given by
ð
dij ¼ ai nj dS; ð5:15Þ
S

where the components of the vectors are with reference to a Cartesian coordinate
system. The dyadic product assures consistency of the signs of the two vectors.
A characterization of this type was first proposed by Vakulenko and Kachanov
[19] for flat cracks, where the influence vector a represented the displacement jump
across the crack surface.
The physical significance of this characterization is that it represents the
oriented nature of the presence of internal surfaces. As illustrated by the examples
of damage in Chapter 3, common internal surfaces are cracks (flat or curved)
generated by interface debonding and matrix failure. The unit normal vector at a
point on the damage entity carries the information on orientation of the surface
(with respect to the frame of reference), while the other vector represents an
appropriate influence induced by activation of the considered point on the surface.
This influence is generally also directed in nature. For the case of mechanical
response, the appropriate influence would be the displacement of the activated
point on the damage entity surface. For a nonmechanical response, such as
thermal or electrical conductivity, the perturbation induced by an internal surface
can also be cast as a vector-valued quantity.
Integrating the dyadic product in Eq. (5.15) over the damage entity surface
provides the total net effect of the entity. For example, if the entity is a flat crack,
then taking a as the displacement vector in the integral gives the crack surface
separation times the crack surface area. This product may be viewed as an
affected volume associated with the crack. For a penny-shaped crack with the
two surfaces separating symmetrically about the initial crack plane, the sole
surviving term of the damage entity tensor represents an ellipsoidal-shaped
volume.
142 Macro-damage mechanics

Referring once again to Figure 5.4, the RVE associated with a generic point P
carries a sufficiently large number of the discrete damage entities to represent
the collective effect on the homogenized constitutive response at the point. The
number of damage entities needed for this representation, and the consequent
RVE size, depend on the distribution of the entities. For instance, if the entities
are sparsely distributed, then the RVE size would be large, while for densely
distributed case a small RVE would suffice. Furthermore, for uniformly distrib-
uted entities of the same geometry, a repeating unit cell containing a single
entity can replace the RVE, while for the cases of nonuniform distribution of
unequal entities, the RVE size will increase until a statistically homogeneous
representation is attained. This implies that further increasing the RVE size will
have no impact on the averages of the selected characteristics. As an example, if
the selected characteristic is the affected matrix volume by a damage entity, as
mentioned above, then the average value of this quantity will vary as the RVE
size increases and will approach a constant value at a certain RVE size. The
minimum RVE size beyond which no appreciable change in the considered
average is found may be taken as the needed RVE. It is apparent that the
RVE is not unique but is subject to the choice made for the particular formula-
tion of the constitutive response of a continuum with damage. Consequently,
there is no unique constitutive theory of a continuum with damage; however,
the use of the concept of an internal state in a given theory requires specifying
RVE in a consistent manner and assuring that the conditions for its existence
are present.
From the cases of damage mechanisms reviewed in Chapter 3 it can be noted
that in composite laminates the damage tends to occur as sets of parallel cracks
within the plies, each oriented along fibers in the given ply. It is therefore
convenient to separate each set of ply cracks according to its orientation,
referred to a fixed frame of reference, and assign it a damage mode number.
Denoting damage mode by a ¼ 1, 2, . . ., n, a damage mode tensor can be
defined as
ðaÞ 1 X 
Dij ¼ dij ka ; ð5:16Þ
V k
a

where ka is the number of damage entities in the ath mode, and V is the RVE
volume. As noted above, if the ply cracks of a given orientation are uniformly
spaced, then the RVE will reduce to the unit cell containing one crack. For
nonuniform distribution of ply cracks, V must be large enough to provide a steady
average of the damage mode tensor components.
As defined by Eq. (5.16) the damage mode tensor will in general be asymmet-
rical. Decomposing the influence vector a along directions normal and tangential
to the damage entity surface S gives,
ai ¼ ani þ bmi ; ð5:17Þ
where n and m are unit normal and tangential vectors on S such that nimi ¼ 0.
5.2 Continuum damage mechanics (CDM) of composite materials 143

Using Eq. (5.17) in Eq. (5.16) the damage entity tensor can be written in two
parts as

dij ¼ dij1 þ dij2 ; ð5:18Þ

where
ð ð
dij1 ¼ ani nj dS; dij2 ¼ bmi nj dS: ð5:19Þ
S S

Here a and b are the magnitudes of the normal and tangential projections,
respectively, of vector ai and vectors ni and mj are unit normal and tangential
vectors, respectively. The damage mode tensor for a given mode can now be
written as

ðaÞ 1ðaÞ 2ðaÞ


Dij ¼ Dij þ Dij ; ð5:20Þ

where

1ðaÞ 1 X 1
2ðaÞ 1 X 2

Dij ¼ dij ; Dij ¼ dij : ð5:21Þ


V k ka V k ka
a a

This separation of the damage mode tensor in two parts allows the analysis to
be simplified to avoid having to deal with asymmetric tensors. For instance,
for damage entities consisting of flat cracks, the two parts of the damage mode
tensor represent the two crack surface separation modes. If an assumption can
be made that only the symmetric crack surface separation (known as mode I or
crack opening mode in fracture mechanics) is significant, then the second term
in Eq. (5.21) can be neglected. This will render the damage mode tensor
symmetrical and it can then be written as
ð 
ðaÞ 1ðaÞ 1X
Dij ¼ Dij ¼ ani nj dS : ð5:22Þ
V k S ka
a

The consequence of this assumption was examined by Varna [20] for one class of
laminates and it was found that not including the crack sliding displacement
(CSD) for ply cracks inclined to the laminate symmetry directions resulted in
errors in estimating the degradation of the average elastic properties of laminates.
However, these errors were found to be small in absolute values while being
significant in percentages. In fact for those ply crack orientations where CSD
dominates, the cracks are difficult to initiate until high loads close to failure load
are applied.
Some damage mechanisms such as crystalline slip may require only the
tangential part. For other situations the sliding between the crack faces can
be negligible, e.g., for intralaminar cracks constrained by stiff plies, and fiber/
matrix debonds. Hence Dij2(a) ¼ 0 under such conditions. For cases where the
144 Macro-damage mechanics

Table 5.1 Thermodynamics variables for damage analysis

Thermodynamic state variables Thermodynamic response variables


1. Strain tensor eij ¼ 12 ui; j þ uj;i
 
1. Cauchy stress tensor sij
2. Temperature T 2. Specific Helmholtz free energy c
3. Temperature gradient gi ¼ T,i 3. Specific entropy 
4. Damage tensors Dij(a) 4. Heat flux vector qi
ðaÞ
5. Damage rate tensors D_ ij

damage entity surfaces conduct tangential displacements only (e.g., CSD by flat
cracks), it is possible to formulate the damage mode tensor as a symmetric tensor.
One example of this is sliding of the fiber/matrix interface in ceramic matrix
composites [16].
With stress, strain, and damage, all expressed as symmetric second-order
tensors, a constitutive theory can now be formulated to have a convenient, usable
form. Such a formulation is described next.

5.2.3 A thermodynamics framework for materials response


Referring once again to Figure 5.4, a formulation of the constitutive response
of a homogenized continuum with damage will now be discussed. In view of the
observed behavior of common composite materials such as glass/epoxy and
carbon/epoxy, only elastic response will be considered. Theoretical treatment of
elastic response of solids is classical and can be found in textbooks. Incorpor-
ating damage is, however, not a simple extension of the classical theory of
elasticity. The CDM approach to be described here is based on thermodynam-
ics and is naturally suited for thermomechanical response. It can be extended to
incorporate nonmechanical effects, such as electrical and magnetic, as well as
chemical. Every extension, however, comes with the price of having to deter-
mine associated response coefficients (material constants) by a certain identifi-
cation procedure. In the treatment presented here, the task of determining
material constants is reduced by use of selected micromechanics. This way of
combining micromechanics with CDM generates useful synergism, justifying
the characterization of the combined approach as synergistic damage mechan-
ics (SDM), to be discussed later. We begin with the conventional CDM frame-
work first.
At the foundation of CDM are the first and second laws of thermodynamics.
Additionally, use is made of the concept of an internal state, which is identified here
as the evolving microstructure depicted in Figure 5.4. As discussed before, this
microstructure is homogenized into a damage field characterized by the set of damage
mode tensors Dij(a). The collection of all variables resulting from thermodynamics with
internal state can now be placed in two categories: state variables and response
functions (see Table 5.1). The thermodynamical response of a composite body, limited
to small strains, is given by a set of five response functions: the Cauchy stress tensor sij,
5.2 Continuum damage mechanics (CDM) of composite materials 145

the specific Helmholtz free energy c, the specific entropy , the heat flux vector qi, and
ðaÞ
a set of damage rate tensors D_ ij , a ¼ 1,2,
 ..., n. The
 thermodynamic state of the body is
given by the strain tensor eij ¼ ð1=2Þ ui; j þ uj;i , with displacement vector ui, absolute
temperature T, temperature gradient gi ¼ T,i, and a set of damage tensors D(a) ij .
Following Truesdell’s principle of equipresence, which states that all state
variables should be present in all response functions unless thermodynamics or
other relevant considerations preclude their dependency, we write

ðaÞ
sij ¼ sij ekl ; T; gk ; Dkl ;

ðaÞ
c ¼ c ekl ; T; gk ; Dkl ;

ðaÞ
 ¼  ekl ; T; gk ; Dkl ; ð5:23Þ

ðaÞ
q ¼ q ekl ; T; gk ; Dkl ;

ðaÞ ðaÞ ðbÞ


D_ kl ¼ D_ kl ekl ; T; gk ; Dkl :

The following balance laws should hold for the continuum body:

Balance of linear momentum:


sij;j þ rbj ¼ r€
xj ; ð5:24Þ

where bj are the components of the body force per unit mass and r is the mass
density.
Balance of angular momentum:

sij ¼ sji : ð5:25Þ

Balance of energy:
ru_  sij e_ ij þ qi;i ¼ rr; ð5:26Þ
where u is the specific internal energy per unit mass and r is the heat supply per
unit mass.
Second law of thermodynamics in the form of the Clausius–Duhem
inequality:
qi gi
sij e_ ij  rc_  rT
_   0; ð5:27Þ
T
where c ¼ u  T.
Time differentiation of c in (5.23) gives,

@c @c _ @c @c _ ðaÞ
c_ ¼ e_ kl þ Tþ g_ i þ D :
ðaÞ kl
ð5:28Þ
@ekl @T @gi @D kl

Substitution of Eq. (5.28) into Eq. (5.27) gives


146 Macro-damage mechanics

   
@c @c _ @c X @c ðaÞ qi gi
sij  r e_ ij  r  þ Tr g_ i  r D_ 
ðaÞ kl
 0; ð5:29Þ
@ekl @T @gi a @D
T
kl

where the summation sign is shown on a to include all damage modes. Now
requiring (5.29) to hold for the independently varying strain, temperature, and
temperature gradient gives the following results

@c
sij ¼ r ; ð5:30Þ
@ekl

@c
¼ ; ð5:31Þ
@T
@c
¼ 0: ð5:32Þ
@gi

Equation (5.32) states that the Helmholz free energy function does not depend on
the temperature gradient and, consequently, Eqs. (5.30) and (5.31) eliminate this
dependency from stress and entropy. The last two equations in the response
function set (Eq. (5.23)) remain unaffected.
Equations (5.30)–(5.32) lead to the following functional dependencies

ðaÞ
sij ¼ sij ekl ; T; Dkl ;

ðaÞ
c ¼ c ekl ; T; Dkl ;

ðaÞ
 ¼  ekl ; T; Dkl ; ð5:33Þ

ðaÞ
q ¼ q ekl ; T; gk ; Dkl ;

ðaÞ ðaÞ ðbÞ


D_ kl ¼ D_ kl ekl ; T; gk ; Dkl :

From Eq. (5.29) the following restriction (also known as internal dissipation
inequality) results
X ðaÞ ðaÞ qi gi
Rkl D_ kl   0; ð5:34Þ
a
T

where R(a) (a)


kl are the thermodynamic forces conjugate to Dkl and are given by

ðaÞ @c
Rkl ¼ r ðaÞ
: ð5:35Þ
@Dkl
Each of these forces is analogous to the crack extension force (i.e., energy release
rate) for a single crack. As an example, for a damage mode component, say
D11 of mode a ¼ 1, the quantity R(1) 11 can be interpreted as the “force” causing
an infinitesimal change in the internal state represented by D(1)
11 . Equation (5.34)
5.2 Continuum damage mechanics (CDM) of composite materials 147

expresses the condition these forces must satisfy as damage evolves under thermo-
mechanical impulses.
The complete thermomechanical response for the composite body is governed
by the set of functions Eq. (5.33) and the internal dissipation inequality Eq. (5.34).
In view of the experimental data, which are mostly available for polymer matrix
composites at room temperature, the thermomechanical framework will be
developed further for the purely mechanical response. Thus, for isothermal condi-
tions (T ¼ 0, gi ¼ 0) the set of response functions is reduced to the following

ðaÞ
sij ¼ sij ekl ; Dkl ;

ðaÞ
c ¼ c ekl ; Dkl ;

ð5:36Þ
ðaÞ ðaÞ ðbÞ
D_ kl ¼ D_ kl ekl ; gk ; Dkl ;

ðaÞ ðaÞ
X
Rkl D_ kl  0 :
a

Since sij are derivable from c, according to Eq. (5.30), it suffices to formulate c and
ðaÞ
D_ kl for a purely mechanical response. Dealing with this scalar-valued function (c) as
the sole response function for a given internal state of damage provides a favorable
situation for further development of the theory. The form of the Helmholz free
energy function can be chosen in different ways. A powerful way is possible by use
of the theory of invariants for polynomial functions [21]. In the following it is
illustrated for one case of orthotropic composites containing one damage mode.
To derive the rate equations for sij, we differentiate Eq. (5.30) with respect to
time and use the functional dependency given in Eq. (5.36), which yields

@2c X @2c
s_ ij ¼ r e_ kl þ r D_ ðaÞ :
ðaÞ mn
ð5:37Þ
@eij @ekl a @e ij @Dmn

Substitution of Eq. (5.35) gives


ðaÞ
@2c X @Rmn
s_ ij ¼ r e_ kl  D_ ðaÞ ; ð5:38Þ
@eij @ekl a
@eij mn

which may be rewritten as


ðaÞ
X
s_ ij ¼ Cijkl e_ kl  Kijmn D_ ðaÞ
mn ; ð5:39Þ
a

where

@2c
Cijkl ¼ r ð5:40Þ
@eij @ekl
and
148 Macro-damage mechanics

ðaÞ
ðaÞ @Rmn @2c
Kijmn ¼ ¼ r ðaÞ
: ð5:41Þ
@eij @eij @Dmn
The matrix Cijkl contains stiffness coefficients of the composite in its virgin mater-
(a)
ial state, whereas coefficients of Kijmn are functions determining the change of state
caused by the internal dissipative mechanisms. The rate shown in Eq. (5.39)
ðaÞ
requires formulation of the scalar function c and the tensor components D_ kl .
The discussion of damage evolution is left for later; the next section will treat the
stress–strain relations at a fixed damage state.

5.2.4 Stiffness–damage relationships


The stiffness coefficient matrix Cijkl of the composite in a given state of damage
is derivable from the Helmholtz free energy function c according to Eq. (5.40).
The scalar valued function c can be written as a polynomial in its variables, i.e.,

ðaÞ
c ¼ cP eij ; Dij ; ð5:42Þ

where cP stands for the polynomial function.


Let us now focus on the particular case of intralaminar cracking in composite
laminates. Figure 5.5 shows an RVE illustrating one set of intralaminar cracks in
an off-axis ply of a composite laminate. Although for clarity of illustration the
cracking is shown only in one lamina, it is understood that in general it exists in
multiple plies of the laminate. The thickness of the cracked plies is denoted by tc, s
is the average crack spacing, t is the total laminate thickness, and W and L
stand for the width and the length, respectively, of the RVE. The volume of the
RVE, the surface area of a crack, S, and the influence vector magnitude, a, are
specified as
V ¼ L:W:t;
W:tc
S¼ ; ð5:43Þ
sin y
a ¼ ktc ;
where k, called the constraint parameter, is an unspecified constant of (assumed)
proportionality between a and the crack size tc (also the cracked-ply thickness).
Here, y is a positive quantity so that the surface area is always positive. Assuming
a to be constant over the crack surface S, one gets from Eq. (5.22)

ðaÞ kt2c
Dij ¼ n i nj ; ð5:44Þ
st sin y
where ni ¼ ðsin y; cos y; 0Þ.
Expansion of the polynomial function in Eq. (5.42) can in general have infinite
terms, which will obviously present an impractical situation. One way to restrict
the functional form is by expanding the polynomial in terms that account for the
5.2 Continuum damage mechanics (CDM) of composite materials 149

Figure 5.5. A representative volume element illustrating intralaminar multiple cracking in


a general off-axis ply of a composite laminate.

initial material symmetry. This is done in the polynomial invariant theory by using
the so-called integrity bases [21]. Such bases have been developed for scalar
functions of various vector and tensor variables. For the case of two symmetric
second-order tensors, such as in Eq. (5.42), the integrity bases for orthotropic
symmetry are given by Adkins [22]. Considering a single damage mode, a ¼ 1, we
have the following set of invariant terms

e11 ; e22 ; e33 ; e223 ; e231 ; e212 ; e23 e31 e12 ;

D11 ; D22 ; D33 ; D223 ; D231 ; D212 ; D23 D31 D12 ;

e23 D23 ; e31 D31 ; e12 D12 ; ð5:45Þ

e31 e12 D23 ; e12 e23 D31 ; e23 e31 D12 ;

e23 D31 D12 ; e31 D12 D23 ; e12 D23 D31 :

For the sake of applying the constitutive theory to thin laminates where only
in-plane strains are of interest, and for small strains, the expansion of the function
c (Eq. (5.42)) can be restricted to no more than quadratic terms in strain compon-
ents e11, e22, and e12
To what extent the damage tensor components are to be
taken in the expansion depends on the nature and amount of information that can
150 Macro-damage mechanics

be acquired for evaluation of the material constants that will appear in the
polynomial function. This issue will be discussed later. To begin with the simplest
possible case, we will include only linear terms in D11, D22, and D12, which are the
nonzero components for intralaminar cracks. Thus the set of invariant terms
reduces to
e1 ; e2 ; e26 ;

D1 ; D2 ; D26 ; ð5:46Þ
e6 D 6 ;
where e1  e11 ; e2  e22 ; e6  e12 ; D1  D11 ; D2  D22 ; D6  D12 . The most general
polynomial form for the Helmholtz free energy, restricted to second-order terms
in the strain components and first-order terms in damage tensor components, is
given by
rc ¼ P0 þ c1 e21 þ c2 e22 þ c3 e26 þ c4 e1 e2

þ c5 e21 D1 þ c6 e21 D2 þ c7 e22 D1 þ c8 e22 D2 þ c9 e26 D1 þ c10 e26 D2



ð5:47Þ
þ fc11 e1 e2 D1 þ c12 e1 e2 D2 g þ fc13 e1 e6 D6 þ c14 e2 e6 D6 g
þ P1 ðep ; Dq Þ þ P2 ðDq Þ;
where P0 and ci, i ¼ 1, 2, . . ., 14 are material constants, P1 is a linear function of
strain and damage tensor components, and P2 is a linear function of damage
tensor components. Setting the free energy to zero for unstrained and undamaged
material, we have P0 ¼ 0, and assuming the unstrained material of any damaged
state to be stress free, we get P1 ¼ 0. The stress components in the Voigt notation
are now given by (from Eq. (5.30))
@c
sp ¼ r ; ð5:48Þ
@ep
where p ¼ 1, 2, 6. A differential in stress can now be written as

@c @c
dsp ¼ r deq þ r dDr ¼ Cpq deq þ Kpr dDr ; ð5:49Þ
@ep @eq @ep @Dr
where
@c
Cpq ¼ r ð5:50Þ
@ep @eq

is the stiffness matrix when dDr ¼ 0, i.e., at constant damage. This is illustrated for
the uniaxial stress–strain response in Figure 5.6. As seen there, the elastic modulus
at any point on the stress–strain curve is the secant modulus, not the tangent
modulus.
Combining Eqs. (5.47), (5.48), and (5.50), one obtains

Cpq ¼ C0pq þ Cð1Þ


pq ð5:51Þ
5.2 Continuum damage mechanics (CDM) of composite materials 151

E0

Figure 5.6. Stress–strain curve of a composite with damage. The secant modulus E varies
with the state of damage.

where

E0x n0xy E0y


2 3
0 0
0 7
2
2c1 c4 0 6 1  nxy nyx
3 6 1  n0xy n0yx 7
6 7
C0pq ¼ 4 2c2 0 5¼6 E0y ð5:52Þ
6 7
7
6 0 7
Symm 2c3 6
4 1  n0xy n0yx 7
5
0
Symm Gxy

represents the orthotropic stiffness matrix for virgin composite material, in which
E0x ; E0y ; n0xy ; G0xy are effective moduli for the undamaged laminate, and

2c5 D1 þ 2c6 D2 c11 D1 þ c12 D2


2 3
c13 D6
Cð1Þ
pq ¼4 2c7 D1 þ 2c8 D2 c14 D6 5; ð5:53Þ
6 7

Symm 2c9 D1 þ 2c10 D2

represents the stiffness change brought about by the damage entities of damage
mode 1.
It can be noted here that Eqs. (5.51)–(5.53) show linear dependence of the
stiffness properties on damage tensor components. This is the consequence of
including only linear terms in these components in the polynomial expansion of
the free energy function, Eq. (5.47). Including higher-order terms will add additional
constants ci, which will need to be evaluated. The evaluation procedure is described
below, but it is remarked here that the formulation of constitutive response is in
no way restricted only to linear dependence on the chosen damage measure.
152 Macro-damage mechanics

Case 1: Cracking in one off-axis orientation


From Eqs. (5.51)–(5.53), it can be seen that the presence of damage entities, even
for one orientation, removes the initial orthotropic symmetry. For the case of
intralaminar cracks in one orientation, as illustrated in Figure 5.5, the nonzero
components of the damage tensor Dij(1) are given from Eq. (5.44) by

ð1Þ kt2c
D1 ¼ D11 ¼ sin y;
st

ð1Þ kt2c cos2 y


D2 ¼ D22 ¼ ; ð5:54Þ
st sin y

ð1Þ kt2c
D6 ¼ D12 ¼ cos y:
st

Inserting these into Eq. (5.51) and using Eq. (5.52) and Eq. (5.54) we obtain
the stiffness matrix of the damaged composite laminate for a fixed state of
damage as

n0xy E0y
2 3
E0x
0 7
6 1  n0xy n0yx 1  n0xy n0yx
6
7
6 7
0
6 7
Cpq ¼6
6 Ey 7
0 7
1  n0xy n0yx
6 7
6 7
ð5:55Þ
4 5
0
Symm Gxy

2c5 þ 2c6 cot2 y c11 þ c12 cot2 y


2 3
c13 cot y
2
kt
þ c sin y6
6 7
2c7 þ 2c8 cot2 y c14 cot y 7:
st 4 5
Symm 2c9 þ 2c10 cot2 y

Case 2: Cross-ply laminates


For the special case of cross-ply laminates, y ¼ 90 , and hence

n0xy E0y
2 3
E0x
0 7
6 1  n0xy n0yx 1  n0xy n0yx
6 2 3
7 2a1 a4 0
2
6 7
7 ktc 6
E0y
6
Cpq ¼6 7þ 2a2 0 5 ð5:56Þ
7
st
4
6 0 7
1  n0xy n0yx Symm 2a3
6 7
6 7
4 5
Symm G0xy

where a1 ¼ c5 ; a2 ¼ c7 ; a3 ¼ c9 ; and a4 ¼ c11 . It can be observed that the ortho-


tropic symmetry is retained for intralaminar cracking in cross-ply laminates.
The engineering moduli can be derived from the following relationships
5.2 Continuum damage mechanics (CDM) of composite materials 153

C11 C22 C212 C11 C22 C212


Ex ¼ C22 ; Ey ¼ C11 ;
ð5:57Þ
nxy ¼ CC1122 ; Gxy ¼ C66 :
Thus, for cross-ply laminates with 90 -ply cracks,
" #2
n0xy E0y kt2c
þ a4
E0x kt2c 1  n0xy n0yx st
Ex ¼ þ 2 a1  ;
1  n0xy n0yx st E0y kt2c sin y
þ 2 a2
1  n0xy n0yx st
" #2
n0xy E0y kt2c
þ a4
E0y kt2c 1  n0xy n0yx st
Ey ¼ þ2 a2  ;
1  n0xy n0yx st E0x kt2c ð5:58Þ
þ2 a1
1  n0xy n0yx st

n0xy E0y kt2c


þ a4
1  n0xy n0yx st
nxy ¼ ;
E0y kt2
0 0
þ 2 c a2
1  nxy nyx st

kt2c
Gxy ¼ G0xy þ 2 a3 :
st

Evaluation of material constants


In the damage–stiffness relations Eq. (5.56) and Eq. (5.58), ai,i ¼ 1,2,3,4 are a set of
four phenomenological constants, which need to be determined to predict stiffness
degradation. As seen from Eq. (5.58), the shear modulus is uncoupled from the
other three moduli and thus can be treated independently. These phenomeno-
logical constants are material and laminate configuration dependent and can be
evaluated for a selected laminate by using data generated either experimentally or
by an analytical or a computational model. As an example, let the moduli of a
given damaged cross-ply laminate at a fixed state of damage, s ¼ s1, be given as
Ex ; Ey ; Gxy ; and nxy , then by using Eq. (5.56), we obtain
" #
s1 t Ex E0x
a1 ¼  ;
2kt2c 1  nxy nyx 1  n0xy n0yx
" #
s1 t Ey E0y
a2 ¼  ;
2kt2c 1  nxy nyx 1  n0xy n0yx
ð5:59Þ
s1 t h 0
i
a3 ¼ Gxy  Gxy ;
2kt2c
" #
s1 t nxy Ey n0xy E0y
a4 ¼ 2  :
ktc 1  nxy nyx 1  n0xy n0yx
154 Macro-damage mechanics

E1/E10
Gl./Ep. [0/903]s

1.0
Calculated
x
x Measured
x
0.9 x

0.8

0.7 x
x
x x x
x x
x x
0.6
Ply discount prediction

0.5
0 100 200 300 s
(MPa)

Figure 5.7. Variation of the longitudinal Young’s modulus with applied stress for a glass/
epoxy [0/903]s laminate. Source: [23].

In the above expressions it should be noted that while the values of ai are fixed for
a given composite laminate (that has been homogenized), the parameter k depends
on the ability of the cracks to perform surface displacements under an applied
mechanical impulse. Thus this parameter may be viewed as a measure of the
constraint to the crack surface separation imposed by the material surrounding
the crack. One way to view this is by considering a crack of a given size embedded
in an infinite isotropic material, in which case the crack surface separation is
unconstrained and can be calculated by fracture mechanics methods. When the
laminate geometry is finite and its symmetry is different from isotropic, the k
parameter will take a value less than that for the infinite isotropic medium. This
consideration allows us to assign k an undetermined value, say k0, for a reference
laminate under reference loading conditions, and evaluate a change from this
value for another crack orientation. This procedure will be discussed in more
detail later.
The CDM approach described here has been used to successfully predict
degradation in the longitudinal and transverse moduli and the Poisson’s ratio
for a variety of laminates, e.g., [0/903]s, [903/0]s, and [0/ 45]s as reported in [13–15,
17, 23]. Figure 5.7 shows the degradation of the longitudinal Young’s modulus
in a [0/903]s glass/epoxy laminate with the applied tensile stress. The predictions
agree with the observed values fairly well in the entire range of cracking.
The predictions by the ply discount method are found to overestimate the total
modulus reduction. The stress–strain curve constructed from the reduced modulus
is shown in Figure 5.8.
5.3 Synergistic damage mechanics (SDM) 155

s
(MPa)
250 Gl./Ep. [0/903]s

200

150

100

50

0
0 0.5 1.0 1.5 2.0 2.5 e (%)

Figure 5.8. Longitudinal stress–strain response for a glass/epoxy [0/903]s laminate.


Source: [23].

5.3 Synergistic damage mechanics (SDM)

The observation that the k-parameter (hitherto referred to as constraint param-


eter) may be viewed as a carrier of the local effects on damage entities within a
RVE, while the ai-constants are material constants, led to a number of studies to
explore prediction of elastic property changes due to damage in different modes.
To be sure, the elastic properties are the averages over appropriate RVEs.
At first it was found that from changes in Ex and nxy due to transverse cracking
in [0/903]s glass/epoxy laminates reported in [24] and assuming no changes in Ey,
changes in Ex for the same glass/epoxy of [0/90]s configuration could be predicted
with good accuracy. Also, in [0/ 45]s laminates of the same glass/epoxy, the
change in Ex could be predicted by setting D1 ¼ D2 (a good approximation,
supported by crack density data). These results have been reported in [14].
Later, a systematic study of the effect of constraint on the constraint parameter
was done by experimentally measuring the crack opening displacement (COD)
in [ y/902]s laminates [25] for different y-values. By relating these values to the
COD at y ¼ 90 normalized by a unit applied strain, the predictions of Ex and nxy
for different y could be made. Another study of the constraint effects was made
by examining [0/ y4/01/2]s laminates, where the ply orientation y was varied.
Once again, using experimentally measured COD for y ¼ 90 as the reference,
the k-parameter for other ply orientations was evaluated from the COD values
and Ex and nxy for different y were then predicted [33].
While the experimental studies supported the idea of using the constraint
parameter as a carrier of local constraints, the scatter in test data and the cost of
156 Macro-damage mechanics

testing do not make the experimental approach attractive. Therefore, another


systematic study of [0m/ yn/0m/2]s laminates was undertaken [28, 29] where com-
putational micromechanics was employed instead of physical testing. An elaborate
parametric study of the constraint parameter allowed developing a master curve
for elastic property predictions. The most recent study [30] examines damage
modes consisting of transverse ply cracks as well as inclined cracks of different
orientations in [0m/ yn/90r]s and [0m/90r/ yn]s laminates. The reference value of k
for y ¼ 90 , as well as changes in it for other ply orientations, are computed by
finite element models of representative unit cells. Predictions made using k as a
micromechanical carrier of local damage-induced perturbations are compared
with available experimental data for [0/90/45/+45]s laminates. This approach
of combining MIDM (local) and MADM (global) approaches is named synergistic
damage mechanics (SDM). The following sections describe the use of SDM. First
we take multidirectional laminates with cracking in two symmetric off-axis orien-
tations, followed by the case of cracking in three orientations.

5.3.1 Two damage modes


Let us consider a case where two modes of damage are active. In this case, the
irreducible integrity bases for cP are given by, with a ¼ 1 and 2,

e11 ; e22 ; e33 ; e223 ; e231 ; e212 ; e23 e31 e12 ;

ð1Þ 2 ð1Þ 2 ð1Þ 2




ð1Þ ð1Þ ð1Þ ð1Þ ð1Þ ð1Þ


D11 ; D22 ; D33 ; D23 ; D31 ; D12 ; D23 D31 D12 ;

ð2Þ 2 ð2Þ 2 ð2Þ 2




ð2Þ ð2Þ ð2Þ ð2Þ ð2Þ ð2Þ


D11 ; D22 ; D33 ; D23 ; D31 ; D12 ; D23 D31 D12 ;
ð5:60Þ
ð1Þ ð1Þ ð1Þ ð2Þ ð2Þ ð2Þ
e23 D23 ; e31 D31 ; e12 D12 ; e23 D23 ; e31 D31 ; e12 D12 ;

ð1Þ ð1Þ ð1Þ ð2Þ ð2Þ ð2Þ


e31 e12 D23 ; e12 e23 D31 ; e23 e31 D12 ; e31 e12 D23 ; e12 e23 D31 ; e23 e31 D12 ;

ð1Þ ð1Þ ð1Þ ð1Þ ð1Þ ð1Þ ð2Þ ð2Þ ð2Þ ð2Þ ð2Þ ð2Þ
e23 D31 D12 ; e31 D12 D23 ; e12 D23 D31 ; e23 D31 D12 ; e31 D12 D23 ; e12 D23 D31 :

For a thin laminate loaded in its plane, the above set can be reduced by consider-
ing only the in-plane strain and damage tensor components. Thus, the remaining
integrity bases in the Voigt notation are given by

e1 ; e2 ; e26 ;

2
2
ð1Þ ð1Þ ð1Þ ð2Þ ð2Þ ð2Þ
D1 ; D2 ; D6 ; D1 ; D2 ; D6 ; ð5:61Þ
ð1Þ ð2Þ
e6 D6 ; e6 D6 :
5.3 Synergistic damage mechanics (SDM) 157

Using the above set of integrity bases, the most general polynomial form for rc,
restricted to second-order terms in the strain components (assuming small strains)
and first-order terms in damage tensor components (assuming small volume
fraction or number density of damage entities in the RVE), is given by

rc ¼ P0 þ c1 e21 þ c2 e22 þ c3 e26 þ c4 e1 e2


n o
ð1Þ ð1Þ ð2Þ ð2Þ
þ e21 c5 D1 þ c6 D2 þ c7 D1 þ c8 D2
n o
ð1Þ ð1Þ ð2Þ ð2Þ
þ e22 c9 D1 þ c10 D2 þ c11 D1 þ c12 D2
n o
ð1Þ ð1Þ ð2Þ ð2Þ
þ e26 c13 D1 þ c14 D2 þ c15 D1 þ c16 D2 ð5:62Þ
n o
ð1Þ ð1Þ ð2Þ ð2Þ
þ e1 e2 c17 D1 þ c18 D2 þ c19 D1 þ c20 D2
n o n o
ð1Þ ð2Þ ð1Þ ð2Þ
þ e1 e6 c21 D6 þ c22 D6 þ e2 e6 c23 D6 þ c24 D6



þ P1 ep ; Dð1Þ
q þ P2 ep ; Dð2Þ
q þ P3 Dð1Þ
q þ P4 Dð2Þ
q ;

where P0 and ci, i ¼ 1, 2, . . ., 24 are material constants, P1 and P2 are linear


functions of strain and damage tensor components, and P3 and P4 are linear
functions only of the damage tensor components. Setting rc ¼ 0 for unstrained
and undamaged material, we have P0 ¼ 0; and assuming the unstrained material of
any damaged state to be stress free, we get P1 ¼ P2 ¼ 0 on using Eq. (5.48).
Considering the virgin material to be orthotropic and proceeding in a similar
manner as above, we obtain the following relations for the stiffness matrix of
the damaged laminate

Cpq ¼ C0pq þ Cð1Þ ð2Þ


pq þ Cpq ; ð5:63Þ
0
where p, q ¼ 1, 2, 6; Cpq is the stiffness coefficient matrix of the virgin laminate
given by Eq. (5.52), and the changes in stiffness brought about by the individual
(1) (2)
damage modes are represented by Cpq and Cpq , which are given by

ð1Þ ð1Þ ð1Þ ð1Þ ð1Þ


2 3
2c5 D1 þ 2c6 D2 c17 D1 þ c18 D2 c21 D6
6 7
Cð1Þ ¼ ð1Þ
2c9 D1 þ 2c10 D2
ð1Þ
c23 D6
ð1Þ
6 7
pq 6 7
4 5
ð1Þ ð1Þ
Symm 2c13 D1 þ 2c14 D2
ð5:64Þ
ð2Þ ð2Þ ð2Þ ð2Þ ð2Þ
2 3
2c7 D1 þ 2c8 D2 c19 D1 þ c20 D2 c22 D6
6 7
Cð2Þ
pq ¼6
ð2Þ
2c11 D1 þ 2c12 D2
ð2Þ
c24 D6
ð2Þ
7:
6 7
4 5
ð2Þ ð2Þ
Symm 2c15 D1 þ 2c16 D2
158 Macro-damage mechanics

(b)
(a)

Figure 5.9. Damage characterization for two damage modes: (a) normal crack spacing syn,
and axial crack spacing sy in a cracked ply; (b) directions of normal vectors for cracks
in +y- and –y-plies.

In general, for N damage modes, Eq. (5.63) can be written as


N
X
Cpq ¼ C0pq þ CðaÞ
pq : ð5:65Þ
a¼1

Let us now consider a special case of a general laminate


 undergoing
 damage in two
symmetrically placed damage modes, such as 0m = yn =’p s , with ’ restricted to
angles that do not cause ply cracking. In such laminates, an in-plane tensile
loading will produce an in-plane stress state in each off-axis ply consisting
of normal stresses along and perpendicular to fibers in that ply and a shear stress
in the plane of the ply. Depending on the values of y, ’, and ply properties,
the stress perpendicular to the fibers could be tensile or compressive. Thus,
on loading, an off-axis ply may or may not develop intralaminar cracks. When
y ¼ 90 , the matrix will undergo multiple cracking in the transverse plies.
For other cases of off-axis ply orientations, multiple cracking is typically observed
to occur for angles from 50 to 90 under an axial tensile load. However, it has
been observed that even in cases where these cracks do not initiate in the off-axis
plies, the laminate moduli change with the applied load due to shear stress-induced
damage within the plies.
The damage state subsequent to ply cracking in the +y and y plies can be
represented by two damage mode tensors. For off-axis ply cracking, it is more
convenient to rewrite the damage mode tensor defined in (5.44) in terms of the
normal crack spacing, syn ¼ sy sin y, where sy is the crack spacing in the axial
direction (see Figure 5.9) for a ply of orientation y. Accordingly, the damage
mode tensors are given by

ðaÞ kt2c
Dij ¼ ni nj : ð5:66Þ
syn t

With reference to Figure 5.9(b) where the orientations of the two damage modes
are shown, the damage mode elements are given by
5.3 Synergistic damage mechanics (SDM) 159

ð1Þ
a ¼ 1 : ni ¼ ðsin y; cos y; 0Þ ;
þ þ þ
ð1Þ ky t2c 2 ð1Þ ky t2c ð1Þ ky t2c
D1 ¼ þ sin y; D2 ¼ þ cos2 y; D6 ¼ þ sin y cos y;
syn t syn t syn t
ð5:67Þ
ð2Þ
a ¼ 2 : ni ¼ ðsin y;  cos y; 0Þ;
  
ð2Þ ky t2c 2 ð2Þ ky t2c ð2Þ ky t2c
D1 ¼  sin y; D2 ¼  cos2 y; D6 ¼   sin y cos y;
syn t syn t syn t

where the superscripts y+ and y– indicate variables for the +y and –y plies,
respectively. Assuming that the intensity and distribution of damage is the same
in both +y and –y plies, we have

þ  þ 
ky ¼ ky ¼ ky ; syn ¼ syn ¼ syn : ð5:68Þ

Substituting Eq. (5.67)–(5.68) into Eq. (5.64), we obtain

ð1Þ ð2Þ ky t2c 


ðc5 þ c7 Þsin2 y þ ðc6 þ c8 Þcos2 y ;

C11 þ C11 ¼ 2 y
sn t
ð1Þ ð2Þ ky t2c 
ðc9 þ c11 Þsin2 y þ ðc10 þ c12 Þcos2 y ;

C22 þ C22 ¼ 2 y
sn t
ð1Þ ð2Þ ky t2c 
ðc13 þ c15 Þsin2 y þ ðc14 þ c16 Þcos2 y ;

C66 þ C66 ¼ 2 y
sn t
ð5:69Þ
ð1Þ ð2Þ ky t 2 
¼ y c ðc17 þ c19 Þsin2 y þ ðc18 þ c20 Þcos2 y ;

C12 þ C12
sn t
ð1Þ ð2Þ ky t2c
C16 þ C16 ¼ sin y cos y½c21 þ c22  ¼ 0;
syn t
ð1Þ ð2Þ ky t2c
C26 þ C26 ¼ sin y cos y½c23 þ c24  ¼ 0 :
syn t

Thus,
2 3
2a1 D1 þ 2b1 D2 a4 D 1 þ b 4 D 2 0
Cð1Þ ð2Þ
pq þ Cpq ¼ 4 2a2 D1 þ 2b2 D2 0 5; ð5:70Þ
6 7

Symm 2a3 D1 þ 2b3 D2

where the superscripts for denoting damage mode have been dropped for convenience,
and ai and bi, i ¼ 1, 2, 3, 4 are the two sets of four material constants, given by
a1 ¼ c 5 þ c 7 ; a2 ¼ c9 þ c11 ; a3 ¼ c13 þ c15 ; a4 ¼ c17 þ c19 ;
ð5:71Þ
b1 ¼ c 6 þ c 8 ; b2 ¼ c10 þ c12 ; b3 ¼ c14 þ c16 ; b4 ¼ c18 þ c20 :

Here, ai and bi are functions of y. Denote


160 Macro-damage mechanics

a1 ðyÞ ¼ a1 sin2 y þ b1 cos2 y;

a2 ðyÞ ¼ a2 sin2 y þ b2 cos2 y;


ð5:72Þ
a3 ðyÞ ¼ a3 sin2 y þ b3 cos2 y;

a4 ðyÞ ¼ a4 sin2 y þ b4 cos2 y :

Then,
2 3
2a1 ðyÞ a 4 ð yÞ 0
Cð1Þ ð2Þ
pq þ Cpq ¼Dy 4 2a2 ðyÞ 0 5; ð5:73Þ
6 7

Symm 2a3 ðyÞ

where

ky t2c
Dy ¼ : ð5:74Þ
syn t
Rewrite Eq. (5.72) as
 
2 2 bi 2
2
ai ðyÞ ¼ ai sin y þ bi cos y ¼ ai sin y 1 þ cot y : ð5:75Þ
ai

Consider for the moment the case when ai  bi. Then,

bi p p
cot2 y 1 for y : ð5:76Þ
ai 4 2

Also, it can be expected that

bi p p
cot2 y  1 for y ;
ai 3 2
ð5:77Þ
p p
i:e:; ai ðyÞ  ai for y :
3 2
For this case, we have laminate stiffness matrix as
2 3
E0x n0xy E0y
0 7
6 1  n0xy n0yx 1  n0xy n0yx
6 2 3
6
7
7 2a1 a4 0
E0y
6 7 6 7
Cpq ¼6 7 þ Dy 6 2a2 0 75; ð5:78Þ
6 0 7 4
1  n0xy n0yx
6 7
6
4
7
5 Symm 2a3
0
Symm Gxy

which is of identical form to the expression for 90 cracking in a cross-ply


laminate, see Eq. (5.56). The moduli can be finally obtained using the relations
in Eq. (5.57). The resultant expressions will be of the same form as in Eq. (5.58),
except that ktc2/st is now replaced with Dy.
5.3 Synergistic damage mechanics (SDM) 161

COMPUTATIONAL MICROMECHANICS
Determine COD and constraint parameter(s)
(
κ (∆u 2 ±q n
b = κq = ) ; ( ∆u2 ) ±q n = ( ∆u2 ) +q n + ( ∆u2 )–q n
90 (∆u 2 90
8

Structural scale: Micro

EXPERIMENTAL/
COMPUTATIONAL SYNERGISTIC DAMAGE MECHANICS

Evaluate damage constants Use SDM to determine stiffness reduction in


using available data present laminate configuration [0m / ± qn /0m/2]s
for reference laminate
Structural scale: Meso
configuration [0/908 /01/2]s

STRUCTURAL ANALYSIS
Analyze overall structural response to external
loading using the reduced stiffness properties

Structural scale: Macro

Figure 5.10. Flowchart showing the multiscale synergistic methodology for analyzing
damage behavior in a class of symmetric laminates with layup [0m/ yn/0m/2]s containing
ply cracks in the +y and y layers.

The overall SDM procedure for this laminate is sketched in Figure 5.10.
As illustrated, it combines micromechanics with CDM for complete evaluation
of the structural response. Micromechanics involves analysis to determine CODs
in cracked plies within a RVE (or unit cell, if applicable) of [0m/ yn/0m/2]s layup,
from which the constraint effect is evaluated for different y and/or m.
The constraint effect is carried over in the CDM formulation through the con-
straint parameter. In a separate step, the damage constants ai are determined
from experimental data for a reference laminate, which is chosen here to be
[0/ 908/01/2]s. A cross-ply laminate of the same class as the ply layup and material
in consideration is a good choice for the reference laminate because either the
experimental data are often available or can be obtained by using any of the
damage models for 90 -ply cracking as described in the previous chapter. For
evaluation of damage constants, expressions in Eq. (5.59) shall be used after
replacing ktc2/st by Dy. With the values of the damage constants and b ¼ ky =k90
known, the stiffness–damage relations given by Eq. (5.78) are employed to predict
stiffness degradation with crack density. The overall structural behavior can be
finally analyzed using the reduced stiffness properties for the laminate.
For computation of average COD values, micromechanics can be performed
using a suitable FE analysis. The 3-D FE model for [0m/ yn/0m/2]s layup, along
162 Macro-damage mechanics

Figure 5.11. Representative unit cell for COD computation for laminates. Reprinted, with
kind permission, from Int J Solids Struct, Vol. 45, C.V. Singh and R. Talreja, Analysis of
multiple off-axis ply cracks in composite laminates, pp. 4574–89, copyright Elsevier (2008).

6
COD (µm)

2
FEM
Experiment

0
20 30 40 50 60 70 80 90
Ply orientation (q)

Figure 5.12. Variation of average COD with change in ply orientation for a [0m/ yn/0m/2]s
glass/epoxy laminate. Reprinted, with kind permission, from Int J Solids Struct, Vol. 45,
C.V. Singh and R. Talreja, Analysis of multiple off-axis ply cracks in composite laminates,
pp. 4574–89, copyright Elsevier (2008).

with boundary conditions and coordinate systems, is shown in Figure 5.11 [29, 31,
32]. As can be seen from Figure 5.12, the computed CODs averaged over the
thickness of the cracked plies agree quite well with the experimental data over
5.3 Synergistic damage mechanics (SDM) 163

(a) (b)

103 u2/tc

103 u2/tc
(c) (d)
103 u2/tc

103 u2/tc

Figure 5.13. COD profiles for cracked plies in [0/ y4/01/2]s laminates: (a) CODs averaged
over +y- and –y-plies; (b) Crack profile for 90º-ply crack compared with an elliptic profile
for an isotropic medium; (c), (d) COD profiles for +y and –y separately: (c) y = 70º,
(d) y = 40º. Parts (c), (d) depict the asymmetry of opening displacements for off-axis
laminates, especially at a ply orientation farther from y = 90 . Reprinted, with kind
permission, from Int J Solids Struct, Vol. 45, C.V. Singh and R. Talreja, Analysis of multiple
off-axis ply cracks in composite laminates, pp. 4574–89, copyright Elsevier (2008).

the whole range of ply orientations considered. This suggests that the 3-D FE analysis
is an accurate tool to evaluate the constraint parameter. The profiles of COD nor-
malized with tc through the ply thickness are shown in Figure 5.13. As expected, for
cross-ply laminates, the profile is symmetric about mid-plane of the cracked ply and
consequently the maximum COD occurs at the mid-plane of the cracked layer.
However, due to the difference in constraint from the surrounding material, this
COD profile is different from an elliptical profile for a single crack in an infinite
isotropic elastic medium subjected to a uniform far-field stress (see Figure 5.13 (b)).
Thus, the magnitude of the average COD for a 90 crack is different from that for an
elliptical crack. For other ply orientations, the crack surface displacements are not
symmetric about the mid-plane as shown in Figure 5.13 (c)–(d). This asymmetry
increases as we go away from cross-ply laminates (y ¼ 90 ) and maximum COD does
not occur midway through the  thickness of the cracked y layers. The aspect ratio of
COD profiles, g ¼ ðu2 Þmax u2 , varies from 1.33 for y ¼ 90 to 1.40 for y ¼ 40
(Figure 5.13 (a)), which is different from the aspect ratio of 1.273 ( ¼ 4/p) for an
164 Macro-damage mechanics

(a) (b)
θ = 70° θ = 70°

Ex nxy
E x0 0
nxy

(c) (d)
θ = 55° θ = 55°

Ex nxy
E x0 0
nxy

Figure 5.14. Variation of longitudinal modulus and Poisson’s ratio for [0/ y4/01/2]s
laminates: (a), (b) for y = 70 ; (c), (d) for y = 55 . Reprinted, with kind permission, from
Int J Solids Struct, Vol. 45, C.V. Singh and R. Talreja, Analysis of multiple off-axis ply
cracks in composite laminates, pp. 4574–89, copyright Elsevier (2008).

elliptic profile. Here (Du2)max represents the maximum COD and u2 represents the
COD averaged over the thickness direction. Numerically, the COD, Du2, is computed
by taking the difference of the node displacements on either side of the crack along a
direction transverse to the local fiber direction (assuming the crack traverses parallel
to the fiber).
The elastic moduli Ex and nxy predicted by the SDM approach are compared
with the experimental data for y ¼ 70 and 55 in Figure 5.14. The agreement with
data is about the same as that obtained by using experimentally measured COD
values to evaluate the constraint parameter. For y ¼ 55 , the stiffness reduction
is caused by matrix cracking as well as shear-induced damage in off-axis plies, as
discussed in [33], where a procedure for calculating the stiffness reduction due
to shear damage was described. SDM predictions for this orientation thus
include both effects. It is interesting to note that the stiffness predictions using
experimentally measured COD values are subject to scatter, which is inherent in
testing. Furthermore, the experimental procedure requires a specialized test set-up
[33, 34], which is costly and takes a certain amount of training to operate. The 3-D
FE computations on a unit cell, on the other hand, are easy to perform and can
provide accurate values of CODs.
The SDM approach is quite handy when the relative stiffness or thickness
of cracked and supporting plies change. The difference in stiffness changes for
5.3 Synergistic damage mechanics (SDM) 165

laminates with different ply thickness or axial stiffness are due to difference in the
relative constraint from the supporting plies, and thus can be characterized
by the changes in the constraint parameter. In [29] a parametric study was
conducted to study the effect 90 of changes in ply thicknesses (m and n) and y
axial stiffness ratio r ¼ E y
A EA over CODs of [0m/ yn/0m/2]s laminates, where EA
and EA90 represent the axial stiffness of y-plies and the 90 -plies, respectively.
Figure 5.15 shows the variation of average COD for changes in these parameters.
The computed average CODs are given by the following parametric equation
 
u2 yn
¼ U:f1 ðyÞ:f2 ðr Þ:f3 ðmÞ:f4 ðnÞ; ð5:79Þ

where U is the average COD for the reference laminate [0/908/01/2]s, and

f1 ðyÞ ¼ sin2 y;
f2 ðr Þ ¼ r c1 ;
c2 ð5:80Þ
f3 ðmÞ ¼ þ c3 ;
m
f 4 ð n Þ ¼ c 4 n c5 ;
are fitted functions where fitting constants (unitless) c1–c5 are given as: c1 ¼
0.0871; c2 ¼ 0.1038; c3 ¼ 0.8949; c4 ¼ 0.247; c5 ¼ 0.99 for [0m/ yn/0m/2]s glass/
epoxy laminates.
The parametric study described above enables us to predict stiffness degrad-
ation in off-axis laminates with different geometry and stiffness values. For
example, one can consider a laminate with stiffer outer plies. The variation of
engineering moduli Ex and nxy for different stiffness ratios (r) for a [0/ 704/01/2]s
laminate is shown in Figure 5.16(a). As expected, stiffer outer plies cause less
severe degradation in the moduli. In contrast to changing the stiffness of outer
plies, one can vary the number of constraining plies (m) or cracked plies (n). The
effect of number of cracked plies on the change in stiffness moduli is shown in
Figure 5.16(b). These results indicate that the cracking ply thickness, i.e., crack
size, has significant effect on stiffness degradation while the thickness of the
constraining plies as well as the change in axial stiffness ratio r have small effect.

5.3.2 Three damage modes


We will now consider a special case of cracking in three off-axis plies, of which two
are symmetrically opposite of each other, and the third is placed along the
transverse direction. Thus cracking is in the +y, –y, and 90 plies. For this case
the stiffness matrix of the damaged laminate is given by

Cpq ¼ C0pq þ Cð1Þ ð2Þ ð3Þ


pq þ Cpq þ Cpq ; ð5:81Þ
0 (1) (2)
where Cpq is given by Eq. (5.52), and Cpq + Cpq is given by Eq. (5.73). The components
of the third damage mode (a ¼ 3) corresponding to cracking in 90 -ply are
166 Macro-damage mechanics

(a)

(b)

(c)

Figure 5.15. Variation of average COD for [0m/ yn/0m/2]s glass/epoxy laminates with
(a) axial stiffness ratio, r (for m = 1, n = 4); (b) number of cracked plies, n; and
(c) number of constraining plies, m. Reprinted, with kind permission, from Int J Solids
Struct, Vol. 45, C.V. Singh and R. Talreja, Analysis of multiple off-axis ply cracks in
composite laminates, pp. 4574–89, copyright Elsevier (2008).
5.3 Synergistic damage mechanics (SDM) 167

(a) (b)
1.2 1.2

1 1

0.8 0.8
Ex 0.6 nxy
0.6
0 nxy
0
Ex
0.4 r = 0.5 0.4 r = 0.5
r=1 r=1
0.2 r=2 0.2 r=2
r=5 r=5
0 0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
Crack density (1/mm) Crack density (1/mm)
(c) (d)
1.2

1 1.2

0.8 1

Ex 0.8
0.6
0 nxy
Ex 0.6
0.4 nxy
0
n=1
n=2 0.4 n=1
0.2 n=4 n=2
0.2 n=4

0 0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
Crack density (1/mm) Crack density (1/mm)

Figure 5.16. Effect of change in axial stiffness ratio (r): (a), (b), and cracked ply thickness (n):
(c), (d), on stiffness degradation in [0m/ 70n/0m/2]s glass/epoxy laminates. Reprinted, with
kind permission, from Int J Solids Struct, Vol. 45, C.V. Singh and R. Talreja, Analysis of
multiple off-axis ply cracks in composite laminates, pp. 4574–89, copyright Elsevier (2008).

ð3Þ k90 t290 ð3Þ ð3Þ


D1 ¼ ; D2 ¼ D6 ¼ 0: ð5:82Þ
s90 t
The integrity bases (Eq. (5.61)) have additional components for D(3)
1. The free
energy function thus gets the following terms added to Eq. (5.62)
ð3Þ ð3Þ ð3Þ ð3Þ
rcða ¼ 3Þ ¼ a01 e21 D1 þ a02 e22 D1 þ a03 e26 D1 þ a04 e1 e2 D1 ; ð5:83Þ

where a0i ; i ¼ 1; 2; 3; 4 are additional material constants. Substituting Eq. (5.83)


into Eq. (5.48), we obtain

2a01 a04
2 3
0
ð3Þ 6
Cð3Þ
pq ¼D1 4 2a02 0 5; ð5:84Þ
7

Symm 2a03

where the contribution to the shear components is zero.


It is important to emphasize here that the relative location of different
damage modes in the whole laminate will cause different losses in stiffness due to
damage in the laminate. To illustrate this let us consider two specific examples of
laminates with damage modes considered in the present section, viz., +y, –y, and 90 .
Figure 5.17(a)–(b) shows representative unit cells for laminates with [0m/ yn/90r]s
168 Macro-damage mechanics

(a) z (b) z

0⬚ layer
0⬚ layer
(u)x= 0 = 0 (u)x= 2 1 = u0 (u)x= 0 = 0 90⬚ layer (u)x= 2 1 = u0
+q layer
+q layer
y –q layer y

90⬚ layer –q layer


x x
(w)z = 0 = 0 (w)z= 0 = 0
(symmetry about mid-plane) (symmetry about mid-plane)

Figure 5.17. Representative unit cell for (a) [0m/ yn/90r]s and (b) [0m/90r/ yn]s laminate
configurations.

and [0m/90r/–yn/+yn]s configurations, respectively. The boundary conditions are


shown for the 3-D FE computation of average COD values. The global laminate
(X, Y, Z) and local crack plane (x1, x2, x3) coordinate systems are also shown.
Considering first the case of a [0m/ yn/90r]s laminate, we note that y modes occur
twice in the whole laminate, above and below the mid-plane N of the laminate, whereas
ðaÞ
90 mode occurs only once. Thus, Cpq ¼ Cpq  C0pq ¼
P
Cpq is given by
n o a¼1
Cpq ¼ 2 Cð1Þpq ðþy Þ þ Cð2Þ
pq ð y Þ þ Cð3Þ
pq ð90Þ : ð5:85Þ

Collecting terms from Eq. (5.73) and Eq. (5.84), while assuming the ai’s to be
independent of y, we get

2a0 1 a0 4
2 3 2 3
2a1 a4 0 0
Cpq ¼ 2Dy 4 2a2 0 5 þ D90 4 2a0 2 0 5; ð5:86Þ
Symm 2a3 Symm 2a0 3

where, for the laminate configuration considered,


ky ð2nt0 Þ2 k90 ð2rt0 Þ2
Dy ¼ ; D90 ¼ ; ð5:87Þ
syn t s90 t
where t0 denotes the thickness of a single ply. The special case when y ¼ 90 , and
DCpq as given by Eq. (5.86) should be equal to that given by a single 90 mode with
crack size of (4n+2r) plies. Consider, for example, the DC11 term. If we assume
that the normal crack spacing is the same in all cracked plies, then DC11 for a
[0m/ yn/90r]s laminate at y ¼ 90 from Eq. (5.86) is given by
n o
ð1Þ ð2Þ ð3Þ
C11 ¼ 2 C11 þ C11 þ C11

¼ 4Dy jy¼90 a1 ð90Þ þ 2D90 a0 1

4ky jy¼90 ð2nt0 Þ2 2k90 ð2rt0 Þ2 0 ð5:88Þ


¼ a 1 þ a1
s90 t s90 t
8t20  2
2n ky jy¼90 a1 ð90Þ þ r 2 k90 a0 1 :

¼ 90
s t
5.3 Synergistic damage mechanics (SDM) 169

Since, for y ¼ 90 , [0m/ yn/90r]s is equivalent to [0/902n+r]s, we can consider their


stiffness changes to be the same. Thus, DC11 can also be written, using Eq. (5.84), as

ð3Þ k904nþ2r ½ð4n þ 2r Þt0 2 0


C11 ¼ C11 ¼ 2a01 D1 ¼ :2a1 ; ð5:89Þ
s90 t
where the sub-subscript “4n+2r” in k904nþ2r represents the crack size for the 90 mode
in a [0/902n+r]s laminate. Equating DC11 from Eq. (5.88) and Eq. (5.89), we have

2n2 ky jy¼90 a1 ð90Þ þ r 2 k90 a01 ¼ k904nþ2r ½ð2n þ r Þt0 2 a01 ; ð5:90Þ
i.e.,
k904nþ2r ð2n þ r Þ2  r 2 k90 0
a1 ð90Þ ¼ a1 : ð5:91Þ
2n2 ky jy¼90
Generalizing, we can write the interrelationship between two sets of constants as

k904nþ2r ð2n þ r Þ2  r 2 k90 0


ai ¼ ai i ¼ 1; 2; 3; 4: ð5:92Þ
2n2 ky jy¼90
Substituting Eq. (5.92) into Eq. (5.86), DCpq for a damaged [0m/ yn/90r]s laminate
is given by

2a01 a0 4 0
2 3

Cpq ¼ D4 2a02 0 5; ð5:93Þ


6 7

Symm 2a03
where

4t20 1 ky n
 
2 k90
o
 2 2
D¼ ð2n þ r Þ k904nþ2r  r k90 þ r 90 ; ð5:94Þ
t syn ky jy¼90 s
where the constraint parameters are given by
     
uy y2n
uy 904nþ2r uy 902r
ky ¼ ; k904nþ2r ¼ ; k90 ¼ : ð5:95Þ
2nt0 ð4n þ 2r Þt0 2rt0

Consider now the case of the [0m/90r/ yn]s laminate configuration. It must be
noted that, unlike [0m/ yn/90r]s laminates, y damage modes in this case are
centrally placed, thereby the corresponding equivalent crack size is 4nt0 (with
averaging over two +y and two –y layers). On the other hand, the crack size for
the 90 damage mode is rt0. The derived stiffness–damage relationships retain the
form of Eq. (5.93). However, D in this case is given by

2t20 1 ky n
 
2 k90
o
 2 2
D¼ 2ð2n þ r Þ k904nþ2r  r k90 þ r 90 ð5:96Þ
t syn ky jy¼90 s

and the corresponding constraint parameters are now given by


170 Macro-damage mechanics

     
uy y2n
uy 904nþ2r uy 90r
ky ¼ ; k904nþ2r ¼ ; k90 ¼ : ð5:97Þ
4nt0 ð4n þ 2r Þt0 rt0

The SDM procedure for this laminate configuration is quite similar to that
described earlier for the [0m/ yn/0m/2]s layup in Figure 5.10. The damage constants
are evaluated from Eq. (5.59) after replacing ktc2/st by D by using the stiffness
degradation data for the reference laminate, which is chosen as [0/903]s in this case.
Depending on the specific 90 -ply placement, the corresponding constraint par-
ameters are calculated using expressions in Eq. (5.95) or Eq. (5.97) by using the
CODs computed from a suitable 3-D FE model. FE modeling details and COD
computations are described in detail in [30, 32]. One important consideration in
COD computation is the interaction between the families of cracks in different
orientations which can modify the crack displacements and consequently the
stiffness changes due to ply cracking [30]. This interaction can also modify the
damage evolution as illustrated in the next chapter. In Figure 5.18, SDM predic-
tions of stiffness moduli of [0/ 70/90]s and [0/ 55/90]s glass/epoxy laminates are
compared with the independent calculations from 3-D FE analysis (see [30–32]
for details on procedure to calculate stiffness changes using 3-D FE analysis).
The SDM predictions for a quasi-isotropic laminate are shown in Figure 5.19.
In the experiments it was observed that the laminates failed before the cracks in
the plies could grow fully through the laminate width. The effect of partially
grown cracks on stiffness changes is accounted for in the SDM analysis by
reducing the crack density for layers by a “relative density factor,” defined as

Actual surface area for partial cracks


rr ¼ : ð5:98Þ
Surface area for full cracks
To find the actual surface area of partial cracks, the information regarding their
actual length (along lamina width) is necessary. Since such data were not reported
in the above experimental study [35], the results are presented for two assumed
values of rr: 0.25 and 0.5.

5.4 Viscoelastic composites with ply cracking

Polymers used as matrix materials in fiber-reinforced composites usually display


viscoelastic behavior. However, the time-dependent behavior of the matrix does
not translate fully in the composite and while it is isotropic in the polymer, the
directionality of the fibers makes it anisotropic in the composite. Viscoelasticity
models for composite materials, particularly for high temperature applications
where the time dependency is more pronounced, are needed. The composite time-
dependent response in most cases can be approximated by linear viscoelasticity.
When ply cracking occurs the viscoelastic response can still be linear but with
modified time-dependency.
5.4 Viscoelastic composites with ply cracking 171

(a)

θ = 70⬚ θ = 70⬚

(b)
θ = 55⬚ θ = 55⬚

Figure 5.18. Variation of longitudinal modulus and Poisson’s ratio for [0/ y/90]s laminates:
(a) y = 70 , (b) y = 55 . The results are compared with independent FE calculations. SDM
predictions are made for two cases: no interaction between y and 90 cracks, and with
maximum interaction between them. Reprinted, with kind permission, from Mech Mater,
Vol. 41, C.V. Singh and R. Talreja, A synergistic damage mechanics approach for composite
laminates with matrix cracks in multiple orientations, pp. 954–68, copyright Elsevier (2009).

(a) (b)
1.1 1.1
+45⬚ cracks +45⬚ cracks
1 –45⬚ cracks 1 –45⬚ cracks
[0/90/ ⫿45]s
[0/90/ ⫿45]s
0.9 0.9
nxy
Ex 0.8 0.8 [0/90]s
0
nxy
E 0x [0/90]s
0.7 0.7

Test 1: [0/90/–45/+45]s
0.6 Test 2: [0/90/–45/+45]s 0.6 Test 1: [0/90/–45/+45]s
Test 3: [0/90/–45/+45]s Test 2: [0/90/–45/+45]s
Test 1: [0/90]s Test 3: [0/90/–45/+45]s
0.5 Test 2: [0/90]s 0.5 Test 1: [0/90]s
Test 3: [0/90]s Test 2: [0/90]s
Test 4: [0/90]s Test 3: [0/90]s
0.4 0.4
0 0.2 0.4 0.6 0.8 1 1.2 0 0.2 0.4 0.6 0.8 1 1.2
90⬚-Crack density (1/mm) 90⬚-Crack density (1/mm)

Figure 5.19. Stiffness reduction for quasi-isotropic [0/90/45/+45]s laminate compared with
experimental data [35]: (a) longitudinal modulus, (b) Poisson’s ratio. Solid lines represent
SDM predictions using assumed values of rr. Reprinted, with kind permission, from Mech
Mater, Vol. 41, C.V. Singh and R. Talreja, A synergistic damage mechanics approach for
composite laminates with matrix cracks in multiple orientations, pp. 954–68, copyright
Elsevier (2009).
172 Macro-damage mechanics

One approach to treating the coupling of viscoelasticity and damage in com-


posite materials is by internal variables based continuum models. Most models of
this nature have been applied to particulate composites. One model that
addresses composite laminates is by Schapery and Sicking [36]. This model
extends the concept of work potential developed for elastic solids to viscoelastic
composite laminates by using the pseudo variables defined by Schapery [37].
However, the Schapery–Sicking model makes certain assumptions that limit the
usefulness of the model. It is assumed that the shear modulus and the transverse
modulus of a ply are linearly related, and that the ply is transversely isotropic.
Furthermore, modeling is done at the ply level, neglecting the constraint effects
of the neighboring plies. The model developed later by Kumar and Talreja [38]
does not suffer from these limitations. The CDM-based modeling of linear
viscoelastic response of composites with damage in accordance with this model
will be discussed below.
Using the well-known Boltzman superposition principle the constitutive law for
a viscoelastic material can be represented as
ðt
@ekl
sij ¼ Cijkl ðt  tÞ dt; ð5:99Þ
0 @t
where Cijkl(t) is the relaxation modulus tensor. Taking the Laplace transform of
both sides and using the convolution theorem, we obtain
ij ¼ C~ijklekl ;
s ð5:100Þ
where the Laplace transform fðsÞ of a function f(t) is defined as
1
ð
fðsÞ ¼ est f ðtÞ dt; ð5:101Þ
0

and

C~ijkl ¼ sCijkl : ð5:102Þ

The constitutive equations (5.100) in the Laplace domain are similar to


the relationships of linear elasticity, except for the dependency on the trans-
formation parameters. Additionally, the equilibrium equations and strain–dis-
placement equations also retain their form in the Laplace domain. This
correspondence between linear elasticity and linear viscoelasticity is the so-
called Correspondence Principle [39], and it applies to solving boundary
value problems as long as the boundary conditions do not change in time.
Thus, if ply cracks are present in a composite laminate, then the Correspond-
ence Principle can still be applied by viewing the cracked laminate as a homo-
geneous solid in which the internal boundaries (cracks) retain the same
boundary conditions in time. This implies that the cracks do not grow or heal
in time. We shall therefore address the linear viscoelastic response of laminates
5.4 Viscoelastic composites with ply cracking 173

at a given (fixed) damage state. Specifically, we will express the relaxation


modulus tensor in Eq. (5.99) as a function of damage.
Following [38] a pseudo energy density function is defined in the Laplace
domain in terms of transformed strain and damage variables, such that

@W eij ; DðaÞ


ij
ij ¼
s : ð5:103Þ
@eij

It is noted that the above equation is valid for a fixed ply crack density. The time-
dependent deformation of the plies themselves will result in time-varying con-
straint on the cracks, leading to the crack surface separation changing in time.
For illustration, let us consider a cross-ply laminate with transverse cracks,
Figure 5.5, with y ¼ 90 . In this case the damage mode tensor has a single nonzero
component, which in the Laplace domain takes the form [38]

ðsÞt902
k
D11 ¼ : ð5:104Þ
st t T
where k ðsÞ represents the constraint parameter in the Laplace domain, t90 is
the thickness of the cracked layer, tT is the total laminate thickness, and st is the
crack spacing. Using the same procedure as in Section 5.1.4, Kumar and Talreja
[38] derived the following transformed constitutive relationships (in the Voigt
notation) as
8 9 2~ ~
38 9
s 0 > e1 >
< 1>
> = 6 C11 C12 7< =
2 ¼ 6 ~ ~
>
s 4 C12 C22 0 75> e2 >; ð5:105Þ
: >
0 C~66
; : ;
s6 0 e6

where

C~pq ¼ C~0pq þ C~ð1Þ


pq : ð5:106Þ

Here C~0pq is the transformed relaxation modulus of the cross-ply laminate without
ð1Þ
cracks and C~pq , given by the following, is transverse cracking (one mode of
damage, a ¼ 1)
2 3
2g11 g12 0
C~ð1Þ ¼ D 4 g12 2g22 0 5; ð5:107Þ
6 7
pq 11
0 2g66

where g11, g12, g22, and g66 are material constants that appear as coefficient terms
in the polynomial expansion of the pseudo strain energy density function W. The
damage function D11 is given by Eq. (5.104).
The time-dependency of the crack opening displacement has been handled in
two ways. In Kumar and Talreja [38] the function W was assumed to depend on
174 Macro-damage mechanics

the initial (t ¼ 0) value of the damage tensor, thereby absorbing all time depend-
ency in the presence of damage into the coefficient terms g11, etc. Later Varna et al.
[41] chose to retain the time dependency in the damage tensor and dealt with the
time-varying constraint to the COD explicitly.
Taking the inverse Laplace transform of Eq. (5.106), one obtains

2rn t90
Cpq ðtÞ ¼ C0pq ðtÞ þ kij ðtÞ; ð5:108Þ
tT

where

t90
rn ¼ ð5:109Þ
2st

is the normalized crack density, and

k11 ðtÞ ¼ 2L1 ðk


g11 =sÞ; k12 ðtÞ ¼ L1 ðk
g12 =sÞ;
ð5:110Þ
k22 ðtÞ ¼ 2L1 ðk
g22 =sÞ; k66 ðtÞ ¼ 2L1 ðk
g66 =sÞ ;

where L–1(*) represents the inverse Laplace transform. Note that k ¼ k0 , its value
at time t ¼ 0, if all time dependency is assumed to be in g11, etc. in the above
equations.
The functions k11, k12, and k22 in Eq. (5.110) can be determined by procedures
similar to those described above for elastic constants. Here these three unknown
functions, which are decoupled from k66, can be evaluated from the time-varying
differences of the axial modulus Ex(t) and Poisson’s ratio nxy(t) from their initial
values, and by assuming no change in the transverse modulus Ey(t), as done in
Kumar and Talreja [38]. The time-variation of the material constants can be
determined from experimental data, if available, or by a micromechanics approxi-
mation. The micromechanics can be analytical, if possible, or numerical, e.g., by a
finite element model.
In Kumar and Talreja [38] cross-ply laminates of given linear viscoelastic ply
properties were considered for validation of the CDM approach described above.
The functions k11(t), k12(t), and k22(t) were determined from the calculated visco-
elastic response of a [0/902]s laminate at a fixed crack density of 0.4 cracks/mm.
These functions were then used to predict the time variations of the axial modulus
and Poisson’s ratio at other crack densities and for other cross-ply laminate
configurations of the same material. The predictions were compared with inde-
pendently calculated time variations of the properties by a finite element model
and an analytical micromechanics model reported in Kumar and Talreja [40].
Predictions agreed well in all cases.
Varna et al. [41] demonstrated the use of SDM for linear viscoelastic response
predictions by explicitly treating the COD variations in time. The time dependency
of COD, i.e., the constraint parameter k and its counterpart in the Laplace
domain k, were calculated by a FE model. These were then inserted in
5.4 Viscoelastic composites with ply cracking 175

Eq. (5.110) to determine k11(t), k12(t), and k22(t). Predictions thus made agreed
well with independently calculated viscoelastic response at different crack densities
and different cross-ply laminate configurations. A parametric study was also
performed in Varna et al. [41] to determine a master function for COD variation.
That function has the following form

E2 ð t Þ d
   
tc
kðtÞ ¼ a þ b þ c 1 ; ð5:111Þ
2ts E1 ð t Þ

where a, b, c, and d are constants, E1 and E2 are the axial and transverse values,
respectively, of the Young’s modulus of the ply material, and ts here is the
thickness of the 0 -plies in the cross-ply laminate. Note that if more generally a
laminate of configuration [S/90n]s is used, then ts will be the thickness of the
sublaminate denoted by S.
For the nonlinear viscoelastic response of composites with damage another
approach is needed since the Correspondence Principle does not apply. A CDM
approach for this case was developed by Ahci and Talreja [42]. The material
system in the experimental study performed there was a carbon fiber (T-650–35)
fabric (8-harness satin, 3K tow size, with UC-309 sizing) in HFPE-II polyimide
thermosetting resin. It was demonstrated that the viscoelastic response of the
virgin material was linear within a range of stress and temperature, beyond which
it became nonlinear. The microcracking within the fiber bundles further enhanced
the nonlinearity of the viscoelastic response. Since the tests measured strain
response under prescribed stress, the CDM formulation was appropriately modi-
fied to consider stress as an independent variable. Thus the Gibbs free energy
function G was formulated such that the strain response can be written as

ðaÞ
@G skl ; Dmn ; T; gs
eij ¼  ; ð5:112Þ
@skl
where the damage mode tensors D(a) ij are as defined by Eq. (5.20) and gs are
the viscoelastic internal state variables. Note that the damage mode tensor is kept
time independent by taking its initial value at time t ¼ 0, analogous to the linear
viscoelastic case in the CDM approach by Kumar and Talreja [38] treated above.
Using polynomial expansion in terms of the integrity bases of the function G,
similar to the procedures for the linear elastic and viscoelastic cases above, and by
incorporating formulation of polymer viscosities available in the literature, the
time-dependent in-plane strain response of an orthotropic composite with trans-
verse cracks is derived as [42]

ep ¼ CEpq þ CD g E
pq D þ Cpq g  Cpq Dg sq ; ð5:113Þ

where p, q ¼ 1, 2, and 6, and superscripts E, D and g to the compliance matrix C


denote the elastic, damage, and viscous contributions, respectively, and the last
term in the parenthesis stands for the interactive contribution between damage
176 Macro-damage mechanics

and viscosity. The damage variable D ¼ D11, which is the only nonzero component
of the damage mode tensor for transverse cracks.
The material constants in the four matrices in Eq. (5.113) must be evaluated
before these relationships can be used to predict the nonlinear viscoelastic
response. The first matrix is simply the elastic response matrix and can be found
by recording the instantaneous strain response to imposed stresses. The second
matrix can be found experimentally in conditions where the viscoelasticity is
negligible, and the procedure could be one of those discussed above for the case
of damage in elastic composites. In the absence of damage, the third matrix in
Eq. (5.113) represents the time-dependent response. Finally, the fourth matrix
needs determining to see how damage enhances the viscoelastic deformation.
A procedure for the complete evaluation of the material constants involved is
thus far from trivial. An experimental procedure aided by finite element modeling
was developed in Ahci and Talreja [42]. The resulting characterization of non-
linear viscoelasticity and damage was used to illustrate the effects of temperature
and stress levels as well as the nature of the damage–viscoelasticity coupling.

5.5 Summary

This chapter has presented the basic concepts of continuum damage mechanics and
illustrated its application to damage in composite materials. The classical field of
continuum thermodynamics combined with internal variables characterization of
damage provides a powerful tool to describe materials response affected by damage.
In its conventional form the CDM relies on material parameter identification by
experimental data. This has been a burden that has been increasingly challenging to
carry because of the multiplicity of damage modes and the ensuing complexity of
response changes. The extension of CDM developed by judiciously aiding it with
selected micromechanics solutions (analytical or numerical) has eased the burden
and thereby increased the attractiveness of CDM. This integrative approach of
combining CDM and micromechanics has been named synergistic damage
mechanics. This chapter has illustrated the application of SDM to multiple (up to
three) damage modes in laminates. Results of parametric studies have also been
presented to offer further insight into the consequences of multiple damage modes.
Although most applications of polymer-based composites have been in struc-
tures operating at temperatures safely below the glass transition temperature of
the polymer, there are instances such as in jet engine casings where high tempera-
tures are induced. Viscoelastic response of the composite in such cases takes place
and must be addressed by itself and when damage occurs. The CDM and SDM
methodologies have been presented for linear viscoelastic composites with
damage. At high temperatures and with extensive damage the viscoelastic
response can become nonlinear. This case has added complexity in identification
of material constants. A methodology that uses experimental data as well as
numerical computations has been discussed.
References 177

References

1. L. M. Kachanov, On the creep rupture time. Izv Akad Nauk SSR, Otd Tekhn Nauk, 8
(1958), 26–31.
2. L. M. Kachanov, Rupture time under creep conditions. Int J Fract, 97:1 (1999), 11–18.
3. Y. N. Robotnov, Creep Problems in Structural Members. (Amsterdam: North-Holland, 1969).
4. S. Murakami and N. Ohno, A continuum theory of creep and creep damage. In Creep
in Structures, 3rd IUTAM Symposium, ed. A. R. S. Ponter and D. R. Hayhurst. (Berlin,
Germany: Springer-Verlag, 1981), pp. 422–44.
5. J. Lemaitre and J. L. Chaboche, Damage mechanics. Chapter 7 in Mechanics of Solid
Materials. (Cambridge: Cambridge University Press, 1990), pp. 346–450.
6. J. Lemaitre, A continuous damage mechanics model of ductile fracture. J Eng Mater
Technol, Trans ASME, 107:42 (1985), 83–9.
7. J. Lemaitre and J. L. Chaboche, Aspect phenomenologique de la rupture par endomm-
agement. Journal de Mecanique Appiquee, 2:3 (1978), 317–65.
8. J. Lemaitre and R. Desmorat, Engineering Damage Mechanics: Ductile, Creep, Fatigue
and Brittle Failures. (Berlin: Springer-Verlag, 2005).
9. J. L. Chaboche, Anisotropic creep damage in the framework of continuum damage
mechanics. Nucl Eng Des, 79 (1984), 309–19.
10. J. L. Chaboche, Continuum damage mechanics: part I – general concepts. J Appl Mech,
Trans ASME, 55:1 (1988), 59–64.
11. J. L. Chaboche, Continuum damage mechanics: part II – damage growth, crack initi-
ation, and crack growth. J Appl Mech, Trans ASME, 55:1 (1988), 65–72.
12. J. W. Ju, Isotropic and anisotropic damage variables in continuum damage mechanics.
J Eng Mech, 116 (1990), 2764–70.
13. R. Talreja, A continuum-mechanics characterization of damage in composite materials.
Proc R Soc London A, 399:1817 (1985), 195–216.
14. R. Talreja, Internal variable damage mechanics of composite materials. In Yielding,
Damage, and Failure of Anisotropic Solids, ed. J. P. Boehler. (London: Mechanical
Engineering Publications, 1990), pp. 509–33.
15. R. Talreja, Damage mechanics of composite materials based on thermodynamics with
internal variables. In Polymer Based Composite Systems for Structural Applications, ed.
A. Cardon and G. Verchery. (London: Elsevier, 1991), pp. 65–79.
16. R. Talreja, Continuum modeling of damage in ceramic matrix composites. Mech Mater,
12:2 (1991), 165–80.
17. R. Talreja, Damage characterization by internal variables. In Damage Mechanics of
Composite Materials, ed. R. Talreja. (Amsterdam: Elsevier, 1994), pp. 53–78.
18. S. Nemat-Nasser and M. Hori, Micromechanics: Overall Properties of Heterogeneous
Materials, 2nd edn. (Amsterdam: North Holland, 1999).
19. A. A. Vakulenko and M. Kachanov, Continual theory of a medium with cracks. Izv AN
SSSR, Mekhanika Tverdogo Tela (Mech Solids), 6:4 (1971), 159.
20. J. Varna, Physical interpretation of parameters in synergistic continuum damage
mechanics model for laminates. Compos Sci Technol, 68:13 (2008), 2592–600.
21. A. J. M. Spencer, Theory of invariants. In Continuum Physics, ed. C. A. Eringen.
(New York: Academic Press, 1971), pp. 239–353.
22. J. Adkins, Symmetry relations for orthotropic and transversely isotropic materials.
Arch Ration Mech Anal, 4:1 (1960), 193–213.
178 Macro-damage mechanics

23. R. Talreja, Transverse cracking and stiffness reduction in composite laminates.


J Compos Mater, 19:4 (1985), 355–75.
24. A. L. Highsmith and K. L. Reifsnider, Stiffness-reduction mechanisms in composite
laminates. In Damage in Composite Materials, ASTM STP 775, ed. K. L. Reifsnider.
(Philadelphia, PA: ASTM, 1982), pp. 103–17.
25. J. Varna, N. V. Akshantala, and R. Talreja, Crack opening displacement and the associ-
ated response of laminates with varying constraints. Int J Damage Mech, 8:2 (1999), 174–93.
26. J. Varna, R. Joffe, and R. Talreja, A synergistic damage-mechanics analysis of trans-
verse cracking in [ y/904]s laminates. Compos Sci Technol, 61:5 (2001), 657–65.
27. J. Varna, R. Joffe, and R. Talreja, Mixed micromechanics and continuum damage
mechanics approach to transverse cracking in [S,90(n)](s) laminates. Mech Compos
Mater, 37:2 (2001), 115–26.
28. C. V. Singh and R. Talreja, Damage mechanics of composite laminates with transverse
matrix cracks in multiple orientations. In 48th AIAA SDM Conference, Honolulu,
Hawaii, USA. (Reston, VA: AIAA, 2007).
29. C. V. Singh and R. Talreja, Analysis of multiple off-axis ply cracks in composite
laminates. Int J Solids Struct, 45:16 (2008), 4574–89.
30. C. V. Singh and R. Talreja, A synergistic damage mechanics approach for composite
laminates with matrix cracks in multiple orientations. Mech Mater, 41:8 (2009), 954–68.
31. S. Li, C. V. Singh, and R. Talreja, A representative volume element based on transla-
tional symmetries for FE analysis of cracked laminates with two arrays of cracks. Int
J Solids Struct, 46:7–8 (2009), 1793–804.
32. C. V. Singh, Multiscale modeling of damage in multidirectional composite laminates.
Ph.D. thesis, Texas A&M University, College Station, TX (2008).
33. J. Varna, R. Joffe, N. V. Akshantala, and R. Talreja, Damage in composite laminates
with off-axis plies. Compos Sci Technol, 59:14 (1999), 2139–47.
34. J. Varna, L. Berglund, R. Talreja, and A. Jakovics, A study of crack opening displace-
ment of transverse cracks in cross-ply laminates. Int J Damage Mech, 2:3 (1993), 272–89.
35. J. Tong, F. J. Guild, S. L. Ogin, and P. A. Smith, On matrix crack growth in quasi-isotropic
laminates – I. Experimental investigation. Compos Sci Technol, 57:11 (1997), 1527–35.
36. Schapery, R. A. and Sicking, D. L., On nonlinear constitutive equations for elastic
and viscoelastic composites with growing damage. In Mechanical Behavior of
Materials, ed. A. Bakker. (Delft, The Netherlands: Delft University Press, 1995), pp. 45–76.
37. Shapery, R. A., On viscoelastic deformation and failure behavior of composite materials
with distributed flaws. In Advances in Aerospace Structures and Materials, ASME-AD-01,
ed. S. S. Wang and W. J. Renton. (Philadelphia, PA: ASTM, 1981), pp. 5–20.
38. R. S. Kumar and R. Talreja, A continuum damage model for linear viscoelastic
composite materials. Mech Mater, 35:3–6 (2003), 463–80.
39. R. M. Christensen, Theory of Viscoelasticity: An Introduction, 2nd edn. (New York:
Academic Press, 1982).
40. R. S. Kumar and R. Talreja, Linear viscoelastic behavior of matrix cracked cross-ply
laminates. Mech Mater, 33:3 (2001), 139–54.
41. J. Varna, A. I. Krasnikovs, R. S. Kumar, and R. Talreja, A synergistic damage mech-
anics approach to viscoelastic response of cracked cross-ply laminates. Int J Damage
Mech, 13:4 (2004), 301–34.
42. E. Ahci and R. Talreja, Characterization of viscoelasticity and damage in high tem-
perature polymer matrix composites. Compos Sci Technol, 66:14 (2006), 2506–19.
6 Damage progression

6.1 Introduction

As discussed in Chapter 1, material constitutive relationships are needed in order


to conduct failure analysis of a structure subjected to service loading. Contrary to
homogeneous materials such as metals whose stress–strain relationships can be
specified a priori, composite materials suffer damage that can alter these relation-
ships. Thus if strains are prescribed, the stress response can be expressed as
sij ¼ Cijkl ðekl Þekl ; ð6:1Þ
where the stiffness matrix Cijkl changes with strain as the composite material
suffers damage. Determination of Cijkl as a function of the applied strain can be
achieved by solving two related sub-problems:
1. Describe stiffness changes as a function of damage: In this step Cijkl is expressed
in terms of some damage characteristic, such as the ply crack density,

Cijkl ¼ Cijkl rðaÞ ; ð6:2Þ


where r(a) is the crack density of damage mode a = 1, 2,. . ., N. As discussed in
Chapter 4, micromechanics solutions for multiple damage modes are not
always possible. Using the CDM formulation for multiple damage modes in
linearized form we can express Eq. (6.2) as
X
N

ðaÞ
Cijkl ¼ C0ijkl þ Cijkl rðaÞ ; ð6:3Þ
a¼1
ðaÞ
where C0ijkl
is the stiffness tensor for undamaged laminate, and Cijkl represents
the stiffness changes brought about by damage mode a.
2. Describe the evolution of crack density as a function of applied loading:

rðaÞ ¼ rðaÞ ðekl Þ: ð6:4Þ

Combining the solution to the two sub-problems above, we obtain


X
N

Cijkl ¼ C0ijkl þ CDAM


ijkl rðaÞ ðemn Þ : ð6:5Þ
a¼1
180 Damage progression

Cracks/In.
40
[0/90n /0],T300/934

30 n=4
n=3
n=2
n=1
20

10

0 20 40 60 80 100 120 Ksi

Figure 6.1. Transverse crack density vs. applied axial stress in [0m/90n]s laminates [1].
Reprinted, with kind permission, from Composites Technology Review, copyright ASTM
International, 100 Barr Harbor Drive, West Conshohocken, PA 19428.

The solution to the first sub-problem has been described in Chapters 4 and 5. This
chapter concerns the second sub-problem.
To take advantage of directional properties of plies, composite laminates are
made from a mix of longitudinal, transverse, and angle plies. Since the unidirec-
tional lamina has low strength in the transverse direction, it is prone to cracking
along fibers. When the applied load is increased beyond the strain (or stress) at
which initiation of cracking occurs, new cracks form in the cracked ply in between
existing cracks. Initially these cracks are far apart and do not interact with each
other. However, quickly they form a roughly periodic array of parallel cracks.
Figure 6.1 shows the increase in density of transverse cracks (i.e., number of
cracks per unit length normal to the crack plane) for typical configurations of
cross-ply laminates when loaded under axial tension. Prediction of such curves has
been an extensive subject of study. The approaches used to model crack initiation
and progression (multiplication) can be divided into two categories: strength-
based models and energy-based models. As the names suggest, the first category
of models involves use of strength (failure) criteria, while energy balance concepts
underlie the second category.

6.2 Experimental techniques

Before discussing experimental observations and measurements in Section 6.3


below, a brief description of each of the techniques is given next. This is to help
provide a background against which to interpret the observations. No attempt is
made here to go into any depth on these techniques, as that would distract from
the focus of this chapter. For further details references [2–4] are suggested.
Over the past forty years many nondestructive evaluation (NDE) techniques
have been developed to detect, monitor, and observe ply cracking damage in
6.2 Experimental techniques 181

composite laminates. The quantities targeted for observations and measurements


concerning ply cracking include the following:
1. Crack initiation strains (or stresses).
2. Increase in number of cracks with applied loading or with number of cycles in
case of fatigue experiments.
3. Changes in stiffness properties with damage.
4. Crack opening displacements (COD) and crack profiles.
5. Final failure strains and damage leading to the final failure.
Brief overviews of the main techniques are provided here.

Edge replication
Although direct optical microscopy can be used [5] for surface observation to evaluate
damage, it requires in most cases specimens to be removed at regular intervals during
loading. If feasible, an optical microscope can be mounted on the testing machine itself,
but it often requires the microscope to be brought close to the specimen surface for
enough resolution. This limitation can be overcome by using edge replication instead.
This is a simple and easy technique if the purpose is to monitor damage (count cracks)
in composites, as was shown early by Stalnaker, Stinchcomb, and Masters [6, 7]. The
approach involves the microscopic examination of surface replicas, which are prepared
by pressing softened rubber tape (or tape with an adhesive, e.g., acetate) against the
specimen edge. With this procedure it is possible to quickly obtain permanent records
of the specimen edge while the specimen itself is held in loading. The popularity of the
technique is evidenced by its utilization by many researchers in the field [2, 8–14]. Two
examples of photomicrographs from edge replicas are shown in Figure 6.2.

Acoustic emission
In this technique, a sensor is used to monitor acoustic emission (AE) signals from
stress waves generated due to some local failure in the material, such as formation of
a ply crack. A major limitation of the approach is its inability to distinguish between
damage types and to provide information concerning crack location and orientation
[14]. However, if the damage mechanism is known based on prior experience with the
material system, this technique can indicate the applied load at which the damage
initiated. Using this to monitor damage and to take edge replicas accordingly has
been found useful. A combination of ultrasonic polar backscattering scans and AE,
which yields accurate crack location, has also been suggested [15, 16]. Figure 6.3
shows the variation of cumulative count of acoustic events and the corresponding
stress–strain response of a ceramic composite subjected to tensile loading.

X-ray radiography
X-ray radiography is a useful technique in finding internal cracks that are not visible
by optical microscopy. For clarity of observation a penetrant fluid that absorbs X-rays
is often necessary to provide sufficient contrast between exposed and unexposed
regions. Pictures of the developed X-ray films can then be enlarged for further clarity.
182 Damage progression

(a) (b)

Transverse Cracking

Transverse Cracking
Longitudinal Cracking

Figure 6.2. Photomicrographs of edge replicas showing details of damage development in


fatigue loading of quasi-isotropic laminates: (a) [0/ 45/90]s and (b) [0/90/ 45]s laminate
[2]. Reprinted, with kind permission, from Damage in Composite Materials, copyright
ASTM International, 100 Barr Harbor Drive, West Conshohocken, PA 19428.

(a) (b)
1200 100

80
900
Cumulative AE counts

Stress (MPa)

60
600
40

300
20

0 0
0 20 40 60 80 100 0 0.05 0.1 0.15 0.2
Stress (MPa) Strain (%)

Figure 6.3. Variation of cumulative acoustic event (AE) counts (a), and stress–strain curve
(b) for a ceramic composite loaded in tension [14]. Reprinted, with kind permission, from
Damage Detection in Composite Materials, copyright ASTM International, 100 Barr Harbor
Drive, West Conshohocken, PA 19428.
6.2 Experimental techniques 183

Figure 6.4. An X-ray radiograph of damage in a quasi-isotropic laminate [19]. Reprinted,


with permission, from Effects of Defects in Composite Materials, copyright ASTM
International, 100 Barr Harbor Drive, West Conshohocken, PA 19428.

The pictures, however, only provide 2-D images of cracks and therefore cannot be
used to find the depth information needed to separate cracks of one layer from
another. For that purpose penetrant enhanced X-ray stereo radiography can be
used. The standard procedure is to make two X-ray radiographs of the object at
slightly different orientations, which is usually done by rotating the X-ray source
relative to the object. The spatial locations of cracks can be observed with a
microscope but cannot be recorded. The approach has been successfully used to
detect fiber fractures, delaminations and matrix cracks [17–19]. To enable imaging
and finer precision detection of internal defects, the technique employed is X-ray
tomography, which uses a medical scanner and yields a three-dimensional image of
an object [20]. An X-ray radiograph of damage in a quasi-isotropic laminate is
shown in Figure 6.4.

Ultrasonic C-scan
Ultrasonic C-scan is a nondestructive inspection technique in which a short pulse
of ultrasonic energy is made incident on a sample. The attenuation of the pulse is
influenced by voids, delaminations, state of resin cure, the fiber volume fraction,
184 Damage progression

the condition of the fiber/matrix interface, and any foreign inclusions present. The
technique is, however, limited by its inability to detect very small defects or cracks
with their planes parallel to the wave direction [14].
In some cases a damaged laminate material may contain a large number of
small and partially grown internal cracks. On increase in loading, these cracks can
grow further and coalesce to form larger cracks. The common NDE techniques
described above cannot provide accurate information on the formation, number,
size, and progression of these internal cracks. For such cases a through-transmis-
sion ultrasonic C-scan imaging with inclined focusing transducers in confocal
configuration has been recently suggested [21, 22]. Other methods based on
vibrations and lamb waves have also been suggested [23, 24].

Technique for COD measurement


In addition to the observations of damage certain measurements also need to be
made as dictated by the models. For instance, the COD of ply cracks is a quantity
that enters the SDM approach described in Chapter 5. Since there are no standard
methods for measuring this quantity, Varna et al. [26–27] developed a set-up for
this purpose. To observe an individual ply crack, the set-up uses a miniature
materials tester (MINIMAT) for loading a thin strip cut out from the cracked
laminate. The open crack is observed under an optical microscope equipped with a
video camera. The video signal transmitted to a TV monitor displays the crack
profile at sufficient magnification ( 2103) to measure the COD. Figure 6.5
presents COD as a function of position along the cracked 90 -ply thickness
(z-direction) and compares it with the theoretical shape prediction obtained using
Linear Elastic Fracture Mechanics (LEFM).

Raman spectroscopy
Recently Katerelos and coworkers [28] have developed an experimental technique
based on Raman spectroscopy, which uses the property that the Raman
vibrational wave numbers (frequencies) of certain chemical groups of commercial
reinforcing fibers, such as aramid or carbon, are stress and strain dependent [16].
Thus, the wave number shift along an embedded fiber can be utilized to determine
stress or strain. The technique can provide a high spatial resolution of  1mm.
However, for the approach to work the matrix needs to be translucent. Also, the
data acquisition can take longer than milliseconds for a single measurement.
Moreover, certain amorphous fibers such as glass have a weak Raman response.
Aramid fibers, on the other hand, scatter the waves well. Therefore, a small
amount of aramid fiber placed in strategic positions within a glass-fiber laminate
can act as Raman sensors of stress and strain. The micromechanical strain
mapping results are then used to derive the properties, i.e., the longitudinal
modulus of elasticity and the magnitude of the residual strains caused by cracking
[29]. The approach has been successfully applied to evaluate ply cracking damage
evolution and resulting stiffness changes [30–32].
6.3 Experimental observations 185

25
[O2/908]s
Hybrid CF/GF
Crack opening displacement u (mm)

20
(Different sections)
LEFM theory

15

10

0
0.0 0.2 0.4 0.6 0.8 1.0
z—t90
Relative position in 90-layer
2t90

Figure 6.5. Crack opening displacement measured using MINIMAT along the 90 -layer
thickness of a hybrid CF/GF cross-ply laminate at various sections through the specimen
width. The solid line depicts predictions from LEFM. Source: J. Varna, L. Berglund,
R. Talreja, and A. Jakovics, A Study of crack opening displacement of transverse cracks in
cross-ply laminates, Int J Damage Mech, Vol. 2, pp. 272–89 (1993).

6.3 Experimental observations

Experimental studies on initiation and growth of intralaminar cracking in composite


laminates have been extensive. Most of the work has focused on 90 -ply cracking in
cross-ply and quasi-isotropic laminates. Chapter 3 reviewed some of those observa-
tions to illustrate the nature of composite damage. A book chapter by Nairn [33] is
also recommended for a good overview of the topic. In the following we review
quantitative data related to the initiation and progression of ply cracking in laminates.

6.3.1 Initiation of ply cracking


The applied loading (stress or strain) at which ply cracking in laminates first
occurs is of interest from materials selection as well as design points of view.
A good laminate configuration (orientation, thickness, and sequence of plies) will
delay initiation of ply cracking to as high a load as possible. Experimental
observations have indicated that all laminate configuration parameters influence
ply crack initiation. Most early studies examined ply cracking in cross-ply lamin-
ates of glass-polyester or glass/epoxy under axial tension. The ply cracking in the
90 -plies could in most cases be observed by looking at the specimen surface under
an optical microscope at low magnification. The near-transparency of composites
186 Damage progression

(a) (b)
3
Longitudinal cracking strain (%) 1.2

Initial cracking strain (%)


Unnotched
2.5 1 Notched

2 0.8
1.5 0.6

1 0.4
0.2
0.5
0
0 40 60 80 100
40 50 60 70 80 90 100
Off-axis angle (␪)
Angle (degrees)
notched unnotched polished unnotched

GFRP [0/␪/0] laminates CFRP (IM7/8552) [02/␪4]s laminates

Figure 6.6. Crack initiation strains for angle ply laminates as a function of ply orientation:
(a) GFRP laminates; (b) CFRP laminates. Part (a) reprinted, with kind permission, from
Composites A, Vol. 28, L.E. Crocker, S.L. Ogin, P.A. Smith and P.S. Hill, Intra-laminar fracture
in angle-ply laminates, pp. 839–46, copyright Elsevier (1997). Part (b) reprinted, with kind
permission, from Springer Science+Business Media: J Materials Science, Intra-laminar cracking
in CFRP laminates: observations and modeling, Vol. 41, 2006, pp. 6599–609, N. Balhi et al.

of glass fibers made this possible. In carbon fiber composites, however, one had to
rely on edge observations until X-ray radiography allowed imaging of the interior
cracks. The axial strain values at initiation of transverse cracking in cross-ply
laminates of glass/epoxy and carbon/epoxy fall typically in the range 0.4–1.0%.
As the off-axis ply angles decrease from 90 the initiation strain increases. For
instance, for 45 -plies cracking barely initiates at 1.0% axial strain [34]. Figure 6.6
illustrates the effect of off-axis angle on crack initiation strain for (a) GFRP [35],
and (b) CFRP [36]. Data for notched samples are also indicated to illustrate the
effect of local stress enhancement. In Figure 6.6 the data for un-notched samples
with and without polished edges illustrate the effect of machining-induced flaws in
initiating cracking. Also of note is the range of strain and the lowest strain for
GFRP versus CFRP. In the latter case, the lowest strain to crack initiation is higher
due to the higher elastic modulus of carbon fibers. In both cases the initiation strain
at low off-axis angles is limited by the fiber failure strain.
The thickness of the cracking ply also has a large effect on einit. Garret and
others [37–44] performed a systematic study of thickness effect on cracking in [0m/90n]s
laminates. They used glass-reinforced polyester [37, 38] and glass-reinforced epoxy
[39–41] as the laminate materials and varied the thickness of the 90 -plies while keeping
the thickness of the 0 -ply constant at 0.5 mm. Figure 6.7 shows the variation of einit
with respect to the total thickness of the 90 -plies. As the thickness of the 90 -plies
increases, einit decreases. When the thickness of the 90 -plies is more than 0.4 mm, the
cracks initiate instantly and span the entire cross section of the 90 -plies. At the other
extreme, if the thickness of the 90 -plies is less than 0.1 mm, cracks may be entirely
suppressed and the laminate eventually may fail due to other damage mechanisms,
e.g., delamination and fiber breakage. The experiments on carbon/epoxy laminates
6.3 Experimental observations 187

Strain to First Microcrack (%)

0
0 1 2 3 4 5
Total Thickness of 90° Plies (mm)

Figure 6.7. The strain to initiate ply cracking in [0m/90n]s glass/epoxy laminates as a function
of the total thickness of 90 -plies [33]. Experimental data from [38]. Reprinted, with
kind permission, from Polymer Matrix Composites, J. A. Nairn, Matrix microcracking in
composites, pp. 403–32, copyright Elsevier (2000).

show similar behavior [41, 45]. In general, for [0m/90n]s laminates, complete crack
suppression is expected only when m / n  10 although the suppression effects may be
felt in the range 2 m / n < 10 [46]. This difference in initiation strain for different
thicknesses is attributed to the relative constraint of the supporting 0 -plies onto the
crack opening in the 90 -plies [47].
The actual ply layup in the laminate may also play a role in determining einit.
For instance, [902/0]s laminates develop cracks sooner than [0/902]s laminates [48].
It is because the 90 -plies in [90/0]s laminates are on the outer surface and may not
experience much support for crack suppression from the inner 0 -plies. For
multidirectional laminates, the situation is even more complex and einit will depend
on the thickness ratio and the stiffness properties of the cracking and supporting
plies. The laminate preparation and processing method also may affect crack
initiation; filament-wound laminates may be more prone to cracking than those
made from prepregs using an autoclave.

6.3.2 Crack growth and multiplication


The growth of ply cracks is usually unstable in composites such as GFRP and
CFRP. The initial transverse cracks are found to grow through the lamina
thickness quickly but are usually arrested at interfaces with adjacent plies of
different orientation, e.g., at the 90/0 interface for cross-ply laminates [38, 40]
and at the 90/45 interface in [0/90/45/+45]s laminates [34]. Continued loading
usually leads to formation of progressively more cracks between the already
formed cracks. Most experimental studies point out that once the ply cracks have
grown through the lamina thickness, they often grow unstably along the fiber
direction through the laminate width, and are thus described as “tunneling
cracks.” In some cases, however, cracks in plies other than 90 may not grow
fully before the laminate fails by delamination [34, 49–51]. Such partial cracks can
be observed in the 45 -plies of quasi-isotropic [0/90/45/+45]s laminates [34].
188 Damage progression

Figure 6.8. Transverse cracking in GFRP specimens with transverse-ply thickness of (a) 0.75
mm, (b) 1.5 mm, and (c) 2.6 mm, strained to 1.6%. Reprinted, with kind permission,
from Springer Science+Business Media: J Materials Science, Multiple transverse fracture in
90 degrees cross-ply laminates of a glass fiber-reinforced polyester, Vol. 12, pp. 157–68,
K. W. Garrett and J. E. Bailey.

Once the ply cracking has initiated, more and more ply cracks start appearing in
between existing cracks, and the crack density rises quickly. As the crack spacing
between adjacent cracks decreases, the cracks start interacting. Closely interacting
cracks typically provide a “shielding effect,” which tends to reduce the stresses
between two adjacent cracks. Therefore, on further loading, the rate of cracking
reduces and finally approaches a saturation value. Thus a typical damage growth
curve consists of three stages: crack initiation, rapid rise in crack density by
multiplication, and reducing rate of crack density evolution until saturation (see
Figure 6.11). Reifsnider and associates [10, 52] described this microcrack satur-
ation as a material state and called it the characteristic damage state (CDS). They
proposed that CDS is a well-defined laminate property and does not depend on
load history, environment, or thermal or moisture stresses. However, later investi-
gations by Akshantala and Talreja [53] have suggested that CDS may not be a
single state under fatigue loading but may depend on the maximum stress applied.
The progression from initial rapid rise until saturation of crack density depends
on the ply material, the ply stacking sequence, and also the laminate fabrication
process. For example, well-made carbon/epoxy laminates typically have a rapid
rise in ply crack density. The saturation crack density is often found to inversely
scale with cracking ply thickness, with thin plies accumulating a large number
of cracks/mm [54, 55]. This is well illustrated in Figure 6.8 where experiments
conducted on GFRP specimens with different ply thicknesses [54] showed that thin
90 -plies can accumulate numerous cracks. The evolution of average crack density
with respect to applied load for the same specimens is shown in Figure 6.9. It can be
observed here that the saturation crack density is roughly proportional to 1/t90. The
damage evolution curves in Figure 6.10 also show that thicker cross-ply laminates
typically show lower crack density at saturation. The damage evolution in outer
and inner 90 -plies is also compared. The laminates with outer 90 -plies are
6.3 Experimental observations 189

1.2

2t90
1
Average crack density (1/mm)

3.2mm
2mm
0.8
1.5mm

0.6 0.75mm

0.4

0.2

0
0 50 100 150 200 250
Applied stress (MPa)

Figure 6.9. Average crack density as a function of applied stress for GFRP cross-ply laminates
with different transverse ply thicknesses (2t90). The experimental data are from [54].

observed to have lower saturation crack densities. A typical damage evolution


curve for ply cracking based on these studies is shown in Figure 6.11.

6.3.3 Crack shapes


When cracks are widely spaced, the maximum principal stress occurs on the plane
midway between existing cracks. Thus, at low crack density there is a tendency for
new cracks to form midway and develop into a periodic array. However, at large
crack density cracks interact causing the maximum principal stress to shift towards
the 0/90 interface close to an existing crack. This may result in curved or oblique
microcracks forming near the 0/90 interface [54, 56, 57]. These cracks make an angle
of 40–60 with existing straight cracks. Lundmark and Varna [58] have recently
reported that curved microcracks form more readily in the low-temperature regime
than at room temperature. In fact, at low temperatures complex crack trajectories
are observed to form (see Figure 6.12). This may result in a highly damaged region
in the laminate encompassing multiple crack types (Figure 6.13).

6.3.4 Effect of cracking


The most direct effect of ply cracking is the reduction of the thermomechanical
properties of the laminate, including changes in the effective values of Young’s
moduli, Poisson’s ratios, and thermal expansion coefficients. The changes in
stiffness properties can in turn lead to change in the behavior of the whole
structure, e.g., the magnitude of its deflection and vibrational frequencies, some-
times making the structure unable to carry out its intended design function. Even
if it does not cause the structure to fail, substantial ply cracking may give rise to
190 Damage progression

(a)
1.4

1.2
[0/902]s
Crack density (1/mm)

1.0

0.8

0.6 [0/904]s
[0/90]s
0.4

0.2

0.0
0 200 400 600 800
Stress (MPa)

(b)
1.0

[90/0]s
[90/0/90]T
0.8
Crack density (1/mm)

[90/02]s
0.6

0.4
[90/04]s
0.2

0.0
0 100 200 300 400 500 600 700 800 900 1000
Stress (MPa)

Figure 6.10. Damage evolution curves for (a) [0/90m]s and (b) [90/0m]s laminates. Reprinted,
with kind permission, from Polymer Matrix Composites, J. A. Nairn, Matrix microcracking
in composites, pp. 403–32, copyright Elsevier (2000).
Crack density (no. of cracks/mm)

Stage I Stage II Stage III

Crack initiation & Multiple crack Saturation of


propagation formation progressive cracking
through
laminate width

Applied load (strain or stress)

Figure 6.11. A typical damage evolution curve for ply cracking in laminates.
6.3 Experimental observations 191

0° layer t0
90° layer
Crack
a Crack type B
Normal t90
type A
crack Partial
a crack

Crack
type C Crack
type D

w w

Figure 6.12. Different crack types observed during tensile testing of a [02/904]s CF/EP
laminate at cryogenic temperature (150 C). Reprinted, with kind permission, from Eng
Fract Mech, Vol. 75, P. Lundmark and J.Varna, Damage evolution and characterization of
crack types in CF/EP laminates loaded at low tempratures, pp. 2631–41, copyright (2008),
with permission from Elsevier.

Figure 6.13. A snapshot of a highly damaged region of a [02/904]s CF/EP laminate at an


applied stress of 343 MPa (0.66% strain) during tensile loading at cryogenic temperature
(150 C). Reprinted, with kind permission, from Eng Fract Mech, Vol. 75, P. Lundmark and
J.Varna, Damage evolution and characterization of crack types in CF/EP laminates loaded
at low tempratures, pp. 2631–41, copyright (2008), with permission from Elsevier.

more deleterious forms of damage such as delamination and longitudinal splits, or


provide pathways for the entry of moisture and corrosive liquids.

6.3.5 Loading and environmental effects


Most experiments are performed using uniaxial tension, but ply cracks will also
form under other loading conditions, such as fatigue, biaxial, or shear loading.
Biaxial loading of [0m/90n]s laminates may show cracks in both 0 - and 90 -plies. If
the material response remains linearly elastic after cracking, then neglecting crack
interaction effects between plies the effect of biaxial loading of [0m/90n]s laminates
may be seen as equivalent to two uniaxial loading cases, one on [0m/90n]s and the
other on [90m/0n]s laminates. On thermal loading, the differential shrinkage
between the 0 - and 90 -plies may also induce biaxial loading [59–65]. In general,
192 Damage progression

Figure 6.14. X-radiograph of thermal stress-induced matrix microcracks in a [+454/–454/


908]s laminate. The 90 -ply matrix cracks run from left to right, and trigger the formation of
many short, stitch-like –45 -ply matrix cracks. Long +45 -ply matrix cracks appear.
Reprinted, with kind permission, from J. A. Lavoie and E. Adolfsson, J Compos Mater,
Vol 35, pp. 2077–97, copyright # 2001 by Sage Publications.

thermal effects can be accounted for in the analysis and predictions of classical
laminate theory. Bailey et al. [41] studied the effect of thermal stresses and Poisson’s
contraction on ply cracking in CFRP and GPRP cross-ply laminates. Thermal
residual stresses typically lower crack initiation strains. Thermal effects are larger
in CFRPs than in GFRPs due to larger differences in both thermal expansion
coefficients and Young’s moduli in directions parallel and perpendicular to fibers
[41]. For instance, a [0/90]s laminate with a ply thickness of 0.5mm shows 0.322%
thermal strains for CFRPs and 0.094% for GFRPs. The Poisson’s effect was found
to be greater in GFRPs because of their larger failure strains. Poisson’s strains can
sometimes be so high that they can induce transverse cracking of the 0 -plies.
The mismatch between the thermal expansions of different plies can also result
in a high-density form of microcracking known as “stitch cracking.” Lavoie and
Adolfsson [66] studied this type of cracking in [+yn/–yn/902n]s laminates (see
Figure 6.14). Stitch cracks appear to form instead of interply delamination at
the tip of a 90 crack in the adjacent constraining ply when the included angle
between the two is greater than 50 . Stitch cracking is also observed in case of
fatigue loading [19]. Variation in intralaminar fracture toughness over tempera-
ture can also change the initiation and evolution of ply cracking [67].
The residual stresses due to moisture can also affect the cracking process
[63, 68]. This could be due to degradation of the matrix during hygrothermal
aging [69, 70]. A combination of moisture and thermal residual stresses can cause
significant stiffness reduction, especially at high crack density [71].
6.3 Experimental observations 193

If laminates are loaded in bending, ply cracks will form on the tension side
[11, 72, 73], and for such loading analyses need to account for the resulting local
stress states [11, 73–77].

6.3.6 Cracking in multidirectional laminates


Cross-ply laminates are not very common in practical applications. In fact,
efficient use of composite laminates in a wide range of applications requires
placing plies in multiple orientations. A common example is a class of laminates
known as quasi-isotropic that can be constructed in different ways; a common
form usually has a mix of plies with 0 , 45 , and 90 orientations, exemplified by
the [0/90/ 45]s configuration. In the off-axis plies of such laminates, cracks
usually initiate at much higher applied axial strains than in 90 -plies and may
sometimes show curved patterns due to inclined principal stress trajectories.
Curved cracks and cracks near free edges typically promote delamination at the
ply interface [78]. Experimental data show that the majority of off-axis cracks
form at the edges of the test coupons and often may not grow across the thickness
and width before the specimen fails by extensive delamination.
Furthermore, in some cases, other damage modes such as delamination or fiber
fracture can occur even before any cracking in the off-axis plies ensues. Johnson
and Chang [79, 80] carried out extensive experiments on a variety of multidirec-
tional laminate configurations and found that for laminates having a ply angle
greater than 45 , ply cracking is a dominant damage mode, with possible delami-
nation at free edges. For instance, in [0/y/0] laminates with y  45 the dominant
damage was ply cracking [31, 35], while in [90/30/–30]s a combination of edge
delamination and fiber fracture led to final failure. Crocker et al. [35] also found
that in [0/45/0] laminates, delamination ensued as soon as ply cracks initiated in
45 -plies.
In case of multidirectional laminates containing 90 -plies, an off-axis ply
adjacent to a 90 -ply shows numerous partial cracks, which may or may not
join to form through cracks on increase of loading. The cracks in the contiguous
ply almost always start from its interface with the 90 -ply. Experiments by
Yokozeki and coworkers [50, 51] on [0/y2/90]s laminates also point out that
the angle of intersection between the 90 - and y-plies and the thickness of the
y-plies may have a significant impact on the initiation and growth of cracking
in these plies.
In more general crack systems in multiple orientations, another important
consideration is the interaction between cracks in two adjacent plies. This inter-
action can enhance cracking in a certain orientation thereby causing further
stiffness degradation. For instance, in quasi-isotropic laminates 45 cracks pro-
mote enhanced cracking in the 90 -plies. This intra-ply crack interaction is a
complex function of the relative crack positions, orientations, crack sizes (ply
thicknesses), and density of cracks in different orientations. This makes determin-
ation of stresses in cracked laminates generally impossible to solve analytically,
194 Damage progression

Figure 6.15. A representative cell illustrating multiple cracking systems in a [0/90/y1/y2]s


laminate.

and numerical computations such as 3-D FEM are then necessary. Figure 6.15
depicts a representative cell of a [0/90/y1/y2]s laminate with cracks in multiple off-
axis orientations.

6.4 Modeling approaches

In the following we describe the main approaches to predicting the evolution of


ply cracking in composite laminates. In the models discussed the ply cracks are
assumed to be fully developed through the ply thickness as well as in the specimen
width direction. In a given array of ply cracks, all cracks are assumed to be of the
same size, shape, and orientation. The problem to address here is the crack
multiplication process as a function of applied loading. As described above, there
are two directions in which this problem can be pursued: one based on a strength
criterion (i.e., point failure), and the other based on an energy criterion (i.e.,
surface formation). This section deals exclusively with cross-ply laminates; multi-
directional laminates are considered later. The basic BVP to solve here is the same
as that approached earlier in Chapter 4, see Figure 4.7. We will use the same
symbols and notations (Section 4.4), unless specified otherwise.

6.4.1 Strength-based approaches


According to these models, microcracks form when the local stress (or strain) state
in a ply reaches a critical level [37–39, 81–88]. The most commonly used values of
critical level are: failure strain (e1T) or failure stress (stu) of a ply in transverse
tension. Other lamina failure criteria such as Tsai–Wu and Hashin’s criteria,
described in Section 2.2.4, are also utilized. Since the stress state at the onset of
transverse cracking is not generally uniform [45], these models fail to account for
differences in crack initiation (as a material point failure process) and crack
progression (as a surface growth process). Consequently, the ply thickness effect
on transverse cracking cannot be properly treated by these criteria.
One problem with strength-based criteria, and more generally with all proced-
ures that require knowledge of the local stress states for failure assessment, is that
6.4 Modeling approaches 195

determining the local stresses analytically is possible only in a few cases, mostly for
cross-ply laminates. The larger problem, however, is the lack of agreement of
strength-based predictions with the experimental data. Use of statistical strength
concepts [83, 84, 86, 89, 90] may improve predictions, but require additional
material data. In the following we provide a brief overview of the strength-based
models.
A basic shear lag analysis applied to determination of crack initiation strain
and crack multiplication was carried out in [91]; it has been covered previously
in Section 4.2. It is a one-dimensional analysis, and therefore does not provide
accurate stress perturbation caused by cracking. The analysis of multiple
cracking in a unidirectional fiber composite was later applied to the case of
transverse cracking in cross-ply laminates by Garrett and Bailey [38] (see Section
4.4.1). According to the analysis, the load shed by the transverse plies in the
crack plane and transferred back to the transverse plies over the distance y is
given as [38]
 
F ¼ 2t0 ws0 1  eby ; ð6:6Þ

where t0 is the thickness of the 0 -ply, w is the specimen width, Ds0 is the maximum
additional stress on the longitudinal ply as a result of cracking in the transverse ply
(occurring in the plane of crack and decaying exponentially away from the crack
surface), and b2 is the shear lag parameter, defined in Eq. (4.62), as
 
1 1
b2 ¼ G90xz0 þ ; ð6:7Þ
E90
x0 lE0x0

where G90 xz0 is the initial in-plane shear modulus of the 90 -ply, l = t0 / t90 is the
ply thickness ratio, and Ex0 and Ex0 are the longitudinal moduli for the 0 - and
0 90

90 -plies, respectively. The transverse ply will fail in tension when


F ¼ 2t90 wstu ; ð6:8Þ
where stu is the transverse ply strength. The first crack is assumed to form
in the middle of the specimen length at Ds0 = (t90 / t0) stu, i.e., at an applied
load of s0 = Ecetu, where Ec is the longitudinal modulus for the composite and
etu is the transverse ply cracking strain. Next, the cracking process will cause
second and third cracks (one above and one below the first crack) to form
simultaneously. From Eq. (6.6) and Eq. (6.8) with y = l, where 2l is the crack
spacing, Ds0 will be
1 stu
s0 ¼ : ð6:9Þ
l 1  ebl
These cracks will perturb the force transferred such that the new cracking process
will occur at
1 stu
s0 ¼ : ð6:10Þ
l 1 þ ebl  2ebl=2
196 Damage progression

4
Crack spacing (mm)

0
0 100 200 300
Applied stress (MN m–2)

Figure 6.16. Prediction of crack spacing as a function of applied stress using the shear
lag model [38]. Reprinted, with kind permission, from Springer Science+Business Media:
J Materials Science, Multiple transverse fracture in 90 degrees cross-ply laminates of a glass
fiber-reinforced polyester, Vol. 12, pp. 157–68, K. W. Garrett and J. E. Bailey.

Similarly, the (N+2)th crack formation will occur when

1 stu
s0 ¼ : ð6:11Þ
l 1 þ e blN  2e2Nbl


The evolution of crack density thus can be generated using the above iterative
scheme. Figure 6.16 shows the model prediction of crack spacing variation with
increasing loading for a glass-polyester specimen of length 130 mm and transverse
ply thickness of 3.2 mm. Although this model represents the general trends of the
experimental results it generally underestimates the average crack spacing. This is
supposedly due to the constant strength of the 90 -ply resulting in new cracks
midway between existing cracks. Later models tried to address this issue by
representing 90 -ply strength as a probabilistic function. Manders et al. [92]
considered a two-parameter Weibull distribution of the strength along the length
of the 90 -ply and represented the risk of rupture per unit volume as
s
v
pðsÞ ¼  ; ð6:12Þ
s
where the constant s* is the scale parameter in terms of stress, and v is the shape
parameter. The cumulative distribution function for failure is given as
6.4 Modeling approaches 197

Crack spacing (mm) 30

20

10

0
0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2
Strain (%)

Figure 6.17. Variation of average crack spacing for a [0/90/0] glass/epoxy composite laminate
as predicted using the probabilistic shear lag model by Manders et al. [92] for. Solid and
open circles represent experimental data from two specimens, while the solid line represents
the model prediction.
2 3 2 3
ð ð
SV ¼ 1  exp4 pðsÞ dV 5 ¼ 1  exp4A pðsÞ dy5; ð6:13Þ
V L

where A is the area, and L is laminate length. Taking a logarithm of all the terms in
Eq. (6.13) gives
ð
ln ð1  SV Þ ¼ A pðsÞ dy  ApL: ð6:14Þ
L

The quantity Ap is obtained by plotting ln(1  SV) against L. The Weibull param-
eters s* and v are obtained by fitting the experimental crack density evolution with
the model. To describe the stress s in the laminate, Manders et al. [92] used initial
shear lag analysis, covered previously in Section 4.4.1. The model predictions of
crack spacing variation as a function of applied strain are shown in Figure 6.17.
A similar analysis was performed by Fukunaga et al. [83] who also used the
Weibull distribution for 90 -ply strength but instead used interlaminar shear lag
analysis, also covered in Section 4.4.1. The effects of thermal residual stresses and
Poisson’s contraction are also included in this analysis. According to their analysis
the applied axial stress and crack spacing (2l) are related as
"   #
Ec  L0 t1 1=v 90
sc ¼
s  sxx R ; ð6:15Þ
Q22 1  Q12 A12 2lt90 d1
Q22 A22

1=v
where s ¼ s90xx0 ðln 2=L0 t1 wÞ with L0, w, and t1 being
Ðl the length, width, and
thickness of a single 90 -ply, respectively, and d1 ¼ 1l 0 ½1  ðcosh bx= cosh blÞdx is
a parameter that reflects the effect of stress nonuniformity on 90 -ply strength.
The corresponding strains can be calculated from the stress–strain relations for the
cracked laminate derived in [83] as
198 Damage progression

   
sc t90 Q22 Q12 A12 tanh al t90 s90
xx R tanh al
ec ¼ 1þ 1 þ : ð6:16Þ
Ec t0 Q11 Q22 A22 al t0 Q11 al
The crack initiation strain can be obtained by setting tan al / al = 0 (as l ! 1) in
Eq. (6.16), and 2l = L0, d1 = 1 in Eq. (6.15). Therefore, the applied strain at first
cracking in the 90 -ply is given by
"   #
1 t1 1=v
e0 ¼
s  s90
xx R : ð6:17Þ
Q22 1  Q 12 A12 t90
Q22 A22

These statistical descriptions of ply strength yielded good results for cross-ply
laminates. But these models cannot account properly for effects of changes in ply
thickness. Also, as mentioned before, shear lag analysis is a one-dimensional stress
analysis and therefore cannot be accurate. More recently, 2-D shear lag analysis
based on Steif’s parabolic displacement variation in 90 -ply [93] have also been
tried [55, 94–96]. A detailed discussion on the probabilistic concept will be covered
later in Section 6.5.

6.4.2 Energy-based approaches


The origin of energy-based approaches to crack extension lies in fracture mechanics
(see Section 2.4 for linear elastic fracture mechanics). In the classical version of
brittle fracture, the energy considerations lead to the condition that crack tip growth
becomes unstable when the energy release rate is equal to or greater than the fracture
toughness of the material containing the crack, i.e., G  Gc. This material property is
obtained by an independent test, which has been standardized in some way. For
perfectly brittle fracture, i.e., when no other energy dissipating mechanism other
than that expended in the crack surface formation exists, the fracture toughness
equals twice the surface energy. In multiple cracking within a composite laminate,
however, the situation is not as assumed in the brittle crack extension of a crack tip
for which the criterion stated above applies. Here, a ply crack on extension is
arrested at the ply interfaces and any further input of energy to the laminate supplied
by external loads goes into the formation of more ply cracks. Although the individ-
ual ply crack goes through the stages of through-thickness growth and growth in the
fiber direction, in most cases the ply cracks form quickly and the analysis is therefore
focused on their multiplication, i.e., an increase in crack density. Because of this, a
modification of conventional fracture mechanics, called finite fracture mechanics
[97], has been proposed. Thus, in contrast to conventional fracture mechanics,
transverse cracking comprises events that involve a finite amount of new fracture
area. For cross-ply laminates, finite fracture mechanics coupled with variational
stress analysis has been used in several works [5, 33, 46, 97–99].
The criterion for finite surface formation under brittle fracture condition can be
written as
  g
A; ð6:18Þ
6.4 Modeling approaches 199

where DG is the energy change (release) during crack surface increase by area
DA and g is the surface energy per unit area of the new crack formed.
In the following we describe the energy-based approaches combined with dif-
ferent approximate stress analyses used.

Shear lag analysis


Among the early energy-based analyses for ply cracking is that proposed by
Parvizi et al. [81]. They recognized the role of constraint of the outer 0 -plies to
transverse cracking in [0/90]s laminates and took account of it in the energy
balance during cracking. Recalling from ACK theory, Section 4.2, a crack does
not form in a specimen loaded in constant tension until
W  US þ UD þ 2gm Vm ; ð6:19Þ
where Vm is the matrix volume fraction and, defined per unit cross-sectional area
of the composite, DW is the work done by applied stress, DUS is the increase in
energy stored in the composite volume, UD is the energy lost by some dissipative
processes during cracking (sliding friction between debonded fibers and matrix),
and gm is the matrix surface energy per unit surface area. For the present case of
cracking in a transverse ply, the above inequality becomes
t90
W  US þ 2gt ; ð6:20Þ
h
where gt is the surface energy of the transverse ply for cracking in a direction
parallel to the fibers, and h = t0 + t90. For a linear elastic body the work of
external forces equals half the stored strain energy, i.e.,
1
US ¼ W: ð6:21Þ
2
Substituting Eq. (6.21) into Eq. (6.20), the cracking will occur when
t90
W  4gt : ð6:22Þ
h
On cracking an additional stress is thrown on to outer uncracked plies, and the laminate
increases in length. The work done during this process can be derived as [81]

2E90 2
x0 Ec etu
W ¼ ; ð6:23Þ
lE0x0 b
where b is as defined earlier in Eq. (6.7). Combining Eqs. (6.22) and (6.23), the
strain to initiate cracking in transverse ply is given by
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
min 2t0 E0x0 gt b
e0 ¼ etu ¼ : ð6:24Þ
hE90 x0 Ec

The model predictions for various transverse ply thicknesses are shown in Figure
6.18 against experimental data for glass/epoxy cross-ply laminates.
200 Damage progression

u (%)
Transverse cracking strain e tu , e tmin
2.5 Experimental etu
min
Theoretical etu
2.0

1.5

1.0

0.5

0.0
0 1 2 3 4
Transverse ply thickness 2d (mm)

Figure 6.18. Values of crack initiation strain emin


tu as a function of ply thickness 2t90 = 2d
(from Eq. (6.24)) and experimental data for various ply thicknesses in glass fiber/epoxy
cross-ply laminates. The horizontal line depicts the limiting values of emin
tu for large inner-ply
thicknesses [81]. Reprinted, with kind permission, from Springer Science+Business
Media: J Mater Sci, Constrained cracking in glass fiber-reinforced epoxy cross-ply laminates,
Vol. 13, 1978, pp. 195–201, A. Parvizi, K. W. Garrett and J. E. Bailey.

(a) (b)
z

A B A C B

s =2l l1 l2
State 1 State 2

Figure 6.19. Progressive multiplication of ply cracks in transverse layer of cross-ply laminate:
(a) state 1 with crack spacing, s = 2l; (b) state 2 with an additional crack in the ligament
AB at location C.

A more sophisticated energy-based analysis of progressive ply cracking using


one-dimensional stress analysis was performed by Laws and Dvorak [100]. (Refer
to Section 4.4.1 for the shear lag analysis.) Consider a cracked cross-ply laminate
in state 1 where cracks are a distance 2l apart, and the ligament AB between
the cracks is as yet uncracked (Figure 6.19(a)). When the applied load reaches a
critical value sc, a new crack appears in this segment at some location C (Figure
6.19(b)). Assuming that the load is kept fixed during formation of this additional
crack, Laws and Dvorak calculated the energy released during cracking of lamin-
ate of width w as
 2  
2t290 whEc 90 E90
x0 bl1 bl2 bl
 ¼ sxxR þ 0 sc tanh þ tanh  tanh ; ð6:25Þ
bt0 E0x0 E90
x0 Ex0 2t90 2t90 t90
6.4 Modeling approaches 201

   
where b2 ¼ Kt90 1=E90 x0 þ 1=lEx0
0
is the shear lag parameter (see Eq. (4.67)),
90
and sxxR is the thermal residual stress in the transverse ply. Substituting Eq. (6.25)
in Eq. (6.18), and using DA = 2wt90, a new crack will form if
 2  
t90 hEc 90 E90
x0 bl1 bl2 bl
sxxR þ 0 sc tanh þ tanh  tanh  g: ð6:26Þ
bt0 E0x0 E90
x0 Ex0 2t90 2t90 t90

The first ply failure stress (the applied stress at crack initiation) will be given by the
limit l ! 1, i.e.,
 1=2
bt0 E0x0 Ec g Ec
sfpf
c ¼  90 s90
xxR : ð6:27Þ
t90 hE90x0 E x0

The thermal residual stress, s90


xxR , and other parameters in the expression above
are known from property data for an undamaged laminate. Therefore,
Eq. (6.27) provides a relation among sfpf c , b, and g. Now sc
fpf
and g can
be determined from experimental data. Therefore, Laws and Dvorak regard
Eq. (6.27) as the relation that determines the shear lag parameter b. Once
b is known, the applied stress needed to cause cracking at location C is given
by
  
Ec 90 bl1 bl2 bl 1=2 Ec 90
sc ðl1 Þ ¼ sfpf
c þ sxxR tanh þ tanh  tanh  90 sxxR :
E90
x0 2t90 2t90 t90 Ex0
ð6:28Þ
In a practical scenario, the location C is random due to spatial variation of the
resistance to crack formation. Let p be the probability density function for the
next crack to occur at a given location. In a laminate which already contains
cracks with normalized crack density, rc = t90 / l, the expected value of the applied
stress to cause additional cracking is then
Ð 2l
E½sc ðrc Þ ¼ 0 pð xÞsc ð xÞ dx: ð6:29Þ

Three possible choices for p(x) are:

Case 1: The next crack occurs midway, so

pð xÞ ¼ dðx  lÞ; ð6:30Þ


where d(x) is the Dirac delta function.
Case 2: All locations are equally likely. Therefore,
1
pð x Þ ¼ : ð6:31Þ
2l
Case 3: p(x) is proportional to the stress at the location. For this case, p(x) is
given by [100] as
202 Damage progression

Figure 6.20. Crack density evolution in a glass/epoxy [0/90]s laminate for three choices of
probability distribution function p(x) [100]. The experimental data are from [8]. Note: [8]
has additionally three data points at the high load end that fall away from model
predictions. Reprinted, with kind permission, from N. Laws and G. J. Dvorak, J Compos
Mater, Vol. 23, pp. 900–16, copyright # 1988 by Sage Publications.

2 3
bx
  cosh
E90 6 t90 7
pð x Þ ¼ s90 x0 6 7:
xxR þ 0 sc 41  bl 5
ð6:32Þ
Ex0
cosh
t90

For case 1, the solution is explicitly given by


  
fpf Ec 90 b b 1=2 Ec 90
E½sc ðrc Þ ¼ sc þ 90 sxxR 2 tanh  tanh  90 sxxR : ð6:33Þ
Ex0 2rc rc Ex0
For cases 2 and 3, the integral in Eq. (6.29) must be evaluated numerically.
The model predictions for these cases of p(x) are shown in Figure 6.20 (with
g = 193 J/m2 and b = 0.9). Based on comparison with the experimental data,
Laws and Dvorak argue that fracture mechanics-based p(x) (case 3) is the most
promising choice. With this choice, model predictions also compare well with
another set of experimental data from [1] for graphite/epoxy laminates.

Variational analysis
Nairn [99] used the variational approach [101] for cracked cross-ply laminates,
including thermal residual stresses, in conjunction with the energy release rate
criterion to predict crack densities in cracked cross-ply laminates. His predictions
showed good agreement with experiments when the critical energy release rate for
matrix cracking, defined as such, was deduced from test data rather than evaluat-
ing it independently.
Another damage evolution model for cross-ply laminates is by Vinogradov and
Hashin [97, 98, 102]. It uses variational analysis [101] for stress computation and
finite fracture mechanics [97] for cracking criterion. Recalling from Section 4.6,
6.4 Modeling approaches 203

where the stress calculations are described in detail, complementary energy change
due to presence of N transverse cracks can be derived as [98]

X
N  2 2 X
N
 ¼  n ¼ s90
xx0 t90 C22 wðrn Þ; ð6:34Þ
n¼1 n¼1
  1
where C22 ¼ ðl þ 1Þ 3l2 þ 12l þ 8 (see Eq. (4.151)), and
60E2

d3 fn    cosh ð2a1 rn Þ  cosð2a2 rn Þ
wðrn Þ ¼  3 
¼ 2a1 a2 a1 2 þ a2 2 ; ð6:35Þ
dx rn a1 sin ð2a2 rn Þ þ a2 sinh ð2a1 rn Þ

and the summation in Eq. (6.34) is over all blocks bounded by adjacent cracks. It
is noted that rn = ln/t90 is the normalized crack spacing and should not be
confused with the crack density. For clarity of notations, the reader is referred
to Section 4.6. Initially when the cracks are far apart (rn ! 1), the function
achieves its asymptotic value given by
 
wð1Þ ¼ 2a1 a1 2 þ a2 2 : ð6:36Þ

Now consider the state of a laminate with N cracks. When a new crack appears,
the energy release can be expressed as [97]
   
n ¼  s Nþ1   s N ; ð6:37Þ

where sN and sN+1 are the stress fields before and after formation of a new (N+1)th
crack. The stresses include both mechanical and thermal effects (if present).
Assuming that the new crack appears instantly, both stress fields are evaluated
at the same external load. Equation (6.37) can be rewritten as
         
n ¼  sNþ1   s0   sN   s0 ; ð6:38Þ

where s0 is the stress field in the undamaged material at the same external load.
Assuming Eq. (6.34) is a good approximation of the energy, the energy release
due to the (N+1)th crack is given by

 2 2 N þ1 
X   90 2 2 X
N  
n ¼ s90
xx0 t90 C22 w rNþ1
i  sxx0 t90 C22 w rNi ; ð6:39Þ
i¼1 i¼1


where s90 N
xx0 is the stress in the undamaged 90 -ply, ri is the nondimensional crack
spacing for the ith block for the Nth cracking step (going from N to N+1 cracks).
Using the energy release rate criterion, Eq. (6.18), taking averages and express-
ing variables as continuous variables, the energy released during cracking can be
expressed as (see [98] for full derivation)
 
 2 d w 2
g ¼  s90
xx0 t C
90 22  :
r ð6:40Þ
d 
r r
204 Damage progression

Using the lower bound of the laminate longitudinal modulus (see Section 4.6) by
Hashin [101]
1 1 t90 w

¼ þ k12 C22 ; ð6:41Þ
Ex E0 h r
where k1 ¼ s90 0
xx 0 =sxx 0 if the temperature change is absent. Using A ¼ L=2
r, we
finally obtain the cracking criterion as
 
1 2 d 1
g ¼ s90 V; ð6:42Þ
2 xx0 dA Ex
where V is the laminate volume. Equation (6.42) represents a particular homo-
thermal case of the general fracture criterion derived by Hashin [103] given by
 
1 @S @a 1 @cp T 2
g¼ s  þ
s T 
s V; ð6:43Þ
2 @A @A 2 @A Tr
where S* is the effective elastic compliance tensor, a* is the effective thermal
expansion tensor, cp is the effective specific heat of a composite, and Tr is the
reference temperature.
The damage evolution predictions using this approach are very accurate when
the probabilistic distribution of the energy release rate is utilized. This will be
discussed in the next section.
Nairn’s original result [48, 99] was quite similar to the fracture criterion
stated above. However, in place of the usual energy release rate (2g) he used
the matrix fracture toughness (Gm) and suggested that it could be obtained
through fitting experimental data for ply cracking. If we assume that the
new crack forms midway between the existing cracks, Nairn’s fracture criter-
ion is
 
E2 aT 2
Gm ¼ s2c 22 þ 2 t90 C22 ½2wðr=2Þ  wðrÞ; ð6:44Þ
Ec C00
where C00 = (1/E2) + (1/lE1) (see Eq. (4.151)). An alternative formulation is to
allow the new crack to form anywhere between two existing cracks. If the prob-
ability of crack formation at any position is proportional to the tensile stress, the
energy release rate is given by
 
E2 aT 2
Gm ¼ s2c 22 þ 2 t90 C22 ½wðdÞ þ wðr  dÞ  wðrÞ; ð6:45Þ
Ec C00
where
r=2
Ð
½wðdÞ þ wðr  dÞ½1  fðr  2dÞdd
0
½wðdÞ þ wðr  dÞ  wðrÞ ¼  wðdÞ: ð6:46Þ
r=2
Ð
½1  fðr  2dÞdd
0
6.4 Modeling approaches 205

300

Applied stress (MPa)

200

100

0
0.0 0.2 0.4 0.6 0.8 1.0
Crack density (1/mm)

Figure 6.21. The applied stress as a function of crack density in a [0/903]s glass/epoxy
composite. The squares are data from [8]. The solid line is the energy release rate analysis fit
using Eq. (6.45) with Gm = 330 J/m2 and assuming that the initial level of the residual
thermal stresses in the 90 -plies was 13.6 MPa [99]. Reprinted, with kind permission,
from J. A. Nairn, J Compos Mater, Vol. 23, pp. 1106–29, copyright # 1989 by Sage
Publications.

For cracking in [90m/0n]s laminates, the expressions for energy release rate remain
the same except that the constant C22 is now given as C22 = (l+1)(3+12l+8l2)
(1/60E2) (see Eq. (4.164)). It is noted that Nairn’s analysis included residual thermal
stresses, which mainly change the crack initiation strain (discussed later in detail).
The parameter Gm is evaluated by fitting the model to the experimental data. The
model predictions for a [0/903]s glass/epoxy laminate with Gm = 330 J/m2 and a
thermal residual stress of 13.6 MPa are shown in Figure 6.21.

Plain-strain formulation
McCartney [104–109] developed a model based on the Gibbs free energy, instead
of complementary strain energy as described above. He used a plane strain
formulation for the estimation of elastic moduli of the damaged laminate, which
has been covered in Section 4.7. Consider the damage progression from a state of
m cracks to n cracks in the 90 -ply and assume that each crack formation occurs
under conditions of fixed applied tractions. Based on energy considerations, crack
formation will occur when
 þ G  0; ð6:47Þ
where DG is the change in Gibbs free energy, and DG is the energy absorbed in the
volume V of laminate due to the formation of new cracks, given by
 ¼ V ½ðon Þ  ðom Þ; ð6:48Þ
where o denotes the damage parameter which characterizes the crack density
in the laminate. The corresponding change in the Gibbs free energy can be
written as
206 Damage progression

ð
G ¼ ½gðon Þ  gðom ÞdV; ð6:49Þ
V

where g(o) represents the Gibbs free energy per unit volume. After some math-
ematical treatment (see [109] for details) the cracking criterion becomes

½^eðon Þ  ^eðom Þ2


þ mA ðon Þðgðon ÞÞ2  mA ðom Þðgðom ÞÞ2  2½ðon Þ  ðom Þ > 0;
1 1

E~ðon Þ E~ðom Þ
ð6:50Þ
where ^eðoÞ; E~ðoÞ; mA ðoÞ; gðoÞ, and G(o) denote axial strain, axial Young’s modu-
lus, in-plane axial shear modulus, applied in-plane shear strain, and the energy
absorption per unit volume for length 2L of laminate, respectively, for given
damage state o. G(o) is here given by

hð90Þ X
M
ð90Þ ð90Þ
ðoÞ d gj ; ð6:51Þ
hL j¼1 j

where 2h(90) and 2h represent the total thickness of 90 -plies and whole laminate,
respectively; M is the number of potential cracking sites in the 90 -plies, which are
ordered from the top to bottom of the plies, taken in increasing order from the
ð90Þ
center of the laminate to the outside; and 2gj is the fracture energy for the jth
potential cracking site. The expressions for other parameters can be found in [109].
The crack initiation strain can be obtained by setting om = o0 (undamaged state)
in Eq. (6.50).

COD-based models
Following the work of Parvizi et al. [81] and Wang and Crossman [110] on
energy release rates to study the formation of cracks in cross-ply laminates, Joffe
and coworkers [111–114] considered fully developed cracks and developed
a methodology to predict the multiplication of transverse cracks based on the
virtual crack closure technique. The idea is to probe the region between two
existing cracks and introduce a virtual crack. For the introduced virtual
crack the work performed to close the crack surfaces is calculated and compared
with the energy needed to create a crack, the critical energy release rate (Gc),
at this position. A crack is taken to form when the work to close the crack
exceeds Gc.
Consider a damaged [0m/90n]s laminate with a periodic system of “N” self-
similar cracks with spacing s = 2l in the 90 -ply (Figure 6.19(a)). At applied

laminate stress s0 (and corresponding far-field stress s90
x0 in the 90 -layer) a new
crack develops midway between two existing cracks and the total number of
cracks becomes 2N with spacing l (Figure 6.19(b)). According to the crack closure
concept the released energy due to these N new cracks is equal to the work needed
6.4 Modeling approaches 207

to close them. If we denote this work by W2N!N and the work to close all cracks
simultaneously by W2N!0, energy balance requires
W2N!0 ¼ W2N!N þ WN!0 ; ð6:52Þ
where the work to close N cracks with spacing s is

90  2
1 s90
WN!0 ¼N
2
s90
xx0 uðzÞ dz ¼ 2Ns90 n ðlÞ
xx 0 t90 u ¼ N xx0 t290 u~n ðlÞ; ð6:53Þ
2 E2
t90

where unit width is assumed and u(z), un ; and uen represent the variation of the
normal crack opening displacement (COD) along the thickness direction, its
average value, and its average value normalized with respect to the remote
stress and transverse modulus for the ply, respectively. un ; and uen are thus
defined as

90
1 un
un ¼ uðzÞ dz; u~n ¼   : ð6:54Þ
t90 s90
xx0 =E 2 t90
0

Similarly, the work to close 2N cracks with spacing l is given by


 2
s90
W2N!0 ¼ 4N xx0 t290 u~n ðl=2Þ: ð6:55Þ
E2

Substituting Eqs. (6.53) and (6.55) into Eq. (6.52), the energy released by forma-
tion of a crack midway between two existing cracks of spacing s is
 2
s90
W2N!N ¼ 2N xx0 t290 ½2~
uðl=2Þ  u~ðlÞ: ð6:56Þ
E2

The cracks form when this work is greater than or equal to the cumulative surface
energy of newly created surfaces, i.e.,

W2N!N  2
N
2t90
Gc : ð6:57Þ

From Eqs. (6.56) and (6.57), the criterion for crack formation is
 90 2
sxx0
t90 ½2~
uðs=2Þ  u~ðsÞ  Gc : ð6:58Þ
2E2
To analyze cracking in an arbitrary position between two pre-existing cracks,
a new crack is introduced in an arbitrary position between the cracks (Figure 6.19
(b)), which leads to a new damage state with one crack spacing equal to s1 and the
second one equal to s2 = s – s1. The cracking criterion in this case is
 90 2
sxx0
t90 ½2~
uðl1 =2Þ  u~ðl1 Þ þ 2~
uðl2 =2Þ  u~ðl2 Þ  Gc : ð6:59Þ
2E2
208 Damage progression

[±q/904]s FEM model, energy approach


0.8
Crack density ρ (cr/mm)

0.6
q=0
q =15
0.4 q = 30
q = 40

0.2

0.0
0 50 100 150 200 250
Stress s0 (MPa)

Figure 6.22. Evolution of crack density as a function of applied stress using energy model
for [ y/904]s laminates. Symbols represent experimental data and lines represent the
average of four runs using the energy model [113]. Reprinted, with kind permission, from
Compos Sci Technol, Vol. 61, R. Joffe, A. Krasnikovs, and J. Varna, COD-based simulation
of transverse cracking and stiffness reduction in [S/90n]s laminates, pp. 637–56, copyright
Elsevier (2001).

The authors applied their analysis for the prediction of crack density evolution
in glass/epoxy [ y/904]s laminates for the case of varying crack spacing.
A Weibull distribution for Gc was utilized and the CODs were calculated
using FE analysis. The predictions are shown against experimental data in
Figure 6.22.
Adolfsson and Gudmundson [11] also developed an energy-based damage
evolution approach using their stress analysis. Basic stress analysis using this
approach is covered in Section 4.9.1, although they updated their analysis to
include bending loads, which can be found in [11, 115]. The energy model for
crack density evolution is based on changes in strain energy due to cracking. From
[11], the strain energy per unit in-plane area of the damaged laminate with n plies
may be written as

1
T
X n

wðcÞ ¼ e  aðcÞ T  eðRÞ CðcÞ e  aðcÞ T  eðRÞ þ hk T; skðRÞ ; ð6:60Þ


2 k¼1

where bold-face letters represent matrices; C(c), a(c), and e(R) are the stiffnesses,
thermal expansion coefficients and residual stress-induced eigenstrains, respect-
ively, of the cracked laminate. Expressions for these crack density quantities were
derived in [115]. In Eq. (6.60), DT is the temperature difference between the curing
and service temperature and hk are the functions containing energy stored in the
laminate due to interlaminar constraints and thermal residual stresses (their
expressions are given in Appendix A of [11]). From the strain energy, the energy
release rate for the ith cracked ply is derived as
6.4 Modeling approaches 209

600

500 (0/90/0)
(0/902 /0)
(0/904 /0)
400 (0/908 /0)
Simulations
Stress (MPa)

300

200

100

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Strain (%)

Figure 6.23. Stress–strain response of graphite/epoxy cross-ply laminates. Symbols


represent experimental data whereas the solid lines represent predictions from the model by
Adolfsson and Gudmundson [11]. Reprinted, with kind permission, from Int J Solids Struct,
Vol. 36, E. Adolfsson and P. Gudmundson, Matrix crack initiation and progression
in composite laminates subjected to bending and extension, pp. 3131–69, copyright
Elsevier (1999).

 
i @U @ AwðcÞ
G ¼ i¼ ; ð6:61Þ
@A @Ai
where Ai is the crack surface area in ply i. The area Ai is given by Ai = Ari with
normalized crack density ri = ti / li, where li is the average spacing of cracks in the
ith ply and A is the laminate in-plane area. From Eqs. (6.60) and (6.61) the energy
release rate Gi for cracking in the ith ply is given by
 T

@aðcÞ @eðRÞ
Gi ¼ i
T þ i
CðcÞ e  aðcÞ T  eðRÞ
@r @r

T @C
X N
ð6:62Þ
1 ðcÞ @hk
 e  aðcÞ T  eðRÞ e  a ðcÞ T  e ðRÞ
 :
2 @ri k¼1
@ri

The above expression contains the derivatives of the effective thermoelastic


properties of the damaged laminate with respect to the ply crack densities.
Calculating these quantities is a more complex task than determining the proper-
ties themselves. For this purpose, either FEM or the approximate analytical
expressions given in Appendix A of [11] have to be utilized. The model predictions
for the stress–strain response of graphite/epoxy cross-ply laminates are compared
with experimental data in Figure 6.23.
210 Damage progression

2.0
Prediction with Gc = 104 J/m2
sc = Ecec Prediction with Gc = 130 J/m2
Experimental data
1.5
Crack density

1.0

0.5

AS/3501–06 [02/902]s

0.0
0.0 0.5 1.0 1.5 2.0
Applied stress

Figure 6.24. Comparison between model predictions and experimental data for the COD
model of Qu and Hoiseth [116] for a cross-ply laminate, sc represents the applied stress on
the composite. Reprinted, with kind permission, from Fatigue Frac Eng Mater, Vol. 24,
J. Qu and K. Hoiseth, Evolution of transverse matrix cracking in cross-ply laminates,
pp. 451–464, copyright Wiley (1998).

Qu–Hoiseth analysis
The evaluation of moduli for damaged cross-ply laminates using the approach
proposed by Qu and Hoiseth [116] is covered in Section 4.5. The cracking criterion
for this model is derived as
      
2ec 2 t90 Ec E2 dr dr
Gc ¼ exp   exp  ; ð6:63Þ
ðE1 þ E2 Þr 2t90 t90
where Gc is the in-plane mode I fracture toughness of transverse ply; ec is the applied
strain; Ec is the plain-strain Young’s modulus of the undamaged cross-ply laminate
in the longitudinal direction; E1 and E2 are longitudinal and transverse moduli of
the ply, respectively; d is the average crack opening displacement of the 90 crack;
and r = t90 / l is the normalized crack density. The threshold strain at which the
transverse matrix cracking initiates can be obtained by setting r ! 0 in Eq. (6.63) as
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Gc ðE1 þ E2 Þ
e0 ¼ dEc E2 : ð6:64Þ

The model predictions for a cross-ply laminate made of AS/3501–06 material for
two values of Gc are shown in Figure 6.24. The experimental data are from [117].

6.4.3 Strength vs. energy criteria for multiple cracking


There is a fundamental difference between strength- and energy-based criteria
when applied to multiple ply cracking in laminates. Strength essentially represents
the failure at a material point when a specified stress component or function of
6.5 Randomness in ply cracking 211

stress components reaches a critical value. This approach is a legacy of homo-


geneous materials such as metals and ceramics where yielding at a point (in a
metal) is assumed to occur according to, e.g., the von Mises criterion, or brittle
failure (in a ceramic) is assumed to occur when the maximum tensile principal
stress reaches a critical value. In case of yielding there is no ambiguity in terms of
the stress components attaining critical value at a point since this type of “failure”
can spread spatially from one point to another as indicated by the contour of the
yield criterion. However, “brittle failure” represents the instability of crack growth,
and unless a crack exists its growth is meaningless. This difficulty has been
conventionally overcome (or bypassed) by assuming that brittle failure according
to a point-failure (strength) criterion is the concurrent formation and instability of
a crack. In an unconstrained failure case, such as the brittle failure of an unre-
inforced ceramic, this approximating assumption causes little difficulty since the
initiation of cracking and its unstable growth are usually not far apart, i.e., they
occur at roughly the same applied load. However, when the constraint to crack
growth is imposed by the presence of reinforcements, or stiff elements in the matrix
generally, then the formation of a crack and the instability of its growth are
determined by different conditions. This fact was realized in [91] when it was found
that the strength criterion was inadequate to predict the formation of multiple
cracking in unidirectional brittle matrix composites. Energy considerations were
then made by recognizing the dissipation of energy in crack surface formation.
Multiple cracking in composite materials is an inherent feature of the failure
process due to the presence of directed interfaces (fiber/matrix and interlami-
nar) that impart mechanisms of stress transfer from the cracking elements to
the noncracking elements, which in turn provide a constraint to the cracks.
Thus, the inadequate incorporation of constrained cracking in a multiple
cracking process is bound to induce error. One example, unfortunately not
uncommon, is using solutions of crack opening displacements in an infinite
medium in models of multiple cracking.

6.5 Randomness in ply cracking

Physical observations of ply cracking indicate that in the early stages of the cracking
process randomness exists in the location of cracks, their size, and how the evolu-
tion (growth and multiplication) of cracking occurs. As the cracking process
evolves, randomness tends to decrease, and as the process approaches saturation,
uniformity in crack spacing results. The causes of randomness are many, most
induced by the manufacturing process. For instance, the fiber volume fraction can
vary spatially. Image analysis reported in [118] showed that in a T300/914 carbon/
epoxy composite with an average fiber volume fraction of 55.9% the local volume
fraction ranged between 15 and 85%. Other common defects are voids and inclu-
sions in the matrix, partially cured regions of matrix, broken fibers, fiber waviness,
unbonded regions of interfaces at fiber surfaces, and between plies.
212 Damage progression

Figure 6.25. Distribution of ply cracks in a [02/904]s laminate at different times in the
load history based on the lattice scheme by Silberschmidt [119, 120]: (a) 100 cycles; (b) 4103
cycles; (c) 105 cycles; and (d) 2105 cycles [120]. Reprinted, with kind permission,
from Springer Science+Business Media: J Mater Sci, Effect of micro-randomness on
macroscopic properties and fracture of laminates, Vol. 41, 2006, pp. 6768–76, V. V.
Silberschmidt.

(a)

(b)
80
70
s− x90x(MPa)

60
50
40
30
20
10
0
0 2 4 6 8 10
(cm)

Figure 6.26. Crack distribution (a) and corresponding variation (b) of the average
longitudinal stress in the 90 -ply along the laminate length, estimated for [0/90]s glass/epoxy
laminates [121]. Reprinted from Compos Sci Technol, Vol. 60, J. M. Berthelot and J. F. Le
Corre, Statistical analysis of the progression of transverse cracking and delamination in
cross-ply laminates, pp. 2659–69, copyright (2000), with permission from Elsevier.

Several attempts exist in the literature to treat random variations of microstruc-


ture. Silberschmidt [119, 120] suggested a lattice scheme which incorporates the
effects of the initial microstructural randomness as well as a dispersed evolution of
damage and its transition to spatially localized matrix cracking. The scheme
involves mapping a dynamic matrix of stress-renormalizing coefficients onto the
lattice of elements covering the cracked (90 ) layer. Figure 6.25 shows the distri-
bution of ply cracks at different times in load history based on the scheme for a
[02/904]s T300–934 laminate during fatigue loading.
Another example of crack distribution for [0/90]s glass/epoxy laminates, tested
under quasi-static tensile loading by Manders et al. [87], is shown in Figure 6.26,
along with the average axial stress distribution along the laminate length, esti-
mated by Berthelot and Le Corre [121]. Clearly the stress state is such that it is
difficult to predict damage evolution using deterministic approaches. Probabilistic
notions are often used to correct the cracking predictions for random effects, as
stated earlier. Although some probabilistic fracture criteria have already been
discussed, we shall focus here on the variational stress model, emphasizing details
of probabilistic considerations.
6.5 Randomness in ply cracking 213

A fracture mechanics-based stochastic model to predict progression of ply


cracking was initially proposed by Wang and coworkers [1, 110, 122–125]. The
authors postulated that cracking in transverse plies of cross-ply laminates is
governed by a characteristic distribution of “effective” flaws, which are essentially
inherent material microcracks that cannot be seen until grown to macroscopic
dimensions. Thus these microcracks act individually as initiators of cracks that
propagate to form fully grown transverse cracks. The distribution of flaw size f(a)
and spacing f(S) along specimen length are assumed to follow the following
normal probability distributions
" #
1 ða  ma Þ2
f ðaÞ ¼ pffiffiffiffiffiffi exp  ;
a 2p 2s2a
" # ð6:65Þ
1 ðS  mS Þ2
f ðSÞ ¼ pffiffiffiffiffiffi exp  ;
S 2p 2s2S

where 2a is the average flaw size, S is the average distance between two adjacent
flaws, and ma, mS, sa, and sS are fitting parameters. The “worst” of the flaws causes
the first ply cracking. With increased loading, smaller flaws cause further trans-
verse cracking. The first transverse crack forms when
Gðsc ; a0 Þ ¼ Gc ; ð6:66Þ
where sc is the longitudinal stress applied to the composite, 2a0 is the initial flaw
size, and Gc is the critical energy release rate, which is assumed to be constant
along the laminate length. The propagation of the flaw will be stable if
Gðsc ; a0 þ aÞ<Gc ; ð6:67Þ
and unstable if
Gðsc ; a0 þ aÞ > Gc : ð6:68Þ
Progressive cracking will ensue when there is enough energy available for multiple
flaws to propagate into fully grown transverse cracks. For crack formation after
the first crack the energy release rate for flaw propagation depends on its relative
distance S from the existing crack and can be expressed as
Gðsc ; aÞ ¼ RðSÞG0 ðsc ; aÞ; ð6:69Þ
where G0 is the energy release rate when no crack is present, and R(S) is the energy
retention factor, with a value between 0 and 1, accounting for the presence of a
neighboring crack. Similarly, for a flaw to propagate between two existing trans-
verse cracks, the energy release rate is

Gðsc ; aÞ ¼ RðSL ÞG0 ðsc ; aÞRðSR Þ ð6:70Þ


where SL and SR are the distance of the flaw from the left and right cracks,
respectively. Chou et al. [126] implemented the approach using a Monte Carlo
scheme. The results showed a fair agreement with experimental data. Essentially,
214 Damage progression

this approach predicts the event when a micro-flaw develops into a fully grown
transverse crack, thereby predicting the multiplication of ply cracks. However, the
approach has not gained wide usage because it requires many unknown param-
eters which are found by fitting to experimental data. As mentioned earlier,
experimental observations show that the transverse cracks usually grow quickly
through the 90 -ply thickness as well as the specimen width. Therefore, more
recent approaches do not try to predict crack propagation; rather they focus on
the multiplication of cracks, i.e., an increase in crack density.
To illustrate how a more recent fracture criterion can be modified to include
probabilistic measures we follow the treatment of Vinogradov and Hashin [98].
Accordingly, the uncertainties in the cracking process can be categorized into
two probabilistic notions: “geometrical” and “physical.” The “geometrical”
uncertainty refers to the probability of a crack to appear at a certain location
between two existing adjacent cracks, while the “physical” aspect deals with the
variation of material resistance to crack formation. The geometrical aspect of
probabilistic cracking can be introduced by considering the statistical variation of
distance between adjacent cracks, i.e.,
1
ð 1
ð

r rpðrÞdr; w ¼ wðrÞp ðrÞ dr; ð6:71Þ
0 0

where p(r) is the probability density function (PDF) of distances between adjacent
cracks. The criterion for first crack formation can be found by substituting r ! 1
in Eq. (6.40) and using Eq. (6.36), to obtain
 2  90 2  2 2

g ¼ s90
xx0 t90 C22 wð1Þ ¼ 2 sxx0 t90 C22 a a þ b : ð6:72Þ

In fact the criterion in Eq. (6.72) is expected to predict the initial stage of the
damage evolution curve. For any material block between two existing adjacent
cracks, the cracking criterion in Eq. (6.40) can be rewritten as [98]
     
 90 2 rþx rx
g ¼ sxx0 t90 C22 w þw  w ð rÞ ; ð6:73Þ
2 2

where x denotes the nondimensional coordinate of the new crack location between
the two existing cracks. Equation (6.73) is a local criterion for crack formation
because it deals with the location of the next crack.
The “physical” nature of damage evolution can be achieved by having a
probabilistic variation of material property g. Thus,

g ¼ GðxÞ: ð6:74Þ

The parameter G(x) can be thought of as local toughness of the material by


arguing that it is easy to form a crack at a section which contains many flaws and
has a weak interface. The variation of g is usually described using a Weibull
distribution, i.e., the PDF of g can be expressed as
6.5 Randomness in ply cracking 215

    
 g  gmin 1 g  gmin
pr ðgÞ ¼ exp  ; g  gmin ; ð6:75Þ
g0 g0 g0

where gmin is the minimum possible value of g, and  and g0 are parameters of the
distribution, usually evaluated by fitting experimental data.
For different laminate systems, i.e., for different mixes of 0 and 90 plies,
e.g., [0n1/90m1]s and [0n2/90m2]s laminates, the distribution parameters
may not be the same. If the parameters for the first laminate configuration,
1, g01, are known (through fitting of experimental data), the parameters
2, g02 for the second laminate configuration can be found from the following
relations
       
2 þ 2 2 1 þ 2 2 þ 1 2 1 þ 1
   
2  m2 1 
  2 ¼   1 ;
1 þ  2 m1 1 þ  1
2 2
 1
 2  ð6:76Þ
1 þ 1


g02 ¼ g01  1  ;
1 þ 2

2

where G(x) represents the standard gamma function for the random variable x.
The derivation can be found in the original article [98].
The simulation procedure for this model can be summarized as follows:
1. Choose a ply material or a laminate configuration of a ply material.
2. Distribute random points for possible crack locations along the laminate
length.
3. Generate a random value of g at each point according to the Weibull
distribution.
4. Fit the model predictions to the experimental data for crack density evolution
to deduce the parameters of the Weibull distribution.
5. Calculate Weibull parameters for other laminate systems using Eq. (6.76), and
predict the crack density evolution for these laminates.
Some examples of the numerical simulation results with the fitted and calculated
parameters of the distribution are shown in Figure 6.27.
A recent strength-based analysis by Berthelot and Le Corre [121] has revealed
that the choice of probabilistic distribution should account for weakness areas in
the material. This analysis for [0/902]s carbon/epoxy laminate shows that a prob-
abilistic distribution of strength which accounts for weakness areas properly
corrects the crack density evolution in the beginning stage (see Figure 6.28).
A divergence between models with and without consideration of weakness areas
is always observed at low crack densities. This is because, initially, cracking is
preferred at weakness areas where the fracture toughness of the material is low
216 Damage progression

(a) Avimid K Polymer/IM6


1

0.9

0.8
Crack density (1/mm)

0.7
[0/903]s
0.6

0.5

0.4
[0/902]s
0.3

0.2

0.1

0
0 200 400 600 800 1000 1200 1400
Stress (MPa)

(b) Hercules AS4/3501–6

1.2

1
Crack density (1/mm)

0.8

0.6

0.4
[0/904]s [0/902]s [0/90]s

0.2

0
0 100 200 300 400 500 600 700 800
Stress (MPa)

Figure 6.27. Prediction of crack density evolution in cross-ply laminates using the
Vinogradov and Hashin model [98] for two material systems: (a) Avimid K polymer/IM6,
(b) Hercules AS4/3501-6. The experimental data are from [127]. Reprinted, with kind
permission, from Int J Solids Struct, Vol. 42, V. Vinogradov and Z. Hashin, Probabilistic
energy-based model for prediction of transverse cracking in cross-ply laminates, pp. 365–
392, copyright Elsevier (2005).

(due to inherent defects) as compared to its average value in the whole laminate.
For glass/epoxy laminates, Berthelot and Le Corre found that delamination
occurs at high crack densities and is the cause of data deviating from the model
prediction.
6.6 Damage evolution in multidirectional laminates 217

(a)
1400

1200 Experimental results


Without weakness areas
Crack density (m–1)

1000 With weakness areas

800

600

400

200

0
250 300 350 400 450 500 550
Average stress (MPa)

(b)
1400

1200 Experimental results


Without weakness areas
Crack density (m–1)

1000 With weakness areas

800

600

400

200

0
250 300 350 400 450 500 550
Average stress (MPa)

Figure 6.28. Evolution of crack density as a function of applied stress for a [0/902]s carbon/
epoxy laminate [121]. The experimental data are from [1]. Reprinted, with kind permission,
from Compos Sci Technol, Vol. 60, J. M. Berthelot and J. F. Le Corre, Statistical analysis
of the progression of transverse cracking and delamination in cross-ply laminates,
pp. 2659–69, copyright Elsevier (2000).

6.6 Damage evolution in multidirectional laminates

Although many generic features of ply cracking are evident in cross-ply laminates,
this class of laminates is used only in limited cases. Most applications require a mix
of lamina orientations in the laminate configuration to generate properties to carry
combinations of normal loads, bending moments and torsion. In a multidirectional
laminate the ply cracking in any ply will generally take place under stresses normal
to and parallel to the fibers, as well as in-plane shear. Experimental investigations
[27, 34, 35, 50, 51, 128] have clarified some of the complexities relating to mode-
mixity of crack growth and interactions between cracks within and among plies.
Consider now a [0/90/y1/y2]s laminate where ply cracks can appear in the 90 -,
y1-, and y2-plies, assuming loading in the 0 -direction. Figure 6.29 illustrates the
development of cracking in multiple orientations in such a laminate configuration.
As indicated there, cracking initiates first in the 90 -plies at an overall strain e90
0 ,
and, on increasing the load, this cracking multiplies. At the strain ey01 the y1-plies
218 Damage progression

90 q1 q2
0 0 0

Figure 6.29. Cracking process in a [0/90/y1/y2]s half-laminate.

x3

x2

tq

s/2
s

State 1: N cracks, crack spacing = s State 2: 2N cracks, crack spacing = s/2

Figure 6.30. Progressive multiplication of off-axis ply cracks in a multidirectional laminate


[129]. Reprinted, with kind permission, from Int J Solids Struct, Vol. 47, C. V. Singh
and R. Talreja, Evolution of ply cracks in multidirectional composite laminates,
pp. 1338–49, copyright Elsevier (2010).

begin cracking (assuming y1 > y2), and with a further increase in the imposed load, an
interactive cracking process continues in the 90 - and y1-ply orientations. At strain ey02
cracking initiates in the y2-plies, and eventually all off-axis plies conduct interactive
crack multiplication. The crack initiation strains and crack multiplication rates
depend on the constraint imposed by the neighboring plies to the cracks in a given ply.
For predicting the evolution of ply cracking the authors of this book have
developed an energy-based approach, which is capable of dealing with cracking
in off-axis plies of orthotropic laminates. In [129] the approach is described and
applied to several ply cracking cases. A brief description of the approach follows.
As illustrated in Figure 6.30, two damage states are considered: state 1 with N parallel
off-axis cracks spaced at distance s, and state 2 where the cracks have multiplied to 2N
and attained the spacing s/2. Evolution of cracking damage is assumed when the work
required in going from state 1 to state 2 (which is the same as the work needed to close N
cracks in going from state 2 to state 1) exceeds a critical value, i.e., if
1
W2N!N  N:Gc : tc ; ð6:77Þ
sin y

where y is the off-axis angle and Gc is the critical (threshold) value of the energy
required for multiple ply crack formation within the given laminate (more discus-
sion about this later). The work required to form N additional cracks in going from
state 1 to state 2 (the same as the work required to close those cracks) is given by
W2N!N ¼ W2N!0  WN!0 ; ð6:78Þ
6.6 Damage evolution in multidirectional laminates 219

where WN!0 and W2N!0 represent the work required to close N cracks in state 1,
and 2N cracks in state 2, respectively, and the two quantities are calculated as (see
[130] for detailed derivation)
1 1 h y 2 y  2 i
WN!0 ¼ N ðtc Þ2
s20 :~un ðsÞ þ sy120
u~yt ðsÞ ; ð6:79Þ
sin y E2
1 1 h  y 2 y s
 y 2 y s
i
W2N!0 ¼ 2N ðtc Þ2
s20 :~
un þ s120
u~t ; ð6:80Þ
sin y E2 2 2
where u~yn ; u~yt are the normalized average crack opening and sliding displacements
(COD and CSD), respectively. These are given by
ð ty =2
uy 1
u~yn ¼  y n  ¼  y  un ðzÞdz;
tc s20 =E2 tc s20 =E2 ty =2
ð ty =2 ð6:81Þ
uy 1
u~yt ¼  y t ¼  y  ut ðzÞ dz;
tc s120 =E2 tc s120 =E2 ty =2

where un and ut represent the relative opening and sliding displacement of the
cracked surfaces, respectively, and overbars represent averages. For the special
case of cracking in the 90 -ply only, the sliding displacement is zero and hence the
criterion for ply crack multiplication is written as
 2
sy20 h y s
i
tc : 2:~
un  u~yn ðsÞ  GIc ; ð6:82Þ
E2 2

where GIc is the critical energy release rate in mode I (crack opening mode).
This is the same relation as that derived for cracking in cross-ply laminates by
Joffe et al. [113] except that they consider centrally placed cracked 90 -plies in
their model and normalize the average COD with half the ply thickness (tc/2).
For cracking in a general off-axis ply, one can use a multi-mode criterion
given as
 M  
wI wII N
þ  1; ð6:83Þ
GIc GIIc
where
 2  2
sy20 tc h y s
i sy120 tc h y s
i
wI ¼ 2:~
un  u~yn ðsÞ ; wII ¼ 2:~
ut  u~yt ðsÞ ; ð6:84Þ
E2 2 E2 2
where GIIc is the critical energy release rate in mode II (crack sliding mode), and
the exponents M and N depend on the material system, e.g, for a glass/epoxy
system, M = 1, N = 2 [130].
In our work [129] we interpret the critical material parameters GIc and GIIc not
in the usual linear elastic fracture mechanics sense where they are defined as the
resistance to advancement of the crack front at the point of unstable crack growth.
220 Damage progression

Instead, we postulate that the work required to go from state 1 to state 2 involves a
range of dissipative processes that all depend on the material condition in a
cracking ply within the given laminate. The material parameter representing the
dissipated energy per unit of ply crack surface is, therefore, not what is obtained in
a standard fracture toughness test for determining GIc or GIIc. To emphasize that
the critical energy terms used here are not the usual fracture toughness values GIc
or GIIc, we will henceforth use the symbols WIc and WIIc instead. These new
quantities are not to be obtained by independent tests, but are to be evaluated
by fitting model predictions (Eq. (6.83)) to the experimental data for a reference
laminate. This way the values obtained will be representative of the energies
associated with multiple cracking within a laminate. A reference laminate is
chosen from the class of laminates (material etc.) for which predictions are to be
made, and for which experimental data are readily available [129, 131]. Further-
more, as described in [129], it is argued that a ply crack within a laminate cannot
be formed unless sufficient energy is available to open its surfaces (i.e., in mode I
cracking). In other words, a pure sliding action will not by itself generate the set of
parallel cracks illustrated in Figure 6.30. This will imply that the second term in
Eq. (6.84) is negligible in comparison to the first term. With these assumptions and
approximations the predictions of crack density evolution agree well with experi-
mental data [129].
The complete procedure to implement the described energy model for micro-
crack initiation and evolution in an off-axis ply of a general symmetric laminate is
outlined below. The procedure is in two parts:
Part I: Estimate WIc

1. From FE simulations, determine the variation of normalized COD and CSD


(Eq. (6.81)) with crack spacing.
2. Assume a value for WIc. Plot the damage evolution for the reference laminate as
follows:
(a) Divide the specimen length into small intervals of length, dX = ty/10 is
chosen here.
(b) Find the multiple crack initiation strain, Eq. (6.82) with the COD value
calculated with a very large spacing (s ! 1).
(c) Assume a small initial crack density, e.g., rinitial = 1/50ty is chosen here.
(d) Choose a random length interval and check for cracking. A new crack forms
when the criterion set in Eq. (6.82) is satisfied. Increase the crack density and
eliminate the cracked length interval from further consideration for ply cracking.
(e) Choose another length interval and repeat the previous step until the
fracture criterion is satisfied.
(f) Increase the applied strain. Repeat steps (d) and (e) using this strain value.
3. Iterate step 2 by varying WIc so that the resulting evolution curve fits the
experimental data for the reference laminate. For example, for predicting the
damage evolution in [0/ y4/01/2]s laminates, we chose [0/908/01/2]s as the refer-
ence laminate.
6.6 Damage evolution in multidirectional laminates 221

4
Model
Crack initiation strain (%) 3.5 Experimental Data

2.5

1.5

0.5

0
40 50 60 70 80 90
Ply Orientation (deg)

Figure 6.31. Variation of crack initiation strain with ply orientation for a glass/epoxy
[0/ y4/01/2]s laminate. The experimental data are from [27].

Part II: Predict the damage evolution for other off-axis plies:

1. From FE simulations, determine the variation of COD and CSD (Eq. (6.81))
with crack spacing for a given off-axis laminate.
2. Using the value for WIc obtained above, predict the damage evolution by
following steps 2(a)–(f) described in Part I.
The above semi-analytical model is coded in a MATLAB program. The input
data include the following laminate properties: ply material (elastic moduli), ply
thicknesses and orientations (i.e., laminate layup), and variation of COD with
respect to crack density (which can be obtained from independent 3-D FE
analysis).
The energy model described above was applied to predict damage evolution
in glass/epoxy [0/ y4/01/2]s, quasi-isotropic ([0/90/∓45]s), and [0m/90n/ yp]s
laminates [129]. Figure 6.31 shows the variation of crack initiation strains with
off-axis ply orientation (y) for [0/ y4/01/2]s laminates. As expected, the crack
initiation strain increases as y decreases and it may exceed 1.5% if y < 45 . In
fact, the experiments by Varna et al. [27] for this laminate revealed that ply
cracks did not form fully for y < 40 . Using the procedure described above, WIc is
evaluated by fitting model predictions with experimental data for a chosen
reference laminate [0/908/01/2]s. The evolution of crack density against applied
strain for these laminates are shown in Figure 6.32 and Figure 6.33 for y = 70
and 55 , respectively. For these laminates, direct application of Eq. (6.84)
does not yield accurate predictions. The reason is that at low crack densities,
the work term wI is almost constant (i.e., independent of crack spacing). How-
ever, on development of sufficient cracks with a distribution in inter-crack
spacing, wI depends on the crack spacing due to interactions between adjacent
cracks. To account for this behavior, Liu and Nairn [127] suggested that the
effective crack spacing be used in place of the average crack spacing s. Thus, wI is
modified as
222 Damage progression

Figure 6.32. Damage evolution in [0/ 704/01/2]s laminates. The experimental data are from
[27]. The crack density is average of crack densities in +70 and 70 -plies. Reprinted, with
kind permission, from Int J Solids Struct, Vol. 47, C. V. Singh and R. Talreja, Evolution of
ply cracks in multidirectional composite laminates, pp. 1338–49, copyright Elsevier (2010).

Figure 6.33. Damage evolution in [0/ 554/01/2]s laminates. The experimental data are from
[27]. The crack density is the average of crack densities in +55 and 55 -plies. Reprinted,
with kind permission, from Int J Solids Struct, Vol. 47, C. V. Singh and R. Talreja, Evolution
of ply cracks in multidirectional composite laminates, pp. 1338–49, copyright Elsevier (2010).

 2    
sy20 tc y fs y
wI ¼ 2:~
un  u~n ðfsÞ ; ð6:85Þ
E2 2

where the parameter f is the average ratio of the crack interval in which a
microcrack forms to the average crack spacing. For [0/ y4/01/2]s laminates, the
predictions shown in Figure 6.32 and Figure 6.33 are made with f = 0.8.
The same model was also used to predict damage evolution in a quasi-isotropic
laminate. The value of WIc for this case was obtained by fitting model predictions
with experimental data for reference [0/90]s laminate. Since 45 -plies in the quasi-
isotropic laminates contained partially grown cracks, this was accounted for in the
analysis by considering cracks in multiple orientations while calculating CODs
from FE analysis. This showed that COD for the 90 -ply increased resulting in
6.7 Damage evolution under cyclic loading 223

Figure 6.34. Evolution of 90 -crack density in [0/90]s and [0/90/ 45]s laminates. The
experimental data are from [34]. Reprinted, with kind permission, from Int J Solids Struct,
Vol. 47, C. V. Singh and R. Talreja, Evolution of ply cracks in multidirectional composite
laminates, pp. 1338–49, copyright Elsevier (2010).

enhanced cracking in the transverse ply. The model predictions for crack density
evolution in this case are compared with experimental data in Figure 6.34. No
modifier for crack spacing was needed for this case, i.e., f = 1.
A parametric study performed for [0m/90n/ yp]s laminates reveals that inter-
actions between the crack systems of different orientations may have a significant
effect on damage evolution. If y is close to 90 , this intra-mode interaction is higher
because of the close proximity of the 90 and y-crack planes. Ply cracks usually
initiate in the 90 -plies first and then in the 60 -layers. Thus, initial simulation
assumes only 90 -cracks; whereas after the initiation of 60 -cracks, a multi-mode
scenario is used in FE modeling. The 60 -cracks influence the damage progression
in the 90 -layer. Model predictions for y = 60 for different values of m, n, and p are
shown in Figure 6.35(a)–(c) for the 90 -, 60 -, and +60 -layers, respectively. For p
= 2, the model predicts that cracks in the 60 - and +60 -layers will initiate earlier
than in the 90 -layer. If y < 45 , this intra-mode interaction is not appreciable, and
for this case the damage evolution in the 90 -layer may not be affected at all.

6.7 Damage evolution under cyclic loading

While the fatigue process in composite materials is treated in the next chapter, here
we shall describe a modeling approach for the evolution of transverse cracking in
cross-ply laminates under cyclic axial tension. This case because of its simplicity
of geometry serves as a good illustration of the fundamental ideas in the fatigue of
composites. The reader is urged to refer to two papers [53, 132] for full details of
the treatment described below.
(a)

(b)

(c)

Figure 6.35. Evolution of crack density in a [0m/90n/ 60p]s laminate for varying ply
thicknesses in (a) 90 -layer, (b) 60 -layer, and (c) +60 -layer. Reprinted, with kind
permission, from Int J Solids Struct, Vol. 47, C. V. Singh and R. Talreja, Evolution
of ply cracks in multidirectional composite laminates, pp. 1338–49, copyright
Elsevier (2010).
6.7 Damage evolution under cyclic loading 225

The guiding principle in any fatigue analysis must be to address the question:
what is the mechanism of irreversibility that causes the accumulation of damage
from one load cycle to another? The common energy-dissipating mechanisms are
plasticity, friction, and surface formation. For composite laminates that are
modeled as layered elastic solids plasticity is not admissible. Frictional processes
within the volume of such composites are possible between crack surfaces if the
surfaces are in contact. Finally, new surface formation without plasticity is pos-
sible by brittle fracture.
As a case for illustration we shall consider a cross-ply laminate with ply cracks
in the central 90 -plies that formed under the first application of an axial tensile
load. The problem posed is: if the load is removed and repeatedly reapplied to the
previous maximum value, when would new cracks form between the pre-existing
cracks? To begin the analysis we note that for the cracked cross-ply laminate a
solution of good accuracy for the stress field in the region between cracks is
available (e.g., [101]). This solution is valid for perfectly and linearly elastic
(no plasticity) laminates. Also, the solution does not apply to partial cracks, i.e.,
cracks that are not fully extended in the thickness and width directions of the 90 -
plies. Thus to use this solution we must retain the symmetry and periodicity of the
cracks assumed in obtaining the solution. This condition eliminates analysis of the
cyclic growth of transverse cracks from partial to full extent. Obviously, a numer-
ical analysis of this case is possible, but that would not provide the analytic fatigue
damage model we intend to develop.
With the analytical stress solution to the cracked cross-ply laminate at hand
we see that unless some irreversible mechanism is included no change in the
crack density can be predicted since any repetition of load in an elastic solid
cannot change the stress field. Therefore, in order to have an analytical stress
solution, i.e., keeping the symmetry of laminate geometry and periodicity of
cracks, and to incorporate irreversibility, a novel idea was proposed in [132].
According to that, all irreversibility leading to damage accumulation is lumped
into delamination surfaces emanating from the transverse crack fronts. Figure
6.36(a) illustrates the resulting model geometry of the cracked laminate while
Figure 6.36(b) shows the repeating unit cell. As shown, the pre-existing trans-
verse cracks are spaced at distance 2a and the delamination on either side of the
cracks is of distance d. The idea behind the model is that the delamination grows
under applied cyclic loading, the same way as a crack does, imparting changes to
the stress field in the region between transverse ply cracks. In this way the model
captures cycle-dependent irreversibility, thereby allowing the fatigue-induced
multiplication of transverse ply cracks to be modeled. Although the formation
of new cracks between pre-existing cracks can be modeled by different criteria, in
[132] a maximum stress criterion for cracking is used, supported by previous
work [133]. It can, however, be shown that if the delamination surfaces are
traction free, then the maximum axial stress between the ply cracks goes down
as the delamination length d increases. This suggests that the irreversibility
captured in such delamination growth is inadequate for the purpose. As argued
226 Damage progression

(a) (b)
X

Nxx
d

Z
X

Z
Region I
Y
Region II

0⬚ 90⬚ t1 t2
90⬚ 0⬚
h

Figure 6.36. Schematic of a cracked cross-ply laminate: (a) uniformly distributed transverse
matrix cracks in 90 -plies with the associated delamination of 0 /90 interface; (b) a unit
cell between two matrix cracks with delaminated region (region I) and perfectly bonded
region (region II).

in [132], the delamination surfaces are indeed under compressive stress [134],
making it plausible that frictional contact exists between those surfaces. The
ensuing frictional sliding was modeled by an interfacial shear stress, which
provided the needed increase in the axial stress for crack formation.
A description of the model now follows.
Referring to Figure 6.36(a) and (b), the stress analysis is performed by a
variational approach along the lines in [101], conducting the minimization of
complementary energy separately for region I and region II. First, an admissible
stress system in the x–z plane of region II is expressed as
LðmÞ 0ðmÞ ðmÞ
sij ¼ sij þ sij ; ð6:86Þ
LðmÞ 0ðmÞ
where sij and sij are the stress components in the cracked laminate and in the
ðmÞ
virgin laminate, respectively, and sij are the perturbations; m = 1 and 2 indicate

the 90 - and 0 -plies, respectively. The axial perturbation stresses in the plies are
assumed to have the following form

sð1Þ ð2Þ
xx ¼ s1 f1 ð xÞ; sxx  s2 ½f2 ð xÞ þ Að xÞ
z; ð6:87Þ
0ðmÞ
where sxx¼ sm has been used; f1(x), f2(x), and A(x) are unknown functions.
Applying the equilibrium in the x direction and the interface iso-strain condition
at z = t1, f2(x) and A(x) in Eq. (6.87) are eliminated, so the axial perturbation
stresses can be expressed by the only unknown function, f1(x). After integrating
equilibrium equations and using Eq. (6.87), all perturbation stress components
6.7 Damage evolution under cyclic loading 227

are expressed as functions of f1(x) by applying interface continuity conditions at


z = t1 and traction-free boundary conditions at z = h. The admissible stress
system for a cracked laminate is then established based on the only unknown
function, f1(x). The corresponding complementary energy functional for linear
elastic materials in a volume V, with only traction boundary conditions in region
II, can be written as [101]:
ð ð
1 0ðmÞ 0ðmÞ 1 ðmÞ ðmÞ
Uc ¼ Uc0 þ Uc0 ¼ sijkl sij skl dV þ sijkl sij skl dV; ð6:88Þ
2 V 2 V

where Sijkl are the components of the compliance tensor. Since the virgin laminate
stresses are constant, Uc0 is not of importance to the analysis. Substituting all
perturbation stress components into Eq. (6.88) gives
ð ðadÞ h i
2 2
Uc0 ¼ ðs1 Þ2 t1 C00 f21 þ t31 C11 ðf01 Þ þ t51 C22 ðf001 Þ þ t31 C02 f1 f001 dx; ð6:89Þ
ðadÞ

where Cij are constants determined by the elastic constants and thickness of each
layer, a is half the crack spacing, and d is the delamination length. Minimizing Uc0
after introducing the nondimensional variable x = x/t1, the following Euler–
Lagrange differential equation in f1 is obtained

d4 f1 d2 f1
4
þp þ qf1 ¼ 0; ð6:90Þ
dx dx2
where p and q are constants determined by Cij. Dependent on the material elastic
property and geometry of the given laminate, two solutions for Eq. (6.90) exist
8 2
9
< A1 coshðaxÞ cosðbxÞ þ A2 sinhðaxÞ sinðbxÞ; p  q <0 =
f1 ðxÞ ¼ 2
4 ð6:91Þ
: A1 coshðaxÞ þ A2 coshðbxÞ; p  q > 0 ;
4

The axial normal stress of interest at the mid-plane in the 90 plies is obtained as
sxx ð0; zÞ ¼ s1 ð1  f1 ð0ÞÞ; t1 <z<t1 : ð6:92Þ
The two constants, A1 and A2, are found using the traction continuity conditions at
x = (a – d) if the stress in region I is known.
To obtain the stress state in region I, a very similar variational approach is
applied. For an admissible stress system, the perturbation stresses are assumed
first as

sð1Þ ð2Þ 
xx ¼ s1 c1 ðxÞ; sxx ¼ s2 ½c2 ð xÞ þ A
z: ð6:93Þ
A cubic variation of shear stress along the interface (z = t1) is enforced to account
for the effect of frictional sliding along delamination
t
sð1Þ
xx ðx; t1 Þ ¼ ðx  aÞ½x  ða  dÞ2 ; ð6:94Þ
a3
228 Damage progression

5.25 × 107
sxx(x = 0, z = 0) (Pa) 5 × 107

4.75 × 107

4.5 × 107

4.25 × 107

0.0002 0.0004 0.0006


3.75 × 107
d (m)

Figure 6.37. Typical variation of the axial normal stress with delamination at a fixed crack density.

where t is an unknown. After integrating the equilibrium equation, and applying


continuity of stress across the interface at z = t1, as well as traction-free boundary
conditions at the crack surfaces, the admissible stress system is built up as
functions of the only unknown, t. By minimizing the corresponding complemen-
tary energy functional,
dUc
¼ 0; ð6:95Þ
dt
the only unknown t is solved, so the stress field in region I is obtained. By applying
the traction continuity conditions at the boundary between region I and region II,
A1 and A2 in Eq. (6.91) are determined. Substituting A1 and A2 into Eq. (6.92),
finally, the axial normal stress at the mid-plane in 90 plies is obtained for the
given unit cell with specified crack spacing and delamination length. A typical
variation of the axial normal stress with delamination at a fixed crack density is
shown in Figure 6.37.
The following power-law relation is assumed to describe the growth of delami-
nation under cyclic loading
 m
dl t
¼B  ; ð6:96Þ
dN l
where t ¼ ðtmax  tmin Þ=smax and l ¼ l=t1 , and where l is the delamination
length, denoted as d in the above stress analysis.
Integration of Eq. (6.96) yields the relationship between delamination length
and the number of cycles, N. From Figure 6.37 we see that after the initial
drop the axial normal stress increases as delamination grows. When the
maximum stress criterion sxx(x = 0, t1 < z < t1) = sc is satisfied, a new
crack forms midway between the pre-existing cracks, and the corresponding
crack spacing (or crack density) is updated. In this way, the multiplication of
matrix crack under fatigue loading is actually controlled by the growth of
delamination, and therefore determined by the number of cycles through
integration of Eq. (6.96). The damage evolution as a function of the number
6.8 Summary 229

2.0
Model prediction
Experiment

1.8
Crack density (/mm)

1.6

1.4

1.2

1.0
0 200 000 400 000 600 000 800 000 1000 000
Number of cycles

Figure 6.38. Transverse crack evolution with cycles for a [0/902]s carbon/epoxy laminate at a
maximum stress of 482.633 MPa and a stress ratio of 0.1.

of cycles in cross-ply laminates is finally quantitatively described in the model.


Figure 6.38 shows the variation of crack density as a function of the number of
cycles under a given cyclic tension.

6.8 Summary

Initiation of cracks within the plies of a laminate, and their growth and multipli-
cation, are part of the field of damage evolution in composites that constitutes a
key element in the performance assessment of structures made of these materials.
This chapter has focused on various stress and failure analysis methods associ-
ated with the prediction of initiation and progression of ply cracks. Since cracks
form from material defects that can be considered random in their size and
spatial distributions, statistical considerations have been included in the
analyses.
For the formation of cracks the criteria used are based either on strength or on
the energy associated with fracture. Both criteria have been treated and compared.
Experimental data, wherever available, have been used to assess the predictions.
Most of the damage evolution work in the literature has been for transverse
cracking in cross-ply laminates. While this cracking mode has been amply dealt
with, more recent work on oblique cracks, i.e., cracks in off-axis plies of multi-
directional laminates, has also been treated.
While the fatigue of composite laminates is the focus of a separate chapter,
where a broad treatment of the subject is given, a section on ply cracking under
cyclic loading has been included here to illustrate how damage accumulation
under repeated loads is modeled.
230 Damage progression

References

1. A. S. D. Wang, Fracture mechanics of sublaminate cracks in composite materials.


In Composites Technology Review. (Philadelphia, PA: ASTM, 1984), pp. 45–62.
2. J. M. Masters and K. L. Reifsnider, An investigation of cumulative damage develop-
ment in quasi-isotropic graphite/epoxy laminates. In Damage in Composite Materials,
ASTM STP 775, ed. K. L. Reifsnider. (Philadelphia, PA: ASTM, 1982), pp. 40–62.
3. K. L. Reifsnider, ed. Damage in Composite Materials, ASTM STP 775, ed. K. L.
Reifsnider. (Philadelphia, PA: ASTM, 1982).
4. J. E. Masters, ed. Damage Detection in Composite Materials, ASTM STP 1128.
(Philadelphia, PA: ASTM, 1992).
5. J. A. Nairn, S. F. Hu, and J. S. Bark, A critical-evaluation of theories for predicting
microcracking in composite laminates. J Materials Sci, 28:18 (1993), 5099–111.
6. D. O. Stalnaker and W. W. Stinchcomb, Load history–edge damage studies in
two quasi-isotropic graphite epoxy laminates. In Composite Materials: Testing and
Design (Proc. 5th Conference), ASTM STP 674. (Philadelphia, PA: ASTM, 1979), pp.
620–41.
7. J. E. Masters, An experimental investigation of cumulative damage development in
graphite epoxy laminates. Ph.D. dissertation, Materials Engineering Science, Depart-
ment, Virginia Polytechnic Institute, Blacksburg, VA, 1980, p. 40–62.
8. A. L. Highsmith and K. L. Reifsnider, Stiffness-reduction mechanisms in composite
laminates. In Damage in Composite Materials, ASTM STP 775, ed. K. L. Reifsnider.
(Philadelphia, PA: ASTM, 1982), pp. 103–17.
9. K. L. Reifsnider and R. Jamison, Fracture of fatigue-loaded composite laminates. Int
J Fatigue, 4:4 (1982), 187–97.
10. K. L. Reifsnider and A. Talug, Analysis of fatigue damage in composite laminates. Int
J Fatigue, 2:1 (1980), 3–11.
11. E. Adolfsson and P. Gudmundson, Matrix crack initiation and progression in composite
laminates subjected to bending and extension. Int J Solids Struct, 36:21 (1999), 3131–69.
12. B. F. Sorensen and R. Talreja, Analysis of damage in a ceramic matrix composite. Int
J Damage Mech, 2:3 (1993), 246–71.
13. B. F. Sorensen, R. Talreja, and O. T. Sorensen, Micromechanical analysis of damage
mechanisms in ceramic-matrix composites during mechanical and thermal cycling.
Composites, 24:2 (1993), 129–40.
14. T. J. Dunyak, W. W. Stinchcomb, and K. L. Reifsnider, Examination of selected NDE
techniques for ceramic composite components. In Damage Detection in Composite
Materials, ASTM STP 1128, ed. J. E. Masters. (Philadelphia, PA: ASTM, 1992), pp.
3–24.
15. K. V. Steiner, Defect classifications in composites using ultrasonic nondestructive
evaluation techniques. In Damage Detection in Composite Materials, ASTM STP 1128,
ed. J. E. Masters. (Philadelphia, PA: ASTM, 1992), pp. 72–84.
16. K. V. Steiner, R. F. Eduljee, X. Huang, and J. W. Gillespie, Ultrasonic NDE tech-
niques for the evaluation of matrix cracking in composite laminates. Compos Sci
Technol, 53:2 (1995), 193–198.
17. G. E. Maddux and G. P. Sendeckyj, Holographic techniques for defect detection in
composite materials. In Nondestructive Evaluation and Flaw Criticality for Composite
Materials, ASTM STP 696. (Philadelphia, PA: ASTM, 1979), pp. 26–44.
References 231

18. G. P. Sendeckyj, G. E. Maddux, and E. Porter, Damage documentation in composites


by stereo radiography. In Damage in Composite Materials, ASTM STP 775, ed. K. L.
Reifsnider. (Philadelphia, PA: ASTM, 1982), pp. 16–26.
19. R. D. Jamison, K. Schulte, K. L. Reifsnider, and W. W. Stinchcomb, Characterization
and analysis of damage mechanisms in tension–tension fatigue of graphite/epoxy
laminates. In Effects of Defects in Composite Materials, ASTM STP 836. (Philadelphia,
PA: ASTM, 1984), pp. 21–55.
20. C. Bathias and A. Cagnasso, Application of x-ray tomography to the non-destructive
testing of high-performance polymer composites. In Damage Detection in Composite Mater-
ials, ASTM STP 1128, ed. J. E. Masters. (Philadelphia, PA: ASTM, 1992), pp. 35–54.
21. K. Maslov, R. Y. Kim, V. K. Kinra, and N. J. Pagano, A new technique for the
ultrasonic detection of internal transverse cracks in carbon-fibre/bismaleimide compos-
ite laminates. Compos Sci Technol, 60:12–13 (2000), 2185–90.
22. V. K. Kinra, A. S. Ganpatye, and K. Maslov, Ultrasonic ply-by-ply detection of matrix
cracks in laminated composites. J Nondestr Eval, 25:1 (2006), 39–51.
23. Y. Zou, L. Tong, and G. P. Steven, Vibration-based model-dependent damage (dela-
mination) identification and health monitoring for composite structures – a review.
J Sound Vib, 230:2 (2000), 357–78.
24. S. K. Seth, et al., Damage detection in composite materials using lamb wave methods.
Smart Mater Struct, 11:2 (2002), 269.
25. R. Talreja, Transverse cracking and stiffness reduction in composite laminates.
J Compos Mater, 19:4 (1985), 355–75.
26. J. Varna, N. V. Akshantala, and R. Talreja, Crack opening displacement and the associ-
ated response of laminates with varying constraints. Int J Damage Mech, 8 (1999), 174–93.
27. J. Varna, R. Joffe, N. V. Akshantala, and R. Talreja, Damage in composite laminates
with off-axis plies. Compos Sci Technol, 59:14 (1999), 2139–47.
28. D. G. Katerelos, L. N. McCartney, and C. Galiotis, Local strain re-distribution and
stiffness degradation in cross-ply polymer composites under tension. Acta Mater, 53:12
(2005), 3335–43.
29. D. T. G. Katerelos, P. Lundmark, J. Varna, and C. Galiotis, Analysis of matrix
cracking in GFRP laminates using Raman spectroscopy. Compos Sci Technol, 67:9
(2007), 1946–54.
30. D. G. Katerelos, M. Kashtalyan, C. Soutis, and C. Galiotis, Matrix cracking in
polymeric composite laminates: Modelling and experiments. Compos Sci Technol,
68:12 (2008), 2310–17.
31. D. G. Katerelos, L. N. McCartney, and C. Galiotis, Effect of off-axis matrix cracking
on stiffness of symmetric angle-ply composite laminates. Int J Fract, 139:3–4 (2006),
529–36.
32. D. T. G. Katerelos, J. Varna, and C. Galiotis, Energy criterion for modelling
damage evolution in cross-ply composite laminates. Compos Sci Technol, 68:12 (2008),
2318–24.
33. J. A. Nairn, Matrix microcracking in composites. In Polymer Matrix Composites, ed. R.
Talreja and J. A. E. Manson. (Amsterdam: Elsevier Science, 2000), pp. 403–32.
34. J. Tong, F. J. Guild, S. L. Ogin, and P. A. Smith, On matrix crack growth in quasi-isotropic
laminates – I. Experimental investigation. Compos Sci Technol, 57:11 (1997), 1527–35.
35. L. E. Crocker, S. L. Ogin, P. A. Smith, and P. S. Hill, Intra-laminar fracture in angle-
ply laminates. Compos A, 28:9–10 (1997), 839–46.
232 Damage progression

36. N. Balhi, et al., Intra-laminar cracking in CFRP laminates: observations and model-
ling. J Mater Sci, 41:20 (2006), 6599–609.
37. K. W. Garrett and J. E. Bailey, Effect of resin failure strain on tensile properties of glass
fiber-reinforced polyester cross-ply laminates. J Mater Sci, 12:11 (1977), 2189–94.
38. K. W. Garrett and J. E. Bailey, Multiple transverse fracture in 90 cross-ply laminates
of a glass fibre-reinforced polyester. J Mater Sci, 12:1 (1977), 157–68.
39. A. Parvizi and J. E. Bailey, Multiple transverse cracking in glass-fiber epoxy cross-ply
laminates. J Mater Sci, 13:10 (1978), 2131–6.
40. A. Parvizi, K. W. Garrett, and J. E. Bailey, Constrained cracking in glass fibre-
reinforced epoxy cross-ply laminates. J Mater Sci, 13:1 (1978), 195–201.
41. J. E. Bailey, P. T. Curtis, and A. Parvizi, On the transverse cracking and longitudinal
splitting behavior of glass and carbon-fiber reinforced epoxy cross ply laminates and
the effect of Poisson and thermally generated strain. Proc R Soc London A, 366:1727
(1979), 599–623.
42. M. G. Bader, J. E. Bailey, P. T. Curtis, and A. Parvizi, eds., The mechanisms of
initiation and development of damage in multi-axial fibre-reinforced plastic laminates.
In Proceedings of the Third International Conference on the Mechancial Behaviour in
Materials (ICM3), Cambridge, UK, Vol. 3. (1979), pp. 227–39.
43. J. E. Bailey and A. Parvizi, On fiber debonding effects and the mechanism of trans-
verse-ply failure in cross-ply laminates of glass fiber-thermoset composites. J Mater Sci,
16:3 (1981), 649–59.
44. F. R. Jones, A. R. Wheatley, and J. E. Bailey, The effect of thermal strains on the
microcracking and stress corrosion behaviour of GRP. In Composite Structures – 1st
International Conference, ed. I. H. Marshall. (Barking, UK: Applied Science Publishers,
1981), pp. 415–29.
45. D. L. Flaggs and M. H. Kural, Experimental-determination of the insitu transverse
lamina strength in graphite epoxy laminates. J Compos Mater, 16 (1982), 103–16.
46. J. A. Nairn and S. Hu, Micromechanics of damage: a case study of matrix micro-
cracking. In Damage Mechanics of Composite Materials, ed. R. Talreja. (Amsterdam:
Elsevier, 1994), pp. 187–243.
47. R. Talreja, Transverse cracking and stiffness reduction in composite laminates.
J Compos Mater, 19:4 (1985), 355–75.
48. J. A. Nairn and S. F. Hu, The formation and effect of outer-ply microcracks in cross-
ply laminates – a variational approach. Eng Fract Mech, 41:2 (1992), 203–21.
49. J. Tong, F. J. Guild, S. L. Ogin, and P. A. Smith, On matrix crack growth in quasi-isotropic
laminates – II. Finite element analysis. Compos Sci Technol, 57:11 (1997), 1537–45.
50. T. Yokozeki, T. Aoki, T. Ogasawara, and T. Ishikawa, Effects of layup angle and ply
thickness on matrix crack interaction in contiguous plies of composite laminates.
Compos A, 36:9 (2005), 1229–35.
51. T. Yokozeki, T. Aoki, and T. Ishikawa, Consecutive matrix cracking in contiguous
plies of composite laminates. Int J Solids Struct, 42:9–10 (2005), 2785–802.
52. J. E. Masters and K. L. Reifsnider, An investigation of cumulative damage deve-
lopment in quasi-isotropic graphite/epoxy laminates. In Damage in Composite
Materials, ASTM STP 775, ed. K. L. Reifsnider. (Philadelphia, PA: ASTM, 1982),
pp. 40–62.
53. N. V. Akshantala and R. Talreja, A micromechanics based model for predicting fatigue
life of composite laminates. Mater Sci Eng A, 285:1–2 (2000), 303–13.
References 233

54. K. W. Garrett and J. E. Bailey, Multiple transverse fracture in 90 degrees cross-ply


laminates of a glass fiber-reinforced polyester. J Materials Sci, 12:1 (1977), 157–68.
55. J. M. Berthelot, Transverse cracking and delamination in cross-ply glass-fiber and
carbon-fiber reinforced plastic laminates: static and fatigue loading. Appl Mech Rev,
56 (2003), 111–47.
56. S. E. Groves, et al., An experimental and analytical treatment of matrix cracking in
cross-ply laminates. Exp Mech, 27:1 (1987), 73–9.
57. S. F. Hu, J. S. Bark, and J. A. Nairn, On the phenomenon of curved microcracks in
[(S)/90(n)](s) laminates – their shapes, initiation angles and locations. Compos Sci
Technol, 47:4 (1993), 321–9.
58. P. Lundmark and J. Varna, Damage evolution and characterisation of crack types in
CF/EP laminates loaded at low temperatures. Eng Fract Mech, 75:9 (2008), 2631–41.
59. H. L. McManus, D. E. Bowles, and S. S. Tompkins, Prediction of thermal cycling
induced matrix cracking. J Reinf Plast Compos, 15:2 (1996), 124–40.
60. H. L. McManus and J. R. Maddocks, On microcracking in composite laminates under
thermal and mechanical loading. Polym Compos, 4:5 (1996), 305–14.
61. C. H. Park and H. L. McManus, Thermally induced damage in composite laminates:
predictive methodology and experimental investigation. Compos Sci Technol, 56:10
(1996), 1209–17.
62. T. G. Reynolds and H. L. McManus, Understanding and accelerating environmentally-
induced degradation and microcracking. In Proceedings of 39th AIAA Structures, Struc-
tural Dynamics and Materials Conference. (Reston, VA: AIAA, 1998, pp. 2103–16.
63. D. S. Adams, D. E. Bowles, and C. T. Herakovich, Thermally induced transverse
cracking in graphite-epoxy cross-ply laminates. J Reinf Plast Compos, 5:3 (1986),
152–69.
64. C. T. Herakovich and M. W. Hyer, Damage-induced property changes in composites
subjected to cyclic thermal loading. Eng Fract Mech, 25:5–6 (1986), 779–91.
65. R. G. Spain, Thermal microcracking of carbon fibre/resin composites. Compos, 2:1
(1971), 33–7.
66. J. A. Lavoie and E. Adolfsson, Stitch cracks in constraint plies adjacent to a cracked
ply. J Compos Mater, 35:23 (2001), 2077–97.
67. X. Huang, J. W. Gillespie, and R. F. Eduljee, Effect of temperature on the transverse
cracking behavior of cross-ply composite laminates. Compos B, 28:4 (1997), 419–24.
68. M. H. Han and J. A. Nairn, Hygrothermal aging of polyimide matrix composite
laminates. Compos A, 34:10 (2003), 979–86.
69. J. E. Lundgren and P. Gudmundson, Moisture absorption in glass-fibre/epoxy lamin-
ates with transverse matrix cracks. Compos Sci Technol, 59:13 (1999), 1983–91.
70. E. Vauthier, J. C. Abry, T. Bailliez, and A. Chateauminois, Interactions between
hygrothermal ageing and fatigue damage in unidirectional glass/epoxy composites.
Compos Sci Technol, 58:5 (1998), 687–92.
71. A. Tounsi, K. H. Amara, and E. A. Adda-Bedia, Analysis of transverse cracking and
stiffness loss in cross-ply laminates with hygrothermal conditions. Comput Mater Sci,
32:2 (2005), 167–74.
72. P. A. Smith and S. L. Ogin, On transverse matrix cracking in cross-ply laminates loaded
in simple bending. Compos A, 30:8 (1999), 1003–8.
73. P. A. Smith and S. L. Ogin, Characterization and modelling of matrix cracking in a (0/90)
2s GFRP laminate loaded in flexure. Proc R Soc London A, 456:2003 (2000), 2755–70.
234 Damage progression

74. S.-R. Kim and J. A. Nairn, Fracture mechanics analysis of coating/substrate systems: part
I: analysis of tensile and bending experiments. Eng Fract Mech, 65:5 (2000), 573–93.
75. S.-R. Kim and J. A. Nairn, Fracture mechanics analysis of coating/substrate systems:
part II: experiments in bending. Eng Fract Mech, 65:5 (2000), 595–607.
76. L. N. McCartney and C. Pierse, Stress Transfer Mechanics for Multiple-ply Laminates
Subject to Bending. NPL Report CMMT(A) 55, February 1997.
77. L. N. McCartney and C. Pierse, Stress transfer mechanics for multiple ply laminates for
axial loading and bending. Proc ICCM-11, 5 (1997), 662–71.
78. S. A. Salpekar and T. K. Obrien, Analysis of matrix cracking and local delamination in
(0/y/–y)s graphite-epoxy laminates under tensile load. J Compos Tech Res, 15:2 (1993),
95–100.
79. P. Johnson and F. K. Chang, Characterization of matrix crack-induced laminate
failure – part I: experiments. J Compos Mater, 35:22 (2001), 2009–35.
80. P. Johnson and F. K. Chang, Characterization of matrix crack-induced laminate
failure – part II: analysis and verifications. J Compos Mater, 35:22 (2001), 2037–74.
81. A. Parvizi, K. W. Garrett, and J. E. Bailey, Constrained cracking in glass fiber-
reinforced epoxy cross-ply laminates. J Mater Sci, 13:1 (1978), 195–201.
82. H. T. Hahn and S. W. Tsai, Behavior of composite laminates after initial failures.
J Compos Mater, 8 (1974), 288–305.
83. H. Fukunaga, T. W. Chou, P. W. M. Peters, and K. Schulte, Probabilistic failure
strength analysis of graphite epoxy cross-ply laminates. J Compos Mater, 18:4 (1984),
339–56.
84. H. Fukunaga, T. W. Chou, K. Schulte, and P. W. M. Peters, Probabilistic initial failure
strength of hybrid and non-hybrid laminates. J Mater Sci, 19:11 (1984), 3546–53.
85. P. W. M. Peters, The strength distribution of 90-deg plies in 0/90/0 graphite-epoxy
laminates. J Compos Mater, 18 (1984), 545–56.
86. P. W. M. Peters, The fiber/matrix bond strength of CFRP deduced from the strength
transverse to the fibers. J Adhes, 53 (1995), 79–101.
87. P. W. Manders, T. W. Chou, F. R. Jones, and J. W. Rock, Statistical analysis of multiple
fracture in [0/90/0] glass fiber/epoxy resin laminates. J Mater Sci, 19 (1983), 2876–89.
88. F. W. Crossman, W. J. Warren, A. S. D. Wang, and J. G. E. Law, Initiation and
growth of transverse cracks and edge delamination in composite laminates: part 2.
Experimental correlation. J Compos Mater Suppl, 14 (1980), 89–108.
89. P. W. M. Peters, Strength distribution of 90 degree plies in 0/90/0 graphite-epoxy
laminates. J Compos Mater, 18:6 (1984), 545–56.
90. N. Takeda and S. Ogihara, In-situ observation and probabilistic prediction of micros-
copic failure processes in CFRP cross-ply laminates. Compos Sci Technol, 52:2 (1994),
183–95.
91. J. Aveston and A. Kelly, Theory of multiple fracture of fibrous composites. J Mater Sci,
8:3 (1973), 352–62.
92. P. W. Manders, T. W. Chou, F. R. Jones, and J. W. Rock, Statistical analysis of
multiple fracture in 0 /90 /0 glass fibre/epoxy resin laminates. J Mater Sci, 18:10
(1983), 2876–89.
93. P. S. Steif, Parabolic shear lag analysis of a [0/90]s laminate. In Transverse Ply Crack
Growth and Associated Stiffness Reduction during the Fatigue of a Simple Cross-ply
Laminate, ed. S. L. Ogin, P. A. Smith and P. W. R. Beaumont, Report CUED/C/
MATS/TR 105, Cambridge University, September 1984, pp. 40–1.
References 235

94. J. M. Berthelot, P. Leblond, A. El Mahi, and J. F. Le Corre, Transverse cracking of


cross-ply laminates: part 1. Analysis. Compos A, 27:10 (1996), 989–1001.
95. J. M. Berthelot, A. El Mahi, and P. Leblond, Transverse cracking of cross-
ply laminates: part 2. Progressive widthwise cracking. Compos A, 27:10 (1996), 1003–10.
96. J. M. Berthelot, Analysis of the transverse cracking of cross-ply laminates: a general-
ized approach. J Compos Mater, 31:18 (1997), 1780–805.
97. Z. Hashin, Finite thermoelastic fracture criterion with application to laminate
cracking analysis. J Mech Phys Solids, 44:7 (1996), 1129–45.
98. V. Vinogradov and Z. Hashin, Probabilistic energy based model for prediction of
transverse cracking in cross-ply laminates. Int J Solids Struct, 42:2 (2005), 365–92.
99. J. A. Nairn, The strain-energy release rate of composite microcracking – a variational
approach. J Compos Mater, 23:11 (1989), 1106–29.
100. N. Laws and G. J. Dvorak, Progressive transverse cracking in composite laminates.
J Compos Mater, 22:10 (1988), 900–16.
101. Z. Hashin, Analysis of cracked laminates: A variational approach. Mech Mater, 4:2
(1985), 121–36.
102. Z. Hashin, Micromechanics of cracking in composite materials. In Continuum Models
and Discrete Systems. Proc 9th Int Symposium, ed. E. Inan and K. Markov. (Singapore:
World Scientific Publishing, 1998), pp. 702–9.
103. Z. Hashin, Thermal-expansion coefficients of cracked laminates. Compos Sci Technol,
31:4 (1988), 247–60.
104. L. N. McCartney, Predicting transverse crack formation in cross-ply laminates.
Compos Sci Technol, 58:7 (1998), 1069–81.
105. L. N. McCartney, Energy-based prediction of progressive ply cracking and strength of
general symmetric laminates using an homogenisation method. Compos A, 36:2 (2005),
119–28.
106. L. N. McCartney, Energy-based prediction of failure in general symmetric laminates.
Eng Fract Mech, 72:6 (2005), 909–30.
107. L. N. McCartney, Physically based damage models for laminated composites. Proc
Inst Mech Engineers L – J Mater Design Appl, 217:L3 (2003), 163–99.
108. L. N. McCartney, Prediction of ply crack formation and failure in laminates. Compos
Sci Technol, 62:12–13 (2002), 1619–31.
109. L. N. McCartney, Model to predict effects of triaxial loading on ply cracking in
general symmetric laminates. Compos Sci Technol, 60:12–13 (2000), 2255–79.
110. A. S. D. Wang and F. W. Crossman, Initiation and growth of transverse cracks and
edge delamination in composite laminates part 1. An energy method. J Compos Mater,
14:Suppl. (1980), 71–87.
111. R. Joffe and J. Varna, Damage evolution modeling in multidirectional laminates and
the resulting nonlinear response. In Proceeedings of International Conference on Com-
posite Materials 12, Paris, France, July 5–9, 1999.
112. J. Varna, R. Joffe, and R. Talreja, A synergistic damage-mechanics analysis of
transverse cracking in [ y/904]s laminates. Compos Sci Technol, 61:5 (2001), 657–65.
113. R. Joffe, A. Krasnikovs, and J. Varna, COD-based simulation of transverse cracking
and stiffness reduction in [s/90n]s laminates. Compos Sci Technol, 61:5 (2001), 637–56.
114. J. Varna, R. Joffe, and R. Talreja, Mixed micromechanics and continuum damage
mechanics approach to transverse cracking in [S,90(n)](s) laminates. Mech Compos
Mater, 37:2 (2001), 115–26.
236 Damage progression

115. E. Adolfsson and P. Gudmundson, Thermoelastic properties in combined bending


and extension of thin composite laminates with transverse matrix cracks. Int J Solids
Struct, 34:16 (1997), 2035–60.
116. J. Qu and K. Hoiseth, Evolution of transverse matrix cracking in cross-ply laminates.
Fatigue Frac Eng M, 21:4 (1998), 451–64.
117. J. Varna and L. A. Berglund, Multiple transverse cracking and stiffness reduction in
cross-ply laminates. J Compos Tech Res, 13:2 (1991), 97–106.
118. C. Baxevanakis, D. Jeulin, and J. Renard, Fracture statistics of a unidirectional
composite. Int J Fract, 73:2 (1995), 149–81.
119. V. V. Silberschmidt, Matrix cracking in cross-ply laminates: effect of randomness.
Compos A, 36:2 (special issue) (2005), 129–35.
120. V. V. Silberschmidt, Effect of micro-randomness on macroscopic properties and
fracture of laminates. J Mater Sci, 41:20 (2006), 6768–76.
121. J. M. Berthelot and J. F. Le Corre, Statistical analysis of the progression of transverse
cracking and delamination in cross-ply laminates. Compos Sci Technol, 60:14 (2000),
2659–69.
122. F. W. Crossman, W. J. Warren, A. S. D. Wang, and G. E. Law, Initiation and growth
of transverse cracks and edge delamination in composite laminates, part 2: experi-
mental correlation. J Compos Mater, 14:Suppl. (1980), 88–108.
123. F. W. Crossman and A. S. D. Wang, The dependence of transverse cracking and delamina-
tion on ply thickness in graphite/epoxy laminates. In Damage in Composite Materials,
ASTM STP 775, ed. K. L. Reifsnider. (Philadelphia, PA: ASTM, 1982), pp. 118–39.
124. A. S. D. Wang, P. C. Chou, and S. C. Lei, Stochastic model for the growth of matrix
cracks in composite laminates. J Compos Mater, 18:3 (1984), 239–54.
125. A. S. D. Wang, N. N. Kishore, and C. A. Li, Crack development in graphite-epoxy
laminates under uniaxial tension. Compos Sci Technol, 24:1 (1985), 1–31.
126. P. C. Chou, A. S. D. Wang, and H. Miller, Cumulative Damage Model for Advanced
Composite Eaterials, Final report by Dyna East Corp, Philadelphia, PA, September 1982.
127. S. L. Liu and J. A. Nairn, The formation and propagation of matrix microcracks in
cross-ply laminates during static loading. J Reinf Plast Compos, 11:2 (1992), 158–78.
128. T. Yokozeki, T. Ogasawara, and T. Ishikawa, Evaluation of gas leakage through
composite laminates with multilayer matrix cracks: cracking angle effects. Compos Sci
Technol, 66:15 (2006), 2815–24.
129. C. V. Singh and R. Talreja, Evolution of ply cracks in multidirectional composite
laminates. Int J Solids Struct, 47:10 (2010), 1338–49.
130. M. Kashtalyan and C. Soutis, Stiffness and fracture analysis of laminated composites
with off-axis ply matrix cracking. Compos A, 38:4 (2007), 1262–9.
131. C. V. Singh, Multiscale modeling of damage in multidirectional composite laminates.
Ph.D. thesis, Department of Aerospace Engineering, Texas A&M University, College
Station, TX (2008).
132. N. V. Akshantala and R. Talreja, A mechanistic model for fatigue damage evolution
in composite laminates. Mech Mater, 29:2 (1998), 123–40.
133. R. Talreja and N. V. Akshantala, An inadequacy in a common micromechanics
approach to analysis of damage evolution in composites. Int J Damage Mech, 7:3
(1998), 238–49.
134. T. W. Kim, H. J. Kim, and S. Im, Delamination crack originating from transverse
cracking in cross-ply laminates under extension. Int J Solids Struct, 27:15 (1991), 1925–41.
7 Damage mechanisms
and fatigue-life diagrams

7.1 Introduction

The fatigue of composite materials presents a tremendous challenge when one


considers the number and variety of parameters that can possibly affect the
governing mechanisms. There is a considerable risk of the fatigue design becoming
empirically based, and quite cost-ineffective, if rational guidelines based on phys-
ical models cannot be developed. To help alleviate this problem, we will in this
chapter develop a mechanisms-based framework for interpretating the fatigue
behavior of composites, beginning with the baseline configuration of unidirec-
tional fiber-reinforced plies and proceeding later to laminate configurations and
other fiber architectures. The framework in the form of fatigue-life diagrams will
allow assessment of the effects of constituent properties, such as fiber stiffness and
matrix ductility, and provide guidelines for fatigue design as well as for developing
mechanism-based life prediction models.
After a review of the fatigue-life diagrams and their utility, we shall discuss the
fatigue design methodologies, taking the examples of aircraft components and
wind turbine blades. Finally, a mechanisms-based modeling of multi-axial fatigue
will be discussed.

7.2 Fatigue-life diagrams

The S-N, or Wöhler diagram, originating from metal fatigue, is a familiar way to
represent the resistance of a given material to the cyclic application of loads.
It describes the observed fact that the material strength, given by the maximum
stress sustained in the first application of load, reduces with repeated application
of load, and is inversely dependent on the number of cycles applied. The strength
value corresponding to a pre-selected large number of cycles, e.g., 106, is custom-
arily taken as the fatigue limit. In some cases, a “true” fatigue limit exists,
representing the stress value below which a fatigue mechanism cannot be initiated,
but in most cases, one uses the operational definition of no failure until the
selected high number of cycles.
In a composite material with two constituents – fibers and matrix – the useful-
ness of the Wöhler diagram may be questioned. The obvious problem is: how do
238 Damage mechanisms and fatigue-life diagrams

Figure 7.1. Schematic illustration of fatigue-life diagram for a unidirectional composite,


showing three regions with different damage mechanisms.

we represent the roles of the two constituents in determining the fatigue-life, and
what roles do the constituents play in determining it? In order to address these
issues, a framework for conceptual interpretation of fatigue of composite materials
was proposed by Talreja [1]. That framework will be discussed below, beginning
with the case of on-axis tension–tension fatigue of unidirectional composites.

7.3 On-axis fatigue of unidirectional composites

Consider a unidirectional composite that is subjected to cyclic tension in the fiber


direction. Let the specimens of the composite be tested under load control, i.e., the
load varies between set limits of maximum and minimum. We shall examine the
mechanisms operating and the failure resulting under this condition. To do this,
let us construct a plot with the horizontal axis as a logarithm of the number of
cycles applied and the vertical axis as the maximum strain attained in the first load
cycle as seen in Figure 7.1. The reasons for plotting strain and not stress are
motivated by the following considerations. (1) The failure in the first cycle occurs
when the composite strain equals the failure strain of fibers irrespective of the fiber
volume fraction. A point with coordinates (log 1, ec), where ec is the composite
failure strain, can thus be plotted in the diagram. (2) The composite fatigue limit is
governed by the matrix fatigue limit, as will be discussed later. The matrix within
the composite is subjected to strain controlled fatigue, due to the fiber constraint,
although the composite is undergoing load controlled testing. Thus, the composite
fatigue limit is expressible in terms of strain.
Since the two extreme values of the fatigue-life are given in terms of strain, there
appears to be no compelling reason not to plot the strain between these limits.
We note that for a given fiber type (e.g., glass or carbon) and matrix material
the stress–strain relationship changes when the fiber volume fraction or fiber
stiffness change. The significance of plotting the maximum strain in the first cycle
lies in the fact that this strain value provides a good reference to the damage
attained in the first cycle and that the subsequent damage and fatigue-life are likely
to depend on this damage state.
7.3 On-axis fatigue of unidirectional composites 239

Figure 7.2. Nonprogressive fiber breakage in region I in fatigue of unidirectional composites.

As first described in Talreja [1], the test data plotted on the axes (log N, emax) can
be viewed as falling in three regions. Region I is the horizontally extending scatter
band (e.g., between 5% and 95% probability of failure) of the composite failure
strain. This region represents a lack of degradation in strength (failure strain),
i.e., the underlying predominant mechanism of fiber failure is nonprogressive.
The arguments to support this postulate will be described below. Region II is
the fatigue-life scatter band, which deviates from region I at a certain number of
cycles and extends down to the fatigue limit. This region is governed by the
progressive mechanism of fiber-bridged matrix cracking, as described below.
Finally, region III is the region of no fatigue failure (in a selected large number
of cycles, say 106) lying below the fatigue limit.
Figure 7.2 depicts three scenarios, the first of which shows fiber breaks
resulting from the first application of a high load that produces a maximum
strain lying within the scatter band of composite failure. These fiber breaks are
caused by the local fiber stress (or strain) exceeding the fiber strength at those
(weak) points during the first application of load. Unloading and reapplying
the load would change the local fiber stresses only if an irreversible (inelastic)
deformation occurs. Assume now that the matrix surrounding a fiber break
undergoes only small inelastic deformation due to the constraint of the stiff
fibers. This is likely if the matrix is relatively brittle, such as an epoxy. The case
of a matrix of high ductility (i.e., ability to flow easily) will be considered later.
For small inelastic deformation in the matrix the repeated application of load
would cause little change in the stresses on fibers surrounding a broken fiber
site. The next fibers to break in a subsequent load cycle could appear near any
of the previously broken fibers – but not necessarily near all previously broken
fibers – when the local stresses exceed their local strengths, see the illustration
at N = N1 in Figure 7.2. The important thing to note is that, due to the small
cycle-to-cycle changes in the local fiber stresses and randomness of the fiber
strength, it is unlikely that the fiber breakage will be a progressive mechanism,
in the sense that the number of fiber failures in a given location increases
monotonically. Thus the final failure, which results from a core of (a few) fiber
failures growing unstably (see illustration in Figure 7.2), could occur in any of
240 Damage mechanisms and fatigue-life diagrams

Crack front
(N = Nf)

Crack front
(N = N1)

Fiber-bridged matrix crack

Figure 7.3. Fiber bridged matrix cracking in region II in fatigue of unidirectional


composites.

Figure 7.4. Matrix cracking between fibers in region III in fatigue of unidirectional
composites.

the potential failure sites, without preference. The consequence of this is that
the composite failure under the specified loading condition can occur at any
number of load cycles. This can be described as a nonprogressive failure
mechanism, with no associated strength degradation. Thus the scatter band in
region I is horizontal.
The mechanism in region II is illustrated in Figure 7.3. At applied maximum
load levels below the lower bound of the region I scatter band, the composite
failure resulting from a cluster of neighboring fiber breaks is unlikely. Instead,
with increasing load cycles, the matrix will undergo fatigue cracking. The matrix
cracks will progress by failing fibers or by debonding and going around them.
A typical fiber-bridged crack would then appear as illustrated in Figure 7.3.
Among several such cracks the one to cause failure earliest would be the one that
grows unstably first.
Region III of the fatigue-life diagram can be viewed as the domain of no fatigue
failure in a pre-selected large (> 106) number of load cycles. One possible scenario
in this region is illustrated in Figure 7.4. As shown in the figure, fatigue cracks
in the matrix are likely to develop, but they remain confined to the parts of the
cross sections between fibers. The driving force for the cracks is insufficient to
advance them by failing and/or debonding fibers. Another likely scenario in this
region is fiber-bridged matrix cracking progressing at rates too low to cause
failure in the pre-selected large number of cycles. A “true” fatigue will exist only
if a mechanism of effectively arresting matrix crack growth is available.
7.4 Effects of constituent properties 241

7.4 Effects of constituent properties

The baseline fatigue-life diagram provides a good starting point for making an
assessment of the roles of constituent properties in determining the fatigue
response of composite materials. In the following we shall discuss some expected
trends in the fatigue-life caused by changes to constituent properties.
Fibers of different materials, e.g., glass, carbon, and SiC, have different axial
moduli and fail at different axial tensile strains. Carbon fibers of different stiffness
are commercially available and these fibers can fail at strains as low as 0.5%, or at
strains exceeding 1.8%, depending on processing and surface treatment. The com-
posite stiffness and failure strain in the fiber direction are mainly determined by the
fiber properties in most composites of practical interest. The failure of fibers is
subject to a significant scatter due to imperfections, in particular the surface defects.
The composite failure is therefore also a statistical process, with added complexities
of stress transfer between matrix and fibers, fiber debonding, fiber misalignment,
stress concentration in fibers near a broken fiber, etc. For our considerations
regarding the effect on composite fatigue of changing fibers and/or matrix proper-
ties we can view the fatigue-life diagram and identify the following trends, as shown
in Figure 7.5, where fiber-bridged cracking is the progressive mechanism.
1. As explained above, region I consists of the horizontally extended scatter band
of composite failure strain, ec. This strain can be raised or lowered by the fiber
failure strain. Thus, the placement of region I with respect to the fatigue limit
em (which is primarily a matrix property) is determined by the choice of fibers.
This in turn means that the extent of region II (ec  em) is also determined
by fibers. This provides a remarkable flexibility in view of the wide variety of
fibers that are commercially available and are being developed. Note also that
region II can be made to vanish (i.e., no fatigue!) by selecting fibers such that ec
is less than or equal to em, as we shall demonstrate later.
2. The progressive fatigue mechanism in region II of the fatigue-life diagram is
influenced significantly by fibers. The closing tractions on the matrix cracks are

Figure 7.5. Trends in fatigue-life diagrams due to constituent properties.


242 Damage mechanisms and fatigue-life diagrams

supplied by the bridging fibers as they are strained. For the same stretching
strain on the fibers, stiffer fibers would supply greater forces than would more
compliant fibers. Thus, lower fatigue degradation rates would result in com-
posites with stiffer fibers, leading to a rightward displacement of region II with
increasing fiber stiffness, as indicated in Figure 7.5.
3. Consider a ductile matrix versus a brittle matrix, in relative terms. A crack in a
ductile matrix would cause higher strains at the crack tips than in a brittle
matrix. Thus, higher crack opening displacement will result in a ductile matrix
than in a brittle one for the same crack length. This would mean higher strains
in the bridging fibers for cracks in composites with more ductile matrices,
resulting in earlier fiber failures than in the case of brittle matrices. In terms
of fatigue-life this would translate into lower fatigue-life for a more ductile
matrix, as indicated by the trend line in Figure 7.5.
4. The effect of constituent properties on the fatigue limit can be viewed in the
following way. First, fatigue in the composite cannot occur unless cracks can
be initiated in the matrix. Thus, the matrix fatigue limit in strain cycling is the
baseline limit (as described above). This limit can be enhanced by fibers that
provide a means of arresting (obstructing) matrix crack growth. However,
stiffer fibers would be more effective in suppressing the crack growth, than
would be more compliant fibers, by supplying higher closing pressure on the
matrix crack planes. This would result in the trend in the fatigue limit indicated
in Figure 7.5.

The interpretation of fatigue behavior of unidirectional composites under cyclic


on-axis tension will now be illustrated by considering test data.

7.5 Unidirectional composites loaded parallel to the fibers

The fatigue-life diagram, described above, is developed by generic considerations


of damage mechanisms. In what follows we will illustrate the diagram for specific
material systems.

7.5.1 Polymer matrix composites (PMCs)


The fatigue-life diagram will be used as a baseline diagram for interpretation of
fatigue of polymer matrix composites. Fatigue of other material systems – metal
matrix composites and ceramic matrix composites – then becomes easy to inter-
pret by way of emphasizing the differences and similarities with respect to polymer
matrix composites.
To begin, consider the fatigue-life data shown in Figure 7.6 for a glass/epoxy
composite of three different fiber volume fractions indicated in the figure. These
data are plotted in the conventional manner with the applied cyclic stress
7.5 Unidirectional composites loaded parallel to the fibers 243

Figure 7.6. Stress/life data of a unidirectional glass/epoxy composite loaded in tension


parallel to the fibers [2]. Reprinted, with kind permission, from Fatigue in Composite
Materials, copyright ASTM International, 100 Barr Harbor Drive, West Conshohocken,
PA 19428.

Vf
0.50
0.33
0.024 0.16
ec

0.016
emax

0.008
em

0 2 4 6 8
log Nf

Figure 7.7. Fatigue-life diagram of a unidirectional glass/epoxy composite loaded in tension


parallel to the fibers [2]. Reprinted, with kind permission, from Fatigue in Composite
Materials, copyright ASTM International, 100 Barr Harbor Drive,West Conshohocken,
PA 19428.

amplitude as the vertical axis. The fatigue-life curves appear as distinct lines for
each case of the fiber volume fraction. The data are replotted in Figure 7.7 using
the maximum strain applied in the first cycle as the vertical axis, in accordance
with the fatigue-life diagram. The data for the three volume fractions are found to
fall together, as seen in the figure. The conceived fatigue-life diagram with the
three regions has been superposed on the data. The horizontal scatter band of
composite failure strain is drawn as region I and fatigue limit of the epoxy resin in
strain control testing, reported by Dharan [2] to be at 0.6% strain, has been drawn
as the assumed fatigue limit of the glass/epoxy composites. A scatter band has
been placed about the fatigue-life data to indicate region II. This plot suggests that
plotting fatigue-life against strain allows more meaningful interpretation, as
argued above. It also provides confirmation of the proposition that the fatigue
244 Damage mechanisms and fatigue-life diagrams

(a) (b)
s Fiber

s
Fiber

Composite
Composite
Matrix Matrix

em ec e em ec
e

Figure 7.8. Effect of fiber stiffness and strain to failure on composite stiffness in a
unidirectional composite in longitudinal loading, where ec is the composite failure strain
and em is the matrix fatigue limit. (a) Composite with low stiffness fibers; (b) composite with
high stiffness fibers.

limit of the matrix is a good indication of the composite fatigue limit, when plotted
in the strain coordinate. The existence of region I is not strongly indicated by the
glass/epoxy data, for reasons that will become clear as we examine other data.
Consider now a set of unidirectional carbon/epoxy composites where each com-
posite has different type of carbon fibers and the same epoxy matrix. Figure 7.8
illustrates two examples of stress–strain behavior, one with relatively low stiffness
fibers (Figure 7.8(a)) and the other with relatively high stiffness fibers (Figure 7.8(b)).
The composite failure strain ec, assumed equal to the fiber failure strain, has been
marked on the strain axis for both cases. Note that the failure strain is lower for
higher fiber stiffness in accordance with observed behavior of carbon fibers. Also
marked is the fatigue limit of epoxy, at the same value in both cases, assuming for
simplicity that it is a matrix property not affected by fibers.
The consequence of changing carbon fiber stiffness is thus essentially to increase or
decrease the extent of region II. Imagine now that a certain carbon fiber is used that
has such high stiffness that the composite failure strain ec < em, the fatigue limit. This
can also be caused by defective fibers of low strength. The consequence of composite
failing at strains below the fatigue limit is that no fatigue progression would be
possible. In other words, region II would not exist, as the fatigue limit would lie
above region I. Thus, only region I (a horizontal scatter band) will appear in the
fatigue-life diagram. Figure 7.9 illustrates this phenomenon by using fatigue-life data
gathered by Sturgeon [3] for a carbon/epoxy composite with high stiffness fibers. The
fatigue limit, assumed for epoxy to be 0.6% strain, has been marked in the figure
along with the average composite failure strain at approximately 0.48% strain.
The next set of test data, plotted in Figure 7.10, is from Awerbuch and Hahn [4]
for carbon/epoxy with fibers of stiffness lower than those in the composite used by
Sturgeon [3] just discussed. Here the data suggest no fatigue degradation when
viewed without the fatigue-life diagram. However, when the scatter band of region
7.5 Unidirectional composites loaded parallel to the fibers 245

emax
em
0.006

ec

0.004

0.002

0 4 8 log Nf

Figure 7.9. Fatigue-life diagram of unidirectional composite reinforced by stiff fibers with
low strain to failure (data from [3]). It is notable that there is no fatigue degradation, since
the mechanism is nonprogressive fiber breakage only.

emax

0.010
ec

em
0.006

0.002
0 2 4 6 log Nf

Figure 7.10. Fatigue-life diagram of unidirectional composite reinforced by medium stiffness


fibers, showing a narrow range of strain where fatigue occurs [4]. Reprinted, with kind
permission, from Fatigue of Filamentary Composite Materials, copyright ASTM
International, 100 Barr Harbor Drive, West Conshohocken, PA 19428.

I and the fatigue limit of epoxy are placed in the diagram, the test data suggest a
different interpretation. The progressive fatigue mechanism of region II is now
seen to exist, albeit in a narrow range of strain.
Another case of differing interpretation with and without the fatigue-life
diagram is shown in Figure 7.11, where data from Sturgeon [5] for a lower stiffness
carbon/epoxy are plotted. Contrary to Sturgeon’s assertion that no fatigue deg-
radation exists for this material, the fatigue-life diagram suggests that, depending
on where the fatigue limit lies, there would be a region II of progressive fatigue
damage. In the figure the fatigue limit is indicated at 0.6% strain as a reference
for comparison with other cases of carbon/epoxy composites.
As the last case of carbon/epoxy fatigue we take the data supplied by
P. T. Curtis (personal communication) plotted in Figure 7.12. Here the regions
of the fatigue-life diagram appear distinctly, illustrating the power of interpret-
ation offered by the diagram. The value of recognizing the existence of region I is
246 Damage mechanisms and fatigue-life diagrams

emax

0.02

ec

0.01

em

0 4 8 log N

Figure 7.11. Fatigue-life diagram of a unidirectional composite with low stiffness fibers
(data from [5]), with a relatively wide range of strain where fatigue occurs.

emax (%)
1.5
εc

1.2

0.9
em
0.6

0 1 2 3 4 5 6 7
log Nf

Figure 7.12. Fatigue-life diagram of a unidirectional carbon fiber/epoxy composite with


a distinct region I (data courtesy of P. T. Curtis, RAE, UK).

emax (%)
ec 2.0

1.5

1.0
em
0.5

0.0
0 1 2 3 4 5 6 7
log Nf

Figure 7.13. Fatigue-life diagram of a unidirectional Kevlar fiber/epoxy composite (data from [6]).

particularly evident. Without this, the error introduced in assessing the fatigue-life
would be significant.
As further illustration of the fatigue-life diagram, Figures 7.13 and 7.14 show
test data for Kevlar®-epoxy and Kevlar®-J-2 polymer, respectively. The J-2
7.5 Unidirectional composites loaded parallel to the fibers 247

emax (%)

2.0

1.5

1.0

0.5

0.0
0 1 2 3 4 5 6 7
log Nf

Figure 7.14. Fatigue-life diagram of a unidirectional Kevlar fiber/J2 polymer composite


(data from [6]).

Figure 7.15. A crack on the surface of a carbon/epoxy unidirectional composite subjected to


tension–tension fatigue. The crack tips are squeezed by the bridging fibers, as illustrated
schematically by the bottom figure. Reprinted, with kind permission, from Springer Science
+Business Media: J Mater Sci, Fatigue damage mechanism in unidirectional carbon
fibre-reinforced plastics, Vol. 34, 1999, pp. 2535–46, E.K. Gamstedt and R. Talerja.

polymer is an amorphous polyamide thermoplastic [6]. The generality of the


fatigue-life diagrams is that they lend themselves to description and interpretation
of any kind of fatigue-life data for composites. A comparison of fatigue perform-
ances between different material candidates can then be made. Commercial mater-
ial systems of unknown constituent composition can thus be compared, and the
best contender for an intended application can be identified.

7.5.1.1 Experimental studies of mechanisms


The fatigue-life diagrams developed in [1] and described above for PMCs were
based on deductive reasoning in the absence of any systematic investigation
of such mechanisms. Since then an investigation was conducted by Gamstedt
and Talreja [7], which, through direct observations of on-axis tension–tension
fatigue of unidirectional carbon/epoxy, gathered evidence to support the diagrams.
For instance, the presence of fiber-bridged cracking postulated in region II was
248 Damage mechanisms and fatigue-life diagrams

Fiber/matrix debonding

Matrix crack

Figure 7.16. A surface replica image showing a matrix crack arrested by fiber/matrix
debonding. Reprinted, with kind permission, from Springer Science+Business Media:
J Mater Sci, Fatigue damage mechanism in unidirectional carbon fibre-reinforced plastics,
Vol. 34, 1999, pp. 2535–46, E.K. Gamstedt and R. Talerja.

found as exemplified by an image of a surface replica shown in Figure 7.15.


The crack opening profile shows squeezing of the crack tips by fibers.
A mechanism for arresting matrix crack growth in region III was also dis-
covered in [7]. Its evidence is displayed in Figure 7.16. As seen there, a matrix
crack growing transverse to the applied load is arrested by extensive fiber/matrix
debonding. Without the energy dissipation in the debonding process the crack will
have a greater driving force for extension.
In [8] the fatigue propagation of fiber-bridged cracks in unidirectional PMCs was
analyzed. As expected, in comparison to an unreinforced matrix crack, the fatigue
growth rate of fiber-bridged cracks was found to be lower, and even showing
decelerating growth. In another study, Gamstedt [9] examined the role of fiber/
matrix interfaces prone to debonding on the composite fatigue under on-axis
tension–tension loading. If the interfaces resist debonding, which may be viewed
as the baseline (normal) case, the fatigue mechanisms described above in developing
the fatigue-life diagram would apply. As debonding increases, the fiber strength
variability becomes increasingly important in determining the damage progression,
and in significant debonding cases, the fiber-bridged crack growth gives way to
debonding-assisted fiber breakage as the progressive mechanism. The slope and
extent of region II of the fatigue-life diagram is then determined by the debonding-
assisted fiber breakage. The flat region I, which is a consequence of the unconnected
fiber breakage, then diminishes, and can eventually become part of region II.
Figure 7.17 illustrates the cases of little or no debonding vs. extensive debonding.
The former case is typical of relatively strong interfaces in carbon/epoxy composites.
When fiber breaks occur, initially at randomly located weak points, they remain
isolated and unconnected due to the lack of debonding. As discussed above in
7.5 Unidirectional composites loaded parallel to the fibers 249

(a) (b)

Figure 7.17. Two cases of the role of fiber/matrix debonding in fatigue damage progression
under on-axis tension–tension loading of unidirectional fiber-reinforced PMCs. Case 1 (a)
illustrates unconnected fiber breaks when little or no debonding occurs while case 2 (b) is
for extensive debonding that connects fiber breaks and causes the fiber breakage to be
progressive. Reprinted, with kind permission, from Springer Science+Business Media:
J Mater Sci, Fatigue damage mechanism in unidirectional carbon fibre-reinforced plastics,
Vol. 34, 1999, pp. 2535–46, E.K. Gamstedt and R. Talerja.

describing region I, any of the initial breaks are likely to form a critical-size crack that
could grow unstably, leading to failure. Thus there is no progressive mechanism and
composite failure results essentially at any number of cycles, resulting in a horizontal
scatter band. In the case of interfaces that are prone to debond, the influence of an
initial fiber break extends to other fibers, depending on the rate at which the debond
crack elongates, and a progressive fiber breakage results. The rate of this damage
progression depends on the resistance to the debond crack growth given by the
interface region and the likelihood of the debond cracks connecting with other cracks.
The overall path to final failure may be described in terms of an average progression
rate. Modeling of this progression is far from an easy task. Attempts at this were
made by Gamstedt [9], who conducted numerical simulation of the fiber breakage
process and clarified the roles of interfacial debonding and fiber strength variability.
Figure 7.18 compares the fatigue-life diagrams of unidirectional carbon/epoxy and
carbon/PEEK (polyetheretherketone) composites loaded in tension–tension loading.
The former material typifies the strong fiber/matrix bonding case (left in Figure 7.17)
while the latter is prone to debonding. The PEEK matrix is a semi-crystalline thermo-
plastic that develops a so-called trans-crystalline structure extending from the fiber
surface. This structure provides weak planes parallel to fibers for debonding to occur.
Although otherwise a tough material, PEEK becomes a source of brittle cracking of the
interfaces. The extensive debonding in carbon/PEEK removes region I of the fatigue-
life diagram by inducing progressive fiber breakage, as explained above.
250 Damage mechanisms and fatigue-life diagrams

Max. strain (%) 1.5

1.2

0.9

Carbon/PEEK
Carbon/epoxy
0.6

0 1 2 3 4 5 6 7
Log cycles

Figure 7.18. Comparison of the fatigue-life diagram for unidirectional carbon/epoxy and
carbon/PEEK composites loaded in tension–tension cycles along fibers. Note the absence
of region I in carbon/PEEK.

7.5.2 Metal matrix composites (MMCs)


The fatigue-life diagram discussed above should be viewed as a baseline diagram
for polymer matrix composites, as the mechanisms considered for its construction
were motivated by observations and conjectures related to this material system.
We shall now discuss what changes in the diagram are plausible when the matrix
material is a metal. Note that the role of fibers is in modifying the fatigue
mechanisms taking place in the matrix, which is the material conducting irrevers-
ible deformation (plasticity). The three regions of the fatigue-life diagram for
MMCs treated by Talreja [10] will be discussed next.

Region I: As discussed earlier, this region is manifested by fiber failures.


For polymer matrix composites we argued that the fiber failures occurred in a
manner that did not have significant progressiveness (accumulation of fiber
failures in a localized zone), leading to the final (composite) failure from
random sites at random number of cycles. An important factor in this deduction
was insufficient irreversible deformation in the matrix to allow cycle-dependent
stress enhancement and failure of fibers. This scenario will change when a
relatively brittle polymer matrix is replaced by a more ductile metal matrix.
The cyclic plastic deformation of a metal matrix around a broken fiber will
redistribute stresses in the neighboring fibers, allowing an accumulative fiber
failure process to occur. This will introduce a localized progressive degradation,
which on reaching a critical level will cause composite failure. Thus, the hori-
zontal scatter band of region I, characteristic of polymer matrix composites, will
now have a downward slope.
Region II: The mechanism of fiber-bridged matrix cracking is also expected to
be the primary progressive mechanism for metal matrix composites. The
7.5 Unidirectional composites loaded parallel to the fibers 251

Figure 7.19. Fatigue-life diagram of SCS6/Ti-15–3 at room temperature.

Figure 7.20. Fatigue-life diagram of SCS6/Ti-15–3 at high temperature (540–550 C).

difference with respect to polymer matrix composites lies in the role of fiber/
matrix debonding. The fiber/matrix bond in metal matrix composites is generally
stronger, leading to shorter extent of the interface failure. Furthermore,
increased ductility of the matrix is expected to provide more crack opening
displacement via crack tip blunting, causing increased failure of the bridging
fibers. The matrix ductility effect will thus be as conjectured above in describing
the trend in region II of the fatigue-life diagram as illustrated in Figure 7.5.
Region III: This region is expected to be the same as in polymer matrix
composites. The composite fatigue limit is also expected to be related to the
matrix fatigue limit.
Let us now examine some test data for metal matrix composites. Figure 7.19
plots fatigue-life at room temperature for a SCS6/Ti-15–3 unidirectional composite
under cyclic tension at two R-ratios. Static failure strain is plotted on the vertical
axis at the first cycle. As discussed above, region I displays some progressiveness by
deviating from a horizontal scatter band. Region II shows greater degradation
rates and tends to the fatigue limit gradually. Note that the matrix fatigue curve,
plotted as a broken line, tends to approximately the same fatigue limit as the
composite. The data for the same composite tested at a high temperature are
shown in Figure 7.20. The general features of the fatigue-life diagram at room
temperature are retained with different quantitative characteristics. Comparisons
252 Damage mechanisms and fatigue-life diagrams

Figure 7.21. Region I for SCS6/Ti-15–3 at high (540–550 C) and low temperature (20 C).

Figure 7.22. Region II for SCS6/Ti-15–3 at high (540–550 C) and low temperature (20 C).

of the two regions are displayed in Figures 7.21 and 7.22. Figure 7.21 shows the
data in region I at room and high temperatures on log–log scales. The best-fit lines
for the data show that higher fatigue degradation (more progressiveness in the
fiber failure mechanism) occurs at high temperature than at room temperature.
Similar region II comparison at the two temperatures is shown in Figure 7.22.
A leftward shift of region II, indicative of more fatigue degradation, is seen with
an increase in temperature. This is likely due to higher matrix ductility (lower yield
stress) at the high temperature, in accordance with the trends in the fatigue-life
diagram described above.

7.5.3 Ceramic matrix composites (CMCs)


In comparison with polymers and metals, ceramics have insignificant irreversi-
bility of deformation. This suggests that there should be little incentive for fatigue
mechanisms to occur in ceramics. However, when ceramics are reinforced by
7.5 Unidirectional composites loaded parallel to the fibers 253

e = 0.1 5 % e = 0.5 % e = 0.8 %

Fiber-bridged
matrix crack

Fibers

500 µm

Figure 7.23. Micrographs of surface replicas of a unidirectional CMC with increasing strain:
0
15%, left; e = 0
5%, middle; e = 0
8% right. Source: [11].

fibers with which they bond weakly, a dissipative mechanism becomes available
from reversed (i.e., back-and-forth, not reversible) frictional sliding at interfaces.
This mechanism is believed to play a major role in causing fatigue of CMCs. In the
following a discussion of the fatigue-life diagram for unidirectional CMCs under
cyclic axial tension is presented.
Before getting into a discussion of fatigue of CMCs it would be useful to briefly
review the damage and accompanying stress–strain response observed under mono-
tonic axial tension. Several works in the literature have reported such data, and we
shall mainly draw upon Sørensen and Talreja [11] to illustrate some characteristics
of interest. Figure 7.23 shows three surface replicas of a SiC/CAS (calcium alu-
minosilicate) glass specimen taken at three applied axial strains at room tempera-
ture. For reference, the failure strain is approximately 1.0%. Note that at 0.15%
strain a few irregularly spaced cracks transverse to fibers are seen. At this low strain
level the cracks are partially developed, i.e., they do not extend across the entire
width of the specimen. As the applied strain increases, the cracks span the width
fully and acquire more regular spacing. At some point the crack spacing approaches
a minimum value, usually described as saturation, as seen in the replica at 0.8%
strain. The stress–strain response in the axial as well as transverse directions is
shown in Figure 7.24, alongside a plot of the AE (acoustic emission) events recorded
during testing. The double stress–strain curves are due to strain gages used on both
sides of the specimen to monitor axiality of loading. Figure 7.25 summarizes the
different stages of damage progression, indicating the ranges in which they operate,
based on microscopy observations and AE event counts.
When considering fatigue, the first question to ask is: what happens in the
material in the second and subsequent load cycles that is different from the
254 Damage mechanisms and fatigue-life diagrams

600

500

400

s (MPa)
300 eL
eT
200

100

0
2.0 1.5 1.0 0.0 0.0 0.25 0.50 0.75 1.00
AE events (×106) e (%)

Figure 7.24. Measured acoustic emission events and stress–strain response. Source: [11].

Monotonic tension
600
SiCf /CAS-II
IV Fiber
RT, 100 MPa/s
fracture
500 IV

III Fiber
400 bridging
Stress (MPa)

III
(Matrix fully
cracked)
300

spl = 285 MPa II


II Matrix crack t
200 development en
m
lop
(initiation of ve
smc = 120 MPa matrix crack) de
e
ag
100 I m
I No Da
damage
0
0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
Strain (%)

Figure 7.25. Schematic drawing of damage mechanisms at different stages in monotonic


tensile loading of a unidirectional CMC loaded in tension parallel to the fibers.

damage caused, if any, in the first load cycle? The answer to this question lies in
knowing the source of irreversibility in the material. For instance, if a material
is cracked (singly or multiply) in the first load cycle, it would not be cracked
further in the next and subsequent cycles unless an irreversible mechanism such as
plasticity or friction is present. As stated earlier, ceramic materials have insignifi-
cant irreversible (plastic) deformation to enable crack growth in cyclic loading.
The likely source of irreversibility is frictional sliding at the interface. Keeping this
in mind we shall examine below the existence and nature of the three regions of the
fatigue-life diagram for CMCs (see Figure 7.26).
7.5 Unidirectional composites loaded parallel to the fibers 255

Figure 7.26. Mechanisms in the three different regions of the fatigue-life diagram for
unidirectional CMCs loaded in tension parallel to the fibers.

Region I: When a first load is applied up to a maximum value within the scatter
band of the composite failure strain, the matrix cracks develop beyond their
saturation stage (see Figure 7.25). This means that a set of fully grown fiber-
bridged cracks exist at the end of the first load cycle. It is likely that a significant
portion of the fibers will be broken due to the high load applied. When the
second and subsequent cycles are applied, the fiber stress will increase due to the
increased debond length caused by the irreversible frictional sliding, supplying
the increased shear stress at the interfaces. Also, the cyclic grinding of the fiber
surfaces by the debris generated by the interfacial wear will damage fibers and
degrade their strength. All this will fail the fibers and when any of the cross
sections going through the crack planes has insufficient number of intact fibers
to bear the applied load, composite failure will occur. Now the critical question
is whether this failure mechanism is progressive or nonprogressive. If failure
were to come from only one cross section, i.e., if only one crack existed, or if one
crack was somehow the favored one, then the accumulative fiber breakage will
result in a progressive failure. However, that is not characteristic of the damage
here. We have multiple cracks, and failure could potentially result from any one
of those cracks. Although each fiber-bridged crack has progressive (accumula-
tive) fiber failure, the rate of progression would likely be discontinuous and
differ from that of other cracks. Thus the final failure would not necessarily
result from the weakest of the cross sections produced by the first load. We
postulate, therefore, that the number of cycles for composite failure in this case
cannot be derived from a single rate equation, and consequently, the failure
process is nonprogressive.
Region II: Consider now the case of loading where the first cycle maximum load
is in a range such that the matrix cracks are either partially developed or are
256 Damage mechanisms and fatigue-life diagrams

fully developed but the bridging fibers are not broken extensively. The partial
cracks will grow in subsequent cycles by debonding and/or breaking fibers at the
crack tips (or fronts). The cyclic frictional sliding at the interfaces will debond
the bridging fibers, causing crack surface separation to increase as the debond
length increases and as more bridging fibers break. With cyclic stressing, there-
fore, partial cracks will grow to their full extent. As more debonding occurs,
stress will be transferred from fibers to the matrix increasingly, resulting in
matrix cracking between two pre-existing cracks. A state of crack saturation
will be approached eventually, unless composite failure results from failure of
all fibers bridging any of the cracks. The damage development in this region of
the fatigue-life diagram has well-defined progressiveness from partial cracking
to full cracking until crack saturation. The terminal point is when at least one
crack loses all its bridging fibers. This point may occur before crack saturation
(toward the lower end of region II) or beyond crack saturation (toward the
higher end of region II). The main difference between this region and region I is
that in the latter saturation cracking with extensive fiber breakage already
occurs in the first load application.
Region III: We have defined this region to be a range of loading (maximum
strain in the first cycle) in which damage may develop but does not lead to
a critical state (failure) in a pre-selected large number of cycles (typically 106).
In a unidirectional composite, failure (separation) will occur when all the
bridging fibers of at least one crack fail, or, alternatively, broken fiber ends
are interlinked by interfacial debonds and matrix cracks to form two separating
surfaces. In either case, interfacial frictional sliding plays an important role
by increasing debonding and thereby stressing fibers to failure. What has been
found, however, is that the wear of the debonded interfaces (breakage of
asperities resulting in smoothing) can make the frictional sliding less effective.
This could slow down the cyclic development of damage to the extent of not
reaching failure in the pre-specified number of cycles. Sørensen et al. [12] found
that the temperature rise of SiC/CAS specimens caused by frictional heating at
debonded interfaces increased sharply just before failure, while for specimens
that did not fail in 108 cycles the temperature rise stopped at some cycles and
then took a downward turn (Figure 7.27). These authors have challenged the
notion of a “true” fatigue limit in CMCs, arguing that damage continues to
develop even after 108 cycles, as long as interfacial debonding occurred in the
first cycle. They contend that the condition for no damage development is no
initiation of fiber-bridged crack, providing the true fatigue limit as the stress
(or strain) at which such crack forms.

The discussion concerning fatigue limit in CMCs is somewhat academic because


the energy dissipating mechanism of frictional sliding at debonded interfaces
depends on the loading frequency and specimen geometry. This is because the
time rate at which conduction of the heat generated by frictional sliding occurs
affects the rate per cycle of energy dissipation by frictional sliding.
7.6 Unidirectional composites loaded inclined to the fibers 257

40
[016] SiCf / CAS II
X
200 Hz, RT
30 smin = 10 MPa
Temperature rise (K)

X
220 MPa

20
Run
out
smax = 260 MPa
10
212 MPa

0
1 102 104 106 108
Cycles N

Figure 7.27. Temperature increase due to frictional heating during cyclic loading of
a unidirectional SIC/CAS II composite.

The fatigue data on CMCs reported in the literature are inadequate for a good
illustration of the usefulness of the fatigue-life diagram. Since the testing was
done with a traditional S-N curve in mind the data generated were too few in
region I to validate its existence. Also, because of the material cost, the data
generated were at too few load levels to have all three regions of the diagram
appear with clarity.

7.6 Unidirectional composites loaded inclined to the fibers

We consider the simple case of a single fiber orientation in a composite placed at


an angle to the applied cyclic tension load as shown in Figure 7.28. As illustrated
in the figure, cracks initiate along the fibers, either at interfaces or in the matrix.
At first the cracks may initiate from defects in the matrix or at interfaces and grow
with cycles through the thickness as well as along the fibers. At some stage in the
evolution of this damage, a single crack may grow to the extent of attaining
unstable growth in the next application of load and separate the composite in
two pieces, as illustrated in the figure.
In comparison to the on-axis loading case, i.e., when the fibers are aligned with
the loading direction, the damage mechanism of the off-axis loading case is
drastically different and dramatically simple. The single progressive mechanism
depicted in Figure 7.28 will produce a continuous curve (and the associated scatter
band) in the fatigue-life diagram starting at the first-cycle failure strain and ending
asymptotically in the fatigue limit. For each off-axis angle the curve will be different.
A schematic depiction of the fatigue-life diagram is shown in Figure 7.29, which
illustrates the fatigue-life dependence on the off-axis angle. The on-axis fatigue-life
diagram is shown for reference in the figure in broken lines.
258 Damage mechanisms and fatigue-life diagrams

Figure 7.28. Cracking in a unidirectional composite under off-axis cyclic tension (left)
and failure from growth of a crack (right).

emax
q

ec
(q = 0)

ec
(0 < q < 90)

log N

Figure 7.29. Fatigue-life diagram for unidirectional composites under off-axis loading.
The diagram in broken lines is for the on-axis loading case.

As Figure 7.29 illustrates, region I (horizontal scatter band) of the fatigue-life


diagram does not exist when the applied cyclic load is inclined to the fiber
direction. The single scatter band starts at the composite failure strain and
asymptotically ends at the fatigue limit. The failure strain as well as the fatigue
limit strain depend on the off-axis angle, y. From the S-N data reported for glass/
epoxy in [13] the fatigue-life diagrams for a few angles are shown in Figure 7.30.
The curves drawn by visual fit to the data show the trends depicted in Figure 7.29.
The fatigue limits extracted from the data are shown plotted in Figure 7.31 where
the fatigue limits deduced from [14] for angle-ply laminates of glass/epoxy
are also plotted. The significant improvement in the angle-ply laminates over
unidirectional composites for the same off-axis angle is discussed in the next
section on fatigue of laminates.
7.7 Fatigue of laminates 259

0.008
ec q /deg
5
(q=5°) 10
30
em 60
0.006

ec
(q=10°)
emax

0.004

ec
(q =30°)

0.002
ec
(q =60°)
edb

0 2 4 6
lg N

Figure 7.30. Fatigue-life data for glass/epoxy unidirectional composites from [13] plotted
as fatigue-life diagrams for different off-axis angles of tension–tension loading. The fatigue
limit of on-axis loading is indicated at 0.6% strain for reference. Reprinted, with kind
permission, from Z. Hashin and A. Rotem, J Compos Mater, Vol. 7, pp. 448–64, copyright
# 1973 by Sage Publications.

0.006

0.004
efl

0.002

0 30 60 90
q (degrees)

Figure 7.31. The off-axis fatigue limit from data in Figure 7.30 (dashed line) is shown
for comparison with the fatigue limit of angle-ply laminates of glass/epoxy taken from data
reported in [14], plotted against the off-axis ply angle.

7.7 Fatigue of laminates

Composite structures are built by placing fibers in different orientations to


effectively carry multi-axial loading. We will here consider how the mechanisms
of fatigue damage are affected by multi-directional fiber placement in a
260 Damage mechanisms and fatigue-life diagrams

laminate subjected to a cyclic tensile load and how that translates into changing
the basic fatigue-life diagram described above.

7.7.1 Angle-ply laminates


These laminates have two fiber orientations placed symmetrically about a princi-
pal direction, e.g., [ yn]s illustrated in Figure 7.32. When loaded in the axial,
i.e., 0 direction, a y-ply in the laminate behaves differently from a unidirectional
composite loaded off-axis at the same y-angle. First, for the same applied axial
strain, the stress state in the ply within the laminate differs from that in the
unidirectional composite. Second, the unidirectional composite fails when a crack
appearing along its fibers grows unstably, while in the laminate the unstable
growth of a ply crack does not cause laminate failure. Instead, the presence of
the interface to the ply (or plies) in the other off-axis orientation leads to multiple
ply cracking. The so-called shear lag process underlying the crack multiplication
has been treated in detail in Chapter 4. Figure 7.32 depicts schematically the
multiple cracking and the consequent (associated) delamination. As a ply crack
front encounters the ply interface, the intense stress state ahead of the crack front
debonds the interface, causing local ply separation (delamination). This delamina-
tion accompanies each ply crack and under cyclic loading undergoes growth along
the interfacial plane. While this occurs, the multiple cracking process within the
plies can continue due to stress redistribution in the plies, if the renewed stress
state is high enough. Under what conditions would the renewed stress state be
critical for producing new cracks is a subject of damage evolution modeling and
will be discussed in a subsequent section of this chapter.
The key elements of the fatigue failure process in laminates are illustrated by the
angle-ply fatigue, and may be summarized as: multiple ply cracking, delamination
formation, and growth, possibly followed by more ply cracking, and delamination
merger in ply interfaces, leading to separation of plies, which in turn can overload

q q Delamination

Matrix
crack

Figure 7.32. Multiple matrix cracks in off-axis ply of a laminate (left) and subsequent
delamination caused by fatigue (right). For clarity, cracks and delamination are shown
for one ply only; the other ply is indicated in broken lines.
7.7 Fatigue of laminates 261

individual plies resulting in ply failure (separation). The laminate failure (separ-
ation in parts) comes from the final event of fiber failures.
Thus while an off-axis loaded unidirectional composite has essentially a two-
stage fatigue process consisting of ply crack initiation and growth until instability,
in laminates the fatigue process generally has several interactive cracking
processes. One can lump them in two parts as the critical and sub-critical failures.
The critical failure is the fiber breakage process, which itself can have its progres-
sion. The sub-critical failure process consists of all events until fiber failures.
The enhancement of the off-axis fatigue limit in angle-ply laminates shown
by data in Figure 7.31 can now be explained by the presence of the sub-critical
failure processes in laminates and its absence in unidirectional composites loaded
at off-axis angles.
The fatigue-life diagram for angle ply laminates is expected to lose region I,
except at the y-angle approaching 0 . The fatigue limit will not be adversely
affected until the sub-critical damage processes become insignificant, such as at
y > 30 for the data in Figure 7.31, and at y = 90 the fatigue limit will be the
same as for a unidirectional composite under transverse loading. It is possible that
at small y-angles some enhancement in the fatigue limit results from the lamin-
ation effect of inducing multiple cracking. Experimental data does not seem to be
available, however, to confirm this.

7.7.2 Cross-ply laminates


Cross-ply laminates, [0n/90m]s, are orthogonal angle-ply laminates, on which
loading is usually applied in one of the two fiber directions. Historically, this
laminate was the first class of laminates for which multiple ply cracking was
experimentally observed and reported in the literature in the late 1970s (see
Chapter 3 for details). In those studies, monotonically increasing tensile load
was applied along the 0 -direction, resulting in multiple ply cracks in the 90 -plies.
Since multiple cracking is a basic feature of damage in composite materials, this
configuration and loading combination has been, and continues to be, a test-bed
for modeling studies of multiple cracking and its consequences on the material
response.
The observations of multiple cracking in cross-ply laminates under monotonic-
ally increasing load do not, however, reveal the key feature underlying progres-
siveness of damage under cyclic loads. Since most studies typically observed the
cracking process on the free edges of a flat specimen, the details of damage
accumulation were missed. The first study to examine the interior of a laminate
with painstaking patience and using X-ray radiography, combined with stereo
radiography, was by Jamison et al. [15]. The key X-ray picture, along with a
schematic to depict interior details, is shown in Figure 7.33.
Other than the transverse cracks, the details seen in Figure 7.33 are not found to
develop sufficiently under monotonic loading. However, these details hold the key
to the further progression of transverse cracking under cyclic loading. As we shall
262 Damage mechanisms and fatigue-life diagrams

Interior
delaminations

Transverse ply
Interior
cracks delaminations
Axial Transverse
Axial splits cracks
splits

Figure 7.33. X-ray radiograph of a carbon/epoxy cross-ply laminate taken after tension–
tension fatigue shows transverse ply cracks, interior delamination, and axial splits. The
accompanying schematic clarifies the interior details [15]. Reprinted, with kind permission,
from Effects of Defects in Composite Materials, copyright ASTM International, 100 Barr
Harbor Drive, West Conshohocken, PA 19428.

discuss below in the section on modeling, the irreversible changes needed from one
load cycle to the next can be lumped into frictional sliding between the surfaces of
the propagating delamination.
In constructing the fatigue-life diagram of a cross-ply laminate under cyclic
axial tension, one needs to ask certain basic questions. First, are conditions
available for region I to be present? As discussed above in explaining the reasoning
behind this region, a statistical nonprogressive fiber breakage mechanism must be
available for this region to exist. The presence of 0 -plies in the laminate makes
this mechanism plausible at applied axial strains near the average fiber breakage
strain that leave the laminate intact at the first application of load.
Second, where is the fatigue limit? To determine the fatigue limit, if not
available from test data, a good estimate will be given by the strain at which
transverse cracking initiates. The reasoning is simply that if no cracks initiate in
the transverse plies, no progression of crack multiplication and subsequent
damage are possible. The models for crack initiation strain have been discussed
in Chapter 6.
What remains to complete the fatigue-life diagram for cross-ply laminates is the
location of region II. If the scatter-band of this region is assumed to be straight
and sloping, a good approximation to begin with, then by fixing its lower end at
the fatigue limit (at, say, 106 cycles), its slope or the point of its deviation from
region I remain to be determined. A model that has been successfully determined
will be discussed below in the section on fatigue-life prediction.
Figure 7.34 shows data reported by Grimes [16] for a carbon/epoxy cross-ply
laminate, plotted on the strain scale. The anticipated fatigue-life diagram, as
discussed above, is superimposed on the data. The scatter-bands drawn are guided
by the data, as insufficient data have been reported to estimate these bands from
probability distributions. The location of the fatigue limit strain has been taken at
the value of stress (converted to strain) reported by the author as the value where
the first transverse cracking was found.
7.7 Fatigue of laminates 263

Debonding in 90˚ plies,


delamination
ec
0.008
emax

edl
0.004

0 2 4 6
log N

Figure 7.34. Experimental data verifying the anticipated fatigue-life diagram of cross-ply
laminates, with data from [16]. Reprinted, with kind permission, from Composite Materials:
Testing and Design, copyright ASTM International, 100 Barr Harbor Drive, West
Conshohocken, PA 19428.

7.7.3 General multidirectional laminates


For practical structures the composite architecture is designed to satisfy multiple
requirements, such as resistance to bending and torsion as well as thermal expan-
sion. A common composite architecture is a laminate consisting of plies, each with
unidirectional fibers, stacked in a sequence such that the resulting structure has the
required combination of properties. An example is [0/ 45/90]s laminate, which is
quasi-isotropic, i.e., it has directionally independent average elastic moduli in its
mid-plane. Many other laminate configurations are possible, but often the number
of ply orientations are kept to three, and the most used ones are 0 , 45 , 45 , and
90 . From a fatigue point of view the assessment of laminates is remarkably
simple. First, irrespective of other ply orientations, the presence of 0 -plies
provides region I, which lies as a scatter-band about the fiber failure strain, as in
the unidirectional on-axis loading case as well as in the cross-ply laminate case.
The fatigue limit for any laminate is determined by the first cracking mechanism.
Therefore, if the 90 -ply orientation is present in a laminate, the strain at which the
first transverse cracking occurs will determine the fatigue limit. This strain value is
affected by the so-called ply constraint, i.e., the ratio of the transverse ply modulus
to the axial modulus of the constraint-providing plies as well as the thickness
ratio of the cracking plies to the constraining plies. For further discussion of the
constraint effect see [17] and Chapter 3. The progressive fatigue damage, repre-
sented by region II, appears as a sloping scatter band, starting at a low number of
cycles (102–103) and asymptotically approaching the fatigue limit at a high number
of cycles (106–107). What remains to determine now is the slope of the region II
band (or line). Obviously, a life prediction model accounting for the (sub-critical)
progressive fatigue damage would be needed to predict this slope. Let us examine
some test data to get some insight into this slope.
Figure 7.35 shows the fatigue-life diagram for a glass/epoxy [0/ 45/90]s lamin-
ate under tension–tension loading along the 0 -direction with data from Hahn and
Kim [18]. To construct the diagram, the failure strain of fibers (same as that of the
264 Damage mechanisms and fatigue-life diagrams

ec
0.015

0.010
emax

ed.1.=0.0046
0.005

0 2 4 6
lg N

Figure 7.35. Fatigue-life data for a glass/epoxy [0/ 45/90]s laminate under tension–tension
loading along the 0 -direction (data from [18]). The fatigue-life diagram is superimposed
on the data.

0.0100
ec

0.0075
emax

ed.1.=0.0046
0.0050

0.0025
0 2 4 6
lg N

Figure 7.36. Fatigue-life data for a carbon/epoxy [0/45/90/452/90/45/0]s laminate


under tension–tension loading along the 0º-direction [19]. The fatigue-life diagram is
superimposed on the data. Reprinted, with kind permission, from Fatigue of Filamentary
Composite Materials, copyright ASTM International, 100 Barr Harbor Drive, West
Conshohocken, PA 19428.

laminate) is taken to place the scatter-band of region I. Since the failure strain data
were not available to determine the failure probabilities, the scatter-band was
drawn based on other similar data regarding fiber failure strain. The fatigue limit
was placed at 0.46% strain based on information concerning the strain at which
transverse cracking was observed. Region II has been placed around the fatigue-
life test data. It is noted that the scatter-band of region II is approximately straight
(or negligibly curved) and meets the region I band at 102–103 cycles. The lower end
of the region II band is at approximately 107 cycles.
Another data set to consider is from [19] for a carbon/epoxy [0/ 45/90]s
laminate under tension–tension loading along the 0 -direction (see Figure 7.36).
The procedure for constructing the fatigue-life diagram is the same as that just
described related to Figure 7.35.
From the two examples of fatigue-life diagrams in Figures 7.35 and 7.36 it is
remarkable that the conceptual framework these diagrams represent is a powerful
means of interpreting and estimating fatigue-life of composite laminates. Under
the restriction of tension–tension cycling along a principal direction, usually a
7.8 Fatigue-life prediction 265

symmetry axis of the given laminate, the diagram can be constructed by looking
for simple indicators of its characteristic features. If, for example, the laminate has
a 0 -ply along which the cyclic tension load is applied, then region I will exist,
providing a flat scatter-band around the failure strain of the composite (which
equals the failure strain of fibers). The next thing to look for is the ply orientation
in the laminate that makes the largest acute angle with the loading axis. In the two
examples above, that angle was 90 . The ply with this orientation starts the fatigue
process by initiating cracks along its fibers. A good approximation of the fatigue
limit is the axial strain at which this cracking occurs. Methods for estimating this
strain have been discussed in Chapter 6. Once the upper limit (failure strain) and
the lower (fatigue) limit of the laminate have been marked in the fatigue-life
diagram, the progressive fatigue mechanism is captured between those limits.
One can model the progressive mechanism and predict fatigue-life, as we shall
discuss in the next section, or simply get a good approximation to it by drawing a
straight line (or scatter-band) going from the point on the fatigue limit line at
selected number of cycles (106–107) to the region line (or scatter-band) at a low
number of cycles. A good guideline for taking the low cycle number is 102 for
glass/epoxy and 103 for carbon/epoxy.

7.8 Fatigue-life prediction

Fatigue-life prediction of composite materials has suffered from concepts and


methodologies developed for metal fatigue that are not quite relevant. Except in
a few cases, such as delamination growth, the analysis of single crack initiation,
propagation, and unstable growth to failure does not apply to composite fatigue.
Thus, fracture mechanics by itself has little use in composite damage. The associ-
ated methodologies of stress intensity factor threshold, the so-called Paris Law,
residual strength, etc. do not help in assessing fatigue damage tolerance and
durability of composite structures.
The classical (pre-fracture mechanics) approaches to metal fatigue, typified by
fatigue-life evaluation based on empirical S-N curves, are just as lacking of useful-
ness for composite fatigue. The obvious fact that fatigue mechanisms are local,
and therefore dependent on microstructure, should suggest that homogenizing a
composite and describing fatigue in terms of average stresses cannot correctly
describe the driving impetus for composite damage. As argued above in explaining
the construction of the fatigue-life diagram, strain correctly describes the limiting
conditions for fatigue. The remarkable success of the fatigue-life diagram
in evaluating the roles of constituents and interfaces between them, as well as in
clarifying the impact of material properties, should justify moving away from the
pure empiricism of the S-N diagrams. The misconception generated by the plot-
ting of strain in the fatigue-life diagram that it requires strain-controlled testing is
indeed unfortunate. The test data are required to be generated in the completely
traditional way, i.e., in load control, while conversion of the average stress to
266 Damage mechanisms and fatigue-life diagrams

the average strain is to be done for plotting the data in the fatigue-life diagram.
The result measured in strain (such as strain corresponding to a given number of
cycles) is to be converted back to stress for use in design.
The power of the conceptual framework of the fatigue-life diagram also lies in
guiding mechanisms-based modeling of life prediction. The diagram assures that
the right mechanisms are addressed in modeling. Needless to say, mechanisms-
based modeling is the only rational approach to fatigue in composites. The
alternative of empiricism is unreliable and costly in the long run.
In the following we shall illustrate mechanisms-based fatigue-life modeling for
cross-ply laminates. So far this laminate configuration is the only one treated by
the approach. Its success, as we shall see below, should provide incentive for
treating more general laminates.

7.8.1 Cross-ply laminates


In Section 7.7.2 above, we discussed the fatigue damage mechanisms and con-
struction of the fatigue-life diagram for cross-ply laminates. To model the pro-
gressive damage in region II let us begin by reviewing the quantification of the
transverse cracking mechanism. Figure 7.37 shows the variation of the transverse
crack density (number of cracks per unit axial length) with the number of cycles
(on logarithm scale) for different levels of maximum load. The characteristic
feature of the damage evolution is that the crack density increases exponentially
at first, followed by the rate of increase decreasing and approaching the saturation
state of zero rate. The level to which the crack density saturates appears to depend
on the load level.
The first step in the life prediction procedure would be to predict the observed
transverse cracking behavior under fatigue. This requires predicting the crack
density at the first application of the maximum load and, next, determining
the increase in crack density with the number of cycles having the same load
excursion. Transverse cracking under monotonic loading was treated in Chapter 6.
Here we will focus on crack multiplication in repeated loading.
Figure 7.38 depicts the crack multiplication process. On first application of the
cyclic load, transverse cracks of a certain spacing form. On next and subsequent
applications of the same load, another crack between the previously formed
cracks would be possible for one of two (assumed) reasons: (1) a flaw exists in
the transverse plies that initiates a new crack, which grows and becomes identical
to the previous cracks, and (2) the stress state between the pre-existing cracks
changes with load cycling and becomes favorable at certain cycles to produce a
new crack. In each case we must identify the mechanism of irreversibility from
one load cycle to the next in order to model damage accumulation leading to
formation of a new crack.
In the first assumed case, we would need some knowledge about the flaws,
possibly induced by the manufacturing process. The size and spatial distributions
7.8 Fatigue-life prediction 267

50
85%
66%
53%
40 35%
28%
Crack density l (in–1)

30

20

10

0
0 1 2 3 4 5 6 7
Cycles (Log n)

Figure 7.37. Density of transverse cracks in a [0/902]s laminate plotted against the log number
of tension–tension cycles in the axial direction. The data points are for different maximum
load values indicated as percentages of the ultimate tensile strength of the composite [20].
Reprinted, with kind permission, from Springer Science+Business Media: Appl Compos
Mater, Vol. 3, 1996, pp. 391–406, X.X. Diao, L. Ye and Y.W. Mai.

Figure 7.38. Transverse cracks produced by the first application of load (left); generation
of a new crack midway between the pre-existing cracks on repeated application of the load
(middle); and completely grown crack at certain number of load applications (right).

of the flaws are likely to be random, requiring probability distributions for their
description. The initiation of cracks from flaws and their propagation with load
cycling need the presence of irreversibility, which is likely to come from the
inelasticity in the matrix (fibers are usually elastic) and/or microcracking in the
crack-tip region. In this scenario of transverse crack progression the modeling
effort will involve statistical and numerical simulation.
The second assumed case relies on stress enhancement in transverse plies in the
region between cracks without resorting to flaws. In this case, the initiation of new
transverse cracks can be assumed to occur when the axial normal stress in the
transverse plies reaches a critical value. Figure 7.39 shows the distribution of the
three stress components in the longitudinal section of the transverse plies as
268 Damage mechanisms and fatigue-life diagrams

90º
x


60 (MPa)
(1)
s xx 50
40
30
20
Crack
Crack 10
x
–3 –2 –1 1 2 3
–10 4
–4 (1)
–20 s zz (interfacial)
s (1)
xz (interfacial)
–30
–40

Figure 7.39. Axial distribution of stresses in the 90 -plies under an axial tensile load
(from [22]). Of interest is the axial normal stress, which is assumed to be constant in the
z-direction. Reprinted, with kind permission, from Damage Mechanics of Composite Materials,
J.A. Nairn and S. Hu, Matrix microcracking, pp. 187–243, copyright Elsevier (1994).

calculated by variational analysis [21, 22]. These stresses result from interaction
between cracks. As seen in Figure 7.39, the axial normal stress attains a maximum
midway between cracks and its value reduces from its pre-crack (constant) value.
Thus, in the absence of flaws, a new crack can form midway between cracks if the
maximum stress there exceeds a critical value. However, since this stress is lower
than the critical value at which the previous cracks formed, new cracks can only
form if the applied load is increased. In a cyclic load of constant amplitude, new
cracks are therefore not possible, if the assumption of no flaws still holds.
The conclusion has to be that in this scenario of transverse cracks in an elastic
composite, damage progression under cyclic loads cannot be achieved.
If the modeling of damage progression in cross-ply laminates is pursued without
entering flaws in the model, then the only plausible place for irreversible mechan-
isms is the interface between the 90 -plies and the 0 -plies. It can be argued that a
transverse crack approaching this interface is bound to cause damage to the
interface in some form or another because of the intense stress field that accom-
panies the crack front. The most likely damage is cracking in the interfacial plane
(delamination). Figure 7.40 depicts this scenario, showing delamination of length
2l at the crack fronts that are 2s distance apart. Two possible sub-scenarios are
7.8 Fatigue-life prediction 269

2l

2s

Figure 7.40. Transverse cracks of spacing 2s in a cross-ply laminate with delamination


emanating from crack fronts and extending a distance l on either side of the transverse
crack. The delamination growth is assumed to be caused by the cyclic axial tension applied
to the laminate.
Max. stress in 90 degree ply (MPa)

Max. stress in 90 degree ply (MPa)

35 300

30 250
25 s–l l
200
s–l l s
20
s 150
15
100
10

5 50

0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
l/s l/s

Figure 7.41. The variation of the maximum axial normal stress between two transverse cracks
is shown when delamination at crack fronts exists. The stress reduces with increase in the
delamination half-length l when the delamination surfaces are traction free (left) and it
increases when a shear stress acts between the delamination surfaces (right).

now possible, one where the delamination surfaces are traction free, and the other
where frictional sliding between the delamination surfaces can take place. In [23]
stress analysis was conducted for transverse cracks and delamination with a
variational mechanics method to estimate the stresses in the 90 -plies between
the transverse cracks. A cubic variation of the shear stress along the delamination
length 2l was assumed and the calculated axial normal stress then showed an
increase in its maximum value with l. Figure 7.41 shows the axial stress maximum
value for traction-free delamination and when a shear stress acts between the
delamination surfaces.
As seen in Figure 7.41, if delamination surfaces are assumed traction free,
then the right conditions would not be present for new cracks to form between
cracks formed in the first load cycle. The incentive for crack formation would
indeed decline as the delamination crack propagates under cyclic loading. On
270 Damage mechanisms and fatigue-life diagrams

2.0

1.5
Crack density (/mm)

1.0

0.5
350 MPa
400 MPa
500 MPa

0.0
0.0 0.2 0.4 0.6 0.8
Number of cycles (106)

Figure 7.42. Transverse crack density in a cross-ply laminate predicted by Akshantala and
Talreja [24] at different cyclic load levels. Note the tendency for saturation at different levels
depending on the load level. Reprinted, with kind permission, from Mater Sci Eng A,
Vol. 285, A micromechanics-based model for predicting fatigue-life of composite laminates,
pp. 303–13, copyright Elsevier (2000).

the contrary, if delamination surfaces are engaged by asperities and/or by


compressive normal stress on the surfaces, then frictional sliding between the
surfaces will result, giving rise to a shear stress. This stress alters the redistri-
bution of stresses between the axial and transverse plies, resulting in the
increase of the maximum axial normal stress as indicated in Figure 7.41. The
number of cycles beyond the first cycle that will elevate this stress to a critical
value for crack formation is then the cycles needed to increase the delamination
length by fatigue growth.
Based on the assumption of frictional sliding of the delamination surfaces,
Akshantala and Talreja [24] developed a procedure by which the transverse crack
density under cyclic loading could be determined. Figure 7.42 shows examples of
the predicted crack density variation with the number of cycles. Note the crack
density variation displays saturation to different levels depending on the maximum
stress in accordance with the experimental data in Figure 7.37, except the initial
exponential rise in the crack density. Since the stress analysis model is for inter-
acting transverse cracks the initial exponential increase in noninteracting cracks
cannot be predicted by the model.
Having predicted the crack density increase with load cycles, Akshantala
and Talreja [24] proceeded to use these data to predict fatigue-life. The
assumption was made that in the progressive damage represented by region
II of the fatigue-life diagram for cross-ply laminates (see Figure 7.34), a certain
maximum crack density is attained at failure of the laminate. This crack
density denoted f lies between the maximum achievable crack density c under
static loading and the minimum crack density fpf at the initiation of multiple
cracking, commonly called the first ply failure (a misnomer since the ply cracks
7.8 Fatigue-life prediction 271

ec
hc hf = A logNf + B

e
Crack density (hf)

A
1
efpf
hfpf

102 106
Cycles to failure (log Nf)

Figure 7.43. The assumed variation of the transverse crack density at failure in fatigue plotted
against the fatigue-life of a cross-ply laminate. The upper limit to this crack density is the
maximum saturation crack density under static load and the lower limit is the minimum
crack density at initiation of multiple cracking. The composite strains corresponding to the
two limits are also indicated. Reprinted, with kind permission, from Mater Sci Eng A,
Vol. 285, A micromechanics-based model for predicting fatigue-life of composite laminates,
pp. 303–13, copyright Elsevier (2000).

but does not fail). The variation of f with failure load cycles was assumed to
be as depicted in Figure 7.43. The equation describing this variation, f = A
log Nf + B, has empirical constants A and B, which are determined by using
the minimum and maximum values of crack densities and their corresponding
number of cycles. The extreme values of the fatigue cycles are the beginning
and end of the progressive damage, i.e., region II. In Figure 7.43, these values
are shown as 102 and 106 cycles for illustration. As suggested at the end of
Section 7.7, a good approximation of the beginning of region II is 102 cycles
for glass/epoxy and 103 for carbon/epoxy composites. Note that the assumed
process of fatigue failure is not that it comes from the transverse cracks of a
certain density, but that the crack density increases from its first-load value to
that crack density. The failure of the laminate must involve delamination
growth and fiber failures. The transverse crack density is simply assumed to
scale with fatigue-life in the assumed way.
Figure 7.44 depicts the procedure for fatigue-life prediction using the calcu-
lated crack density from the stress analysis model [23] and the assumed
relationship of the crack density to failure cycles as displayed in Figure 7.43.
At a given cyclic load, the fatigue-life is given by the minimum number of
cycles until the crack density equals the limiting value. Thus, the graphical
method for determining this value is to find the point of intersection of the
crack density increase curve and the straight line describing the limiting
crack density (see Figure 7.44). The number of cycles corresponding to the
intersection point is then the fatigue-life at that load level. The fatigue-life
prediction thus obtained is compared with the test data for a cross-ply lamin-
ate in Figure 7.45.
272 Damage mechanisms and fatigue-life diagrams

3.0

hf = – 0 .425 Nf + 3.35
hc
2.5
Crack density (/mm)

2.0

1.5

0.6%
1.0 0.65%
hfpf
0.7%
0.76%
0.5
0.9%
1.0%
0.0 1.2%
100 101 102 103 104 105 106
Number of cycles

Figure 7.44. The crack density data points represent the initial (first load) values and their
increase with cyclic loading at different load levels (indicated by the corresponding first-
cycle peak strains). The crack density relationship to failure cycles, schematically described
in Figure 7.43, is also shown with the calculated end-values of crack density. Fatigue-life
at a given load level is given by the cycles at intersection of the calculated crack density
with its failure value.

2.5

2.0
Maximum applied strain (%)

Region 1

1.5

1.0
Region II

0.5 Fatigue limit

Experiment data
Model
0.0
100 101 102 103 104 105 106 107
Cycles to failure

Figure 7.45. The predicted fatigue-life of a glass/epoxy [0/90]s laminate compared with test
data from [25].
7.9 Summary 273

7.8.2 General laminates


In Chapter 6 we treated the damage progression of a broad class of laminates
consisting of plies in multiple orientations subjected to axial tensile loads. From
the models described there one can calculate the load at initiation of multiple
cracking and crack densities resulting at the first application of the maximum load
in the given cyclic load. For further progression of damage under repeated
application of that load one must ask: what is the mechanism responsible for
irreversibility that accumulates damage with each load cycle? For cross-ply lamin-
ates we addressed this question and described the model that incorporates the
answer to the question. There is every reason to believe that in laminates with
multiple off-axis ply orientations the delamination occurring at the off-axis crack
fronts conducts frictional sliding and thereby provides the needed irreversibility
(energy dissipation). For cross-ply laminates the stress analysis conducted in [21]
for transverse cracks with delamination cannot simply be extended to multiple
crack orientations. In fact even without delamination the stress analysis of cross-
ply laminates is not easily extended to the multiple crack orientation case.

7.9 Summary

This chapter has provided a systematic conceptual framework for interpretation


of the fatigue process in composite materials. No effort has been made to give
a comprehensive exposition of the vast literature on the subject. Instead, emphasis
has been placed on understanding of the physical mechanisms underlying fatigue
and incorporating this in a systematic way in the representation called fatigue-
life diagrams. These diagrams provide a healthy departure from the empiricism
dominant in the fatigue literature. They also generate bases for material selection
and give useful guidelines for mechanisms-based modeling.
The field of fatigue damage modeling and life prediction needs a great deal of
further work. The challenges of addressing the mechanisms of fatigue damage
accumulation have discouraged most research efforts that have instead taken the
path of empirical approach. In fact even in the century old field of metal fatigue
empirical approaches are common. The so-called Paris Law of cyclic crack growth
is basically a curve fit to the observed data. However, its simplicity has attracted
practical use but has impeded advances in fundamental understanding. It is
therefore not surprising that the composites community has so far also opted
for empiricism. It is worth noting that because of the large number of parameters
in composite materials (constituent properties, ply orientations, woven and other
complex architectures, etc.) the empirical path is highly inefficient. As demon-
strated by the fatigue-life diagrams, focusing on the essential mechanisms allows
one to transcend the apparent complexity of the composite microstructure and
fiber architecture.
274 Damage mechanisms and fatigue-life diagrams

In the next chapter we shall outline the areas in composite fatigue that must
receive attention by the composites community and must be supported by govern-
ment and industry.

References

1. R. Talreja, Fatigue of composite materials: damage mechanisms and fatigue-life


diagrams, Proc R Soc London A, 378 (1981), 461–75.
2. C. K. H. Dharan, Fatigue failure mechanisms in a unidirectionally reinforced compos-
ite material. In Fatigue in Composite Materials, ASTM STP 569. (Philadelphia, PA:
ASTM, 1975), pp. 171–88.
3. J. B. Sturgeon, Fatigue and creep testing of unidirectional carbon fiber
reinforced plastics. In Proceedings of the 28th Annual Technical Conference of the
Society of the Plastics Industry. (Washington, DC: Reinforced Plastics Division, 1973),
pp. 12–13.
4. J. Awerbuch and H. T. Hahn, Fatigue and proof-testing of unidirectional graphite/
epoxy composite. In Fatigue of Filamentary Composite Materials, ASTM STP 636, ed.
K. L. Reifsnider and K. L. Lauraitis. (Philadelphia, PA: ASTM, 1977), pp. 248–66.
5. J. B. Sturgeon, Fatigue Testing of Carbon Fibre Reinforced Plastics, Technical Report,
Royal Aircraft Establishment, Farnborough (1975).
6. R. B. Croman, Tensile fatigue performance of thermoplastic resin composites
reinforced with ordered Kevlar® aramid staple. In Proceedings of the Seventh Inter-
national Conference on Composite Materials, vol. 2, ed., Y. Wu, Z. Gu, and R. Wu.
(Oxford: International Academic Publishers, 1989), pp. 572–7.
7. E. K. Gamstedt and R. Talreja, Fatigue damage mechanisms in unidirectional carbon
fibre-reinforced plastics. J Mater Sci, 34 (1999), 2535–46.
8. E. K. Gamstedt and S. Östlund, Fatigue propagation in fibre-bridged cracks in unidir-
ectional polymer-matrix composites. Appl Compos Mater, 8 (2001), 385–410.
9. E. K. Gamstedt, Effects of debonding and fiber strength distribution on fatigue-
damage propagation in carbon fibre-reinforced epoxy. J Appl Polym Sci, 76 (2000),
457–74.
10. R. Talreja, A conceptual framework for interpretation of MMC fatigue. Mater Sci Eng
A, 200 (1995), 21–8.
11. B. F. Sørensen and R. Talreja, Analysis of damage in a ceramic matrix composite.
Int J Damage Mech, 2 (1993), 246–71.
12. B. F. Sørensen, J. W. Holmes, and E. L. Vanswijgenhoven, Does a true fatigue limit
exist for continuous fiber-reinforced ceramic matrix composites? J Amer Chem Soc, 85
(2002), 359–65.
13. Z. Hashin and A. Rotem, A fatigue failure criterion for fiber reinforced materials.
J Compos Mater, 7 (1973), 448–64.
14. A. Rotem and Z. Hashin, Fatigue failure of angle ply laminates. AIAA J, 14 (1976), 868–72.
15. R. D. Jamison, K. Schulte, K. L. Reifsnider, and W. W. Stinchcomb, Characterization
and analysis of damage mechanisms in tension–tension fatigue of graphite/epoxy
laminates. In Effects of Defects in Composite Materials, ASTM STP 836. (Philadelphia,
PA: ASTM, 1984), pp. 21–55.
References 275

16. G. C. Grimes, Structural design significance of tension–tension fatigue data on com-


posites. In Composite Materials: Testing and Design (Proc. 4th Conference), ASTM STP
617. (Philadelphia, PA: ASTM, 1977), pp. 106–19.
17. R. Talreja, Transverse cracking and stiffness reduction in composite laminates.
J Compos Mater, 19 (1985), 355–75.
18. H. T. Hahn and Y. Kim, Fatigue behavior of composite laminate. J Compos Mater,
10 (1976), 156–80.
19. J. T. Ryder and E. K. Walker, Effect of compression on fatigue properties of a quasi-
isotropic graphite/epoxy system. In Fatigue of Filamentary Composite Materials, ASTM
STP 636. (Philadelphia, PA: ASTM, 1977), pp. 3–26.
20. X. X. Diao, L. Ye, and Y. W. Mai, Simulation of fatigue performance of cross-ply
composite laminates. Appl Compos Mater, 3 (1996), 391–406.
21. Z. Hashin, Analysis of cracked laminates: a variational approach. Mech Mater, 4,
(1985), 121–36.
22. J. A. Nairn and S. Hu, Matrix microcracking. In Damage Mechanics of Composite
Materials, ed. R. Talreja. (Amsterdam: Elsevier Science, 1994), pp. 187–243.
23. N. V. Akshantala and R. Talreja, A mechanistic model for fatigue damage evolution in
composite laminates. Mech Mater, 29 (1998), 123–40.
24. N. V. Akshantala and R. Talreja, A micromechanics based model for predicting
fatigue-life of composite laminates. Mater Sci Eng A, 285 (2000), 303–13.
25. C. J. Jones, R. F. Dickson, T. Adam, H. Reiter, and B. Harris, The environmental
fatigue behaviour of reinforced plastic. Proc R Soc London A, 396 (1984), 315–38.
8 Future directions

In Chapter 1 we discussed the durability assessment of composite structures, the


overall goal for the subject of this book. As outlined there in Figure 1.1, the
mechanisms of damage and their effects on deformational response constitute
the main thrust of the field of damage mechanics, which is at the core of durability
assessment. After discussing the physical nature of damage observed experimen-
tally in Chapter 3, the next two chapters treated the two main approaches in
damage mechanics – micro-damage mechanics (MIDM) and macro-damage
mechanics (MADM), both aimed at predicting deformational response at fixed
damage. Damage evolution was treated in Chapter 6, while Chapter 7 was devoted
to fatigue, a subject that requires special attention due to the conceptual difficul-
ties it poses.
In closing the book we wish in this chapter to review what has been
achieved and what directions the field of damage and failure of composite
materials should pursue to further advance toward durability assessment and
beyond.

8.1 Computational structural analysis

Obviously, complex structural geometries require computational structural


analysis. The analytical modeling of damage initiation and evolution, and its
effects on deformational response of composite laminates, discussed in previous
chapters, were developed for idealized simple cases. Direct application of these
models is limited to structures with simple geometry and loading conditions. For
complex geometries, such as an airplane wing or a wind turbine blade, usually
subjected to multi-axial mechanical loads, and possibly combined with thermal
and moisture environments as well as manufacturing-induced residual stresses,
computational approaches are inevitable. In industry, one often uses commercial
software, e.g., ANSYS, ABAQUS, and NASTRAN, and the obvious need is to
integrate damage and failure analyses into these codes. Efforts have been made
to attempt some simple test cases where FE analysis of composites is combined
with damage using failure criteria [1]. A series of World Wide Failure Exercises
(WWFE) [2–4] have been conducted to compare several composite failure models
with experimental data and provide guidance for their usage in composite design.
8.1 Computational structural analysis 277

INPUT
Geometry, laminate configuration, material
properties and loading conditions

PRELIMINARY STRESS ANALYSIS

Create geometric Apply loading &


Meshing Stress analysis
model service conditions

DAMAGE ANALYSIS

Identify regions where Predict damage Evaluate stiffness


damage might have initiation & properties of
developed evolution damaged regions

UPDATED STRESS ANALYSIS

Update stiffness of damaged regions Perform stress analysis of whole structure again

OUTPUT
Failure characteristics, stress-strain response,
deformation behavior, life and durability

Figure 8.1. Flowchart illustrating computational stress analysis of a composite structure in


combination with damage analysis.

While previous exercises focused on the ultimate failure, the ongoing WWFE-III
offers opportunities to examine the initiation and progression of sub-critical events
and their effect on the mechanical response. Yet no comprehensive code exists
which can combine stress analysis with damage prediction. The current codes
predict structural failure based on lamina-based failure criteria and not on the basis
of sub-critical damage such as ply cracking. To improve this situation, the authors
are currently developing a user subroutine (UMAT) in ABAQUS with an ability to
dynamically update the stiffness properties of a region using the energy-based
fracture criterion described in Chapter 6. The aim is to predict the initiation of
ply cracking and its evolution, as well as the resultant stiffness properties, using the
synergistic damage mechanics approach described in Chapter 5.
While Figure 1.1 outlined the overall methodology for durability assessment,
Figure 8.1 here illustrates the procedure for performing computational stress
analysis of a composite structure in combination with damage analysis. At first,
stress analysis is conducted assuming no damage is present in the whole structure.
At this point the necessary data input includes the geometry of the structure, the
configuration (layup, thicknesses of plies, etc.) and material of the composite
278 Future directions

laminate, and the loading and service conditions. Then, based on the stress
pattern, regions are identified which are potential sites for damage under the
prescribed loading and service conditions. These regions include any stress con-
centrators and could be based on previous design experience. In this step, possible
damage mechanisms are also identified. For instance, the region close to a hole or
cut-out in the structure is expected to develop damage in the form of ply cracks
and delamination. Next, using the preliminary stress states, damage initiation and
its evolution are predicted with the methods and analyses treated in earlier
chapters. The damage prediction step may involve a substructural analysis and
be carried out over multiple length scales. The multiscale modeling aspect will be
discussed later in the present chapter. The stiffness properties of damaged regions
are then updated using the analytical predictions, and a new stress analysis is
performed at the same applied load levels. On increased loading, the same iterative
process is repeated until the analysis predicts failure based on the design criteria.
Depending on the purpose and need, the output of the computational structural
analysis may include detailed failure characteristics, overall stress–strain response,
deformation characteristics (deflections, vibrational frequencies, etc.), as well as
damage tolerance and life prediction of the structure.
The main challenge in any computational scheme is to properly incorporate
damage and its effects. It is not uncommon to find works where damage with all its
complexities is reduced to a parameter D, which is then arbitrarily assigned a value
between 0 and 1, often based on computational convenience and seldom account-
ing for the physics behind damage. Much progress is still needed in taking the
theoretical advances in damage mechanics into computational design procedures.
One of the main hurdles is properly conducting multiscale computational analysis
of damage. In the following we shall address issues in multiscale modeling of
damage, which present some fundamental challenges arising in the computational
treatments of composite failure.

8.2 Multiscale modeling of damage

The computational power available today has given impetus to the so-called multi-
scale modeling, which follows an intuitive notion that the physical phenomena
occurring at lower length scales determine the material response at higher scales.
Fiber-reinforced composites and, more generally, heterogeneous solids appear to
be natural candidates for this modeling idea. Thus, numerous studies have resulted
to address deformation, damage, and failure of these materials and to some extent
also their nonmechanical properties. A vast literature on multiscale modeling exists,
and a recent collection of works [5] covers a wide range of such treatments.
Since in a heterogeneous solid the microstructural entities (heterogeneities)
are embedded in the matrix, a given property (e.g., mechanical or thermal
response characteristic) must be defined at a scale much larger than the charac-
teristic size of the heterogeneities. This leads to the suggestion that, at a chosen
8.2 Multiscale modeling of damage 279

Figure 8.2. Durability analysis procedure for composite structures.

t t t
u u1 u2

Figure 8.3. Heterogeneous solid in pristine state (left) and two damage states, type 1
(middle) and type 2 (right).

scale, a representative volume element (RVE) exists such that by appropriate


averaging over this volume the property of interest can be computed. The RVE
idea is a necessary fundamental concept in any proper multiscale treatment. The
reader is referred to texts in micromechanics such as that by Nemat-Nasser and
Hori [6] for in-depth treatment of the RVE concept and the associated aver-
aging schemes.
Denoting the RVE scale as mesoscale, and the scale of microstructural entities
as microscale, the three-level micro-meso-macro hierarchy forms the basis of a
multiscale modeling treatment. In fact if the “microstructure” in a heterogeneous
material contains nano-scale reinforcements (e.g., particles, fibers or tubes), then
further differentiation of the lower-end scale to nano- and micro-levels can be
made. While the hierarchical multiscale modeling is a viable approach for micro-
structures that are “stationary” (Figure 8.3), i.e., they do not change their charac-
teristic size under a loading impulse, it is not a given that the approach will remain
valid when energy dissipative mechanisms permanently alter the microstructure.
Our concern here is multiscale analysis of the energy dissipative mechanisms
underlying damage in composite materials for the purpose of assessing structural
integrity and durability, and for this case further examination of the multiscale
approach seems warranted.
In the context of computational structural analysis described above in
Section 8.1 and in a previous review of this subject [7] we refer to Figure 8.2,
280 Future directions

which was also discussed in Chapter 1. Here we draw attention to the damage
mechanics part of the scheme. As indicated, micro-meso-macro multiscale model-
ing is at the core of this analysis. The three-scale differentiation when damage
exists does not necessarily coincide with the multiscale differentiation adopted
before damage. This critical point was treated in [8], from which the following
discussion is drawn.

8.2.1 Length scales of damage


Let us consider three RVEs depicted in Figure 8.3 to discuss characteristic scales
of damage and their relation to the scales of microstructure. The RVE to the left in
the figure is for pristine (undamaged) composite. When the bounding surface of
this RVE is subjected to a prescribed traction t, the combined deformation of the
matrix and the heterogeneities (shown symbolically as filled circles) contained
within the RVE produces a displacement u of points on the bounding surface.
This displacement field can be written as u = u0 + du0, where u0 represents the
deformation of the matrix in the absence of fibers and du0 is the perturbation in it
caused by the heterogeneities in the absence of damage. The middle RVE in
Figure 8.3 illustrates a damage scenario – let us call it type 1 damage – where
some of the heterogeneities (e.g., inclusions) have partially or fully separated from
the matrix. The interfacial cracks, thus formed, then perturb further the deform-
ation field within the RVE if the traction t on the RVE surface is sufficient to
activate the cracks (i.e., displace the crack surfaces). The displacement of points on
the bounding surface is now u1. Finally, the RVE to the right in the figure depicts a
damage scenario – type 2 damage – where the cracks within the RVE are restricted
to the matrix and are geometrically unconnected to the heterogeneities. The
displacement response to surface traction t on the RVE in this case is
denoted u2. The perturbation field du0 can, in principle, be determined by know-
ledge of the properties of matrix and heterogeneities, as well as configuration
variables such as size, shape, and spacing of heterogeneities. Depending on the
model employed, one may use limited configuration information such as volume
fraction of heterogeneities, or more enriched information such as statistical cor-
relation functions that describe their relative size and placement. A thorough
treatment of the morphological characterization of microstructures is found
in [9], while [10] gives an extensive review of models that aim at carrying the
morphological information to RVE level averages.
The models in general have multi-level features and often rely on homogeniza-
tion concepts. In the absence of damage, the models may be characterized as
“stationary” microstructure models, i.e., the microstructure configuration remains
unchanged under application of RVE surface traction t. In calculating the RVE
surface displacement response to the prescribed traction t in the presence of
damage, one can take two alternative approaches. One approach is to homogenize
the matrix and stationary microstructure and embed the damage entities (e.g.,
cracks) in the homogenized composite. The displacement response can then be
8.2 Multiscale modeling of damage 281

written as ui = u + diu, where diu results from the perturbation in the displace-
ment field of the homogenized composite caused by damage (i = 1 or 2 for
damage type 1 or type 2, respectively). This approach is common and has many
versions, a familiar one being the Mori–Tanaka estimation procedure [11]. It is
noted that the Mori–Tanaka procedure is a homogenization method for station-
ary microstructures, i.e., for undamaged composites. For further reference, we
shall call it the homogenized microstructure (HM) approach. The other approach
is to retain the discrete nature of the stationary microstructure in estimating the
RVE displacement response. In this approach, called here the discrete microstruc-
ture (DM) approach, the displacement response can be written as ui = u0 + diu0,
where diu0 results from combined perturbation in the displacement response of the
matrix caused by the stationary microstructure and damage, the subscript index i
referring still to the damage type.
The two approaches to estimating the RVE surface displacement, just
described, are approximate and will generally yield different results. In the HM
approach, the explicit association of a damage entity to the microstructure is lost,
since the microstructure in which the damage entity resides has been homogenized,
while in the DM approach the heterogeneities as well as damage entities are
explicitly present. Thus, in the HM approach, damage entities of type 1 and type 2
are both surrounded by the homogenized microstructure, thereby their association
with the microstructure (as to how the microstructure affects their initiation) is
lost. In this sense, the HM approach is insensitive to which of the two types of
damage is treated. On the other hand, in conducting the DM approach a signifi-
cant difference exists depending on whether damage of type 1 or type 2 is
considered. Since the type 1 damage is geometrically tied to the stationary micro-
structure, the perturbation caused by it in the local fields can be analyzed by
viewing it as a modification to the perturbation induced by the heterogeneities.
This would not be the case for type 2 damage, as it is unconnected to the
heterogeneities but affected by them. In fact, for this reason, the HM approach
would be preferable for type 2 damage.
The considerations described above have been made for a specific purpose: to
clarify the characteristic scales associated with the micro-level (heterogeneities)
and their relation to the meso-level (RVE) scale. As noted, the micro–meso
bridging in the absence of damage is clear and unambiguous. In fact, if the matrix
has a heterogeneous structure itself, then knowing the characteristic scales of that
substructure it would be possible to homogenize it with a multiscale (sub-micro to
micro) approach. Thus, in general for stationary heterogeneities, the multiscale
modeling is plausible and systematic, at least conceptually. This is far from the
case when damage in the form of distributed internal surfaces exists. When these
internal surfaces are geometrically connected to the microstructure, such as in
type 1 damage, their characteristic micro-level scales can be deduced from those of
the heterogeneities and the micro–meso bridging is then relatively simple. How-
ever, few cases of damage in composite materials belong to this type, i.e., where
the damage entities remain connected to the heterogeneities. Although this might
282 Future directions

Debonding induces matrix cracking

(a)

Matrix cracking causes debonding

(b)

Figure 8.4. A typically observed micrograph (left) of a cross section of a unidirectional fiber
composite loaded in tension normal to fibers. To the right are depicted two plausible
scenarios underlying the picture on the left.

occur in the initial stages of damage, the damage entities formed grow away from
the heterogeneities. In the case of type 2 damage, where the damage entities have
scales unrelated to those of the heterogeneities, the resulting complexities of scales
question the viability of the hierarchical multiscale modeling, as discussed in more
detail in the following.

8.2.2 Hierarchical multiscale modeling


We shall now focus on the specific case of composite laminates in which individual
plies are of a matrix material reinforced with unidirectional fibers. Thus the
stationary microstructure here consists of fibers distributed in the matrix and
corresponds to the picture on the left in Figure 8.3 for a heterogeneous solid
in pristine state. The microstructural length scale, or micro-scale, is the size of a fiber
(radius or diameter). The next scale in the hierarchy, the meso-scale, is the RVE size,
which depends on the distribution of fibers. Assuming all fibers are straight and
parallel, then for the case of uniformly distributed fibers in the cross section a
repeating unit of a fiber embedded in a surrounding matrix, i.e., a unit cell, replaces
RVE. For nonuniform distribution of fibers, reference is made to treatments in [9]
and [10]. Simpler estimates for ply properties can be found in any of the common
texts on composite mechanics, e.g., [11]. Finally, the macro-scale for a composite
laminate is a structural scale that depends on the structural geometry.
When a composite laminate suffers damage under loading, a variety of length
scales develop that may or may not be connected with the initial length scale of
undamaged composite. Although this issue has been discussed above in general
terms, we will treat it now specifically for composite laminates.

8.2.2.1 Damage in unidirectional composites: transverse loading case


Figure 8.4 (left) shows a typical view of damage observed in the cross section of a
unidirectional fiber composite of a polymeric matrix loaded in tension normal to
8.2 Multiscale modeling of damage 283

fibers. The damage appears to be a mix of fiber/matrix debonding (type 1 damage)


and matrix cracking (type 2 damage) described above. However, a closer study of
fiber/matrix debonding [12–14] suggested that this failure mechanism could be a
consequence of cavitation-induced brittle cracking in the matrix. Type 2 damage,
on the other hand, is matrix flow-induced ductile cracking in this case. Thus, in
polymer matrix composites, damage of type 1 and type 2 could both be different
realizations of matrix failure with drastically different governing scales, as dis-
cussed below.
In Figure 8.4 (right) are shown two plausible scenarios of cracking that can
underlie the observed damage in Figure 8.4 (left). The first one, marked (a), is for
fiber/matrix debonding that progresses out of the fiber surface into the matrix.
Such matrix cracks then coalesce forming a continuous fiber-to-fiber crack. The
second scenario assumes formation of matrix cracks first, which on growing
towards fibers induce fiber/matrix debonding. This sequence of cracking would
also produce a continuous fiber-to-fiber crack. Which of the two scenarios holds
depends on the transverse loading-induced local stress states, which in turn
depend on the microstructure configuration, i.e., fiber volume fraction and distri-
butions of fiber diameter, inter-fiber spacing, etc. A discussion of the effect of local
stress states on matrix damage follows.
A cross-sectional region of a transversely loaded composite is illustrated in
Figure 8.5. For nonuniformly distributed fibers in the cross section, stress analysis
studies conducted in conjunction with the work reported in [12] suggested that for
points in the matrix close to the fiber surface, such as that indicated in the figure,
the deformation of the matrix is nearly or fully dilatational, while for points in the
matrix away from fibers, such as the other point indicated in the figure, the matrix
deformation has a significant distortional component. The mix of dilatational and
distortional deformation will depend on the constraint to deformation imposed by
the presence of fibers. For instance, for points at nearly-touching fiber surfaces
and in matrix regions that are squeezed between three fibers, the deformation will
approach the dilatational state, while at points in the matrix sufficiently away
from fibers such that the local stress perturbation induced by fibers is negligible,
the deformation state will have a high degree of distortion. For points near a fiber
surface where conditions for dilatational deformation are favorable, the same
studies [12, 13] proposed that cavitation of the polymer matrix is induced by the
hydrostatic tension (the insert in Figure 8.5(b)). The cavities thus formed expand
stably at first and become unstable when the dilatation energy reaches a critical
value. The unstable cavity growth in a region with lack of sufficient distortional
energy results in brittle cracking, which finds its way into the fiber/matrix inter-
face. The consequent debonding grows as an arc-shaped crack along the fiber
surface (Figure 8.5(b)) and subsequently diverts into the matrix along a direction
normal to the local maximum tensile stress (Figure 8.4(a)). Further progression of
damage produces the view seen in Figure 8.4 (left).
An alternative damage scenario is illustrated in Figure 8.5(c), where a point
away from fibers undergoes significant distortional deformation, which eventually
284 Future directions

(a) s

Dilatational

s Distortional

(b) (c)
s s

Distortional

s s

Figure 8.5. (a) Points in a transversely loaded heterogeneous solid with dilatational and
distortional deformation; (b) cavitation and subsequent cracking induced by hydrostatic
stress state; and (c) cracking in matrix caused by distortional flow.

localizes in shear-intensive bands, leading to ductile cracking. As the crack


growth advances to brittle regions near fibers, fiber/matrix debonding occurs
(Figure 8.4(b)). As in the other damage scenario, further progression of damage
can produce the same view seen in Figure 8.4 (left).
Modeling of the two damage scenarios depicted above involves analyses at two
different sets of characteristic scales. For the cavitation-induced brittle cracking
resulting in fiber/matrix debonding (the first of the damage scenarios, Figure 8.5
(b)), the characteristic scale at which cavitation begins is the molecular scale of the
polymer matrix. Once debonding occurs, the scale of damage is the diameter of the
fiber on whose surface the arc-shaped debond crack grows. For the second
damage scenario (Figure 8.5(c)), the sequence of deformation and failure mechan-
isms can be long and complex depending on the polymer morphology and the
stress triaxiality (defined in an appropriate way to describe the mix of deviatoric
and hydrostatic stress). These mechanisms can be classified as brittle, quasi-brittle,
or ductile to characterize the relative degree of material flow involved in the
cracking process. The characteristic scales of damage will then vary accordingly.
A large body of literature exists on models that address the deformation and
8.2 Multiscale modeling of damage 285

failure of polymers ranging from molecular level treatments to continuum mech-


anics analyses incorporating the details of the mechanisms in various explicit and
implicit ways. Review of this literature is not the purpose here, but in the context
of the multiscale damage in heterogeneous solids it is important to note that the
presence of heterogeneities alters the nature of the problem. While for an unre-
inforced polymer a multiscale modeling effort must address all scales of morph-
ology that are activated by the loading, in the case of a polymer composite,
activities are enhanced at certain scales and subdued at others, depending on the
local stress perturbation caused by the heterogeneities. Thus, the hierarchy of
scales and their relative roles are generally different in unreinforced and in
heterogeneous polymers.
Although the discussion here has been focused on polymers as matrix materials,
the inferences drawn will largely apply to reinforced ductile metals as well.

8.2.2.2 Cross-ply laminate: transverse ply cracking


Transverse ply cracking in a cross-ply laminate is an extensively studied area and
has been reviewed in previous chapters; here the focus of discussion will be the
characteristic scales of damage. When a cross-ply laminate is loaded in tension
along the longitudinal plies, the first event of damage is the transverse crack
formation in one of the two ways discussed above. The crack grows across fibers
first, and then along fibers, eventually spanning the thickness and width of the
transverse ply. Up to this point, the scales associated with the damage are as
discussed above for transverse loading-induced damage in unidirectional compos-
ites. On encountering the ply interfaces the transverse crack fronts bring about
interfacial stress perturbation, which traverses a certain distance along the longi-
tudinal plies, whence the stresses return to the pre-cracking state, unless a perturb-
ation by another crack intervenes, in which case another equilibrium stress state
results. This so-called shear lag distance is the distance on either side of the
transverse crack where reduction of the axial stress in the transverse ply prevents
another transverse crack from forming, and increased loading is needed to pro-
duce such a crack. This phenomenon of stress transfer from cracked transverse
plies to the longitudinal plies is responsible for multiple cracking in the transverse
plies. The conditions for single crack formation versus multiple cracking were first
explained in the landmark paper now commonly known as the ACK theory [15].
From the viewpoint of the characteristic scales of damage, which is what is of
concern here, the situation changes drastically when damage evolves from the
phase of single transverse crack formation to multiple transverse cracking. The
scales associated with single crack formation were discussed above. In multiple
cracking the governing scale is that of the shear lag distance along the 0/90 ply
interface. This distance is determined by the ply properties and the thickness of
plies in the two orthogonal directions. In other words, it is the homogenized ply
properties and laminate configuration that determine the scale of multiple
cracking, while it was the ply constituent properties and reinforcement morph-
ology (fiber volume fraction, fiber diameter, inter-fiber spacing, etc.) that governed
286 Future directions

Stationary microstructure

RVE for damage characterization

Evolving microstructure

Step 1 Homogenization of P
stationary microstructure

V
Step 2

P
Homogenization of
evolving microstructure
ai
nj

Continuum after homogenizing Fully homogenized continuum


the stationary microstructures Characterization of a
damage entity

Figure 8.6. Two-step homogenization of a composite body with damage is depicted. The
characterization of damage and the associated RVE are also illustrated.

formation of a single transverse crack. This fact does not seem to be fully
appreciated by most multiscale modeling efforts, which tend to treat transverse
cracking with the scales of heterogeneities (fiber size and spacing, and associated
distributions).

8.2.3 Implication on multiscale modeling: Synergistic damage mechanics


The two cases discussed above are intended to illustrate the complexity and
richness of the damage phenomena in composite materials. More cases have been
treated elsewhere [7, 8, 16, 17]. It seems clear from these studies that a hierarchical
multiscale approach for undamaged composites cannot generally be extended to
cover all cases of damage. This may suggest that multiscale modeling for the
purpose of structural integrity and durability should be approached on its own
rather than tying it to the hierarchical approach for undamaged heterogeneous
solids. Efforts in this direction were proposed and labeled as “synergistic damage
mechanics” (SDM) [18]. Since the publication of that work, a systematic demon-
stration of the viability of the approach has been made [19–23].
In Section 5.2 the SDM approach was described in detail. Here, for the sake of
completeness of the multiscale modeling discussion we include two figures from
Chapter 5. Figure 8.6 (same as Figure 5.4) shows the two-step homogenization
procedure involved in the characterization of damage. As depicted in the figure,
the stationary microstructure is homogenized first and represented by appropriate
constitutive relations. The evolving microstructure, consisting of damage entities,
8.3 Cost-effective manufacturing and defect damage mechanics 287

is homogenized next and regarded as an internal structure embedded in the


homogenized solid of stationary microstructure. The new continuum is now
represented by a thermodynamics framework in which internal state variables
characterize damage by a set of second-order tensors. The internal variables are
defined by descriptors that require averaging over a RVE. The RVE size is then
the meso scale, while the characteristic size of the damage entities is the micro
scale. The size-scales of the stationary microstructure enter separately in the step 1
homogenization.
In this scheme the micro-scale level is the single damage entity size and the
meso-scale level is the RVE size, as described above. The micro-level descriptor
is the damage entity tensor dij described in Chapter 5, while its average over the
ðaÞ
RVE, denoted Dij for a selected damage mode a is the meso-scale descriptor,
also described in Chapter 5. Because of the way the damage entity tensor is
constructed, it is possible to explicitly incorporate micro-level information into
this descriptor. In this way the influence of the surrounding heterogeneous
solid, i.e., the “microstructure,” on the damage entity can be analyzed by a
convenient means and thereby transmitted to the meso level. A parameter,
called the “constraint parameter,” has been devised to effectively accomplish
this task, either experimentally or by computational micromechanics. This has
all been described in detail in Chapter 5 for single and multiple modes of
damage. Figure 5.10 from Chapter 5 is reproduced here as Figure 8.7 for a
convenient recollection of the SDM procedure.
The multiscale modeling of damage and the accompanying SDM methodology
are the appropriate way to treat the effect of damage on the composite material
response and by extension to assess structural durability. The hierarchical multi-
scale treatment of heterogeneous solids, although suited for estimating their
overall response, cannot generally be extended to account for damage.

8.3 Cost-effective manufacturing and defect damage mechanics

A composite structure, constructed by any practical manufacturing process, is


rarely perfect. The defects induced during manufacturing can be in the fiber
architecture, e.g., fiber misalignment, irregular fiber distribution in the cross
section, and broken fibers; in the matrix, e.g., voids; and in the interfacial regions,
e.g., debonding and delaminations. Such defects must be analyzed to determine
their effects on the composite structural integrity and durability. The results of
such analyses can be used in two ways: (a) to develop acceptance/rejection criteria
for the manufactured part; and (b) to design the part to meet performance
requirements accounting for the defects. The former is what industry largely
practices today, if at all. A host of nondestructive inspection (NDI) techniques
are potentially possible to detect defects. These techniques can be applied to
implement thresholds for product quality, e.g., a maximum void volume fraction
or a maximum delamination surface area.
288 Future directions

COMPUTATIONAL MICROMECHANICS
Determine COD and constraint parameter(s)
(
κ (∆u 2 ±q n
b = κq = ) ; ( ∆u2 ) ±q n = ( ∆u2 ) +q n + ( ∆u2 )–q n
90 (∆u 2 90
8

Structural scale: Micro

EXPERIMENTAL/
COMPUTATIONAL SYNERGISTIC DAMAGE MECHANICS

Evaluate damage constants Use SDM to determine stiffness reduction in


using available data present laminate configuration [0m / ± qn /0m /2]s
for reference laminate
Structural scale: Meso
configuration [0/908 /01/2]s

STRUCTURAL ANALYSIS
Analyze overall structural response to external
loading using the reduced stiffness properties

Structural scale: Macro

Figure 8.7. Flowchart showing the multiscale synergistic methodology for analyzing
damage behavior in a class of symmetric laminates with layup [0m/ yn/0m/2]s containing ply
cracks in the +y and y layers.

The field of designing a composite part with known defects is far from mature.
What need to be developed are accurate analyses of effects of real-life defects as
well as a strategy for incorporating the results in a cost-effectiveness assessment of
the manufacturing process. As is the case, most of the cost of a part lies in the
manufacturing process. In the following, we shall first review a cost-effectiveness
assessment procedure, clarifying the different elements that make up the proced-
ure. We shall then focus on the mechanics of materials approaches for analysis of
real-life defects. For illustration we shall review some recent results on (i) elastic
property changes due to matrix voids, and (ii) effects on the propensity for crack
extension induced by interlaminar voids. In closing we shall make recommenda-
tion for future work.

8.3.1 Cost-effective manufacturing


Figure 8.8 describes the interrelated elements involved in the process of assessing
the cost-effectiveness of a composite structure with respect to its long-term per-
formance. To begin, the manufacturing process selected for a given composite
structure is described by material and geometry parameters, processing param-
eters and their time variations, as well as descriptors of machining, tooling,
joining, assembly, etc., as needed. The product resulting from the manufacturing
process is characterized in terms of the “material state” and its corresponding
8.3 Cost-effective manufacturing and defect damage mechanics 289

Figure 8.8. Procedure for the cost-effectiveness assessment of composite structures.

properties. The conventional material state description consists of constituent


properties and their relative proportions, e.g., volume fractions, and of the fiber
architecture, e.g., ply thickness, orientation, and stacking sequence in a laminate,
or fabric type (e.g., 5- or 8-harness satin), thickness, and layup in a woven fabric
composite, etc. In addition to this, the material state needs to be described by
certain defect descriptors. As we shall see, homogenized descriptions of the
constituents and defects do not suffice for the cost-effectiveness assessment of a
manufacturing process.
The defect descriptors needed would depend on the manufacturing process.
Examples of such descriptors are distributions (or other statistics) of fiber mis-
alignment angles, of void size and location, of fiber/matrix interfacial disbonds, of
delamination size and location, etc. The appropriate set of defect descriptors,
along with the conventional material state descriptors, makes up a complete
characterization of the manufactured composite material. Depending on the service
environment in which the composite structure is to function, i.e., the design
requirements imposed on the structure, the cost-effectiveness assessment will
consider the necessary material properties and their relationships with the material
and defect descriptors. A cost/performance trade-off exercise, and any iteration on
it, will result in an optimized cost-effective product. In most applications, where
long-term performance is the critical design consideration, one needs to look at
the degradation of initial (end-of-manufacturing) properties of interest under the
service environment. Thus the cost/performance trade-off will consider the
residual properties. A common approach is to consider only the initial (i.e., pre-
service) properties even for the long-term case, assuming implicitly that the
residual properties will relate to the defects (and the cost) the same way as the
initial properties do. This assumption is in fact questionable since the initial
properties may not show sensitivity to some of the material defects that may turn
out to be significant in governing the long-term properties.
There is a variety of manufacturing processes used for composite structures,
e.g., autoclave molding, liquid compression molding, resin transfer molding,
290 Future directions

filament winding, chemical vapor deposition, etc. Each of these processes pro-
duces defects in the manufactured part that are usually characteristic of that
process. The machining, joining, and assembly methods used for composite struc-
tures produce defects that are generally different from those produced during
molding, winding, and vapor deposition. For instance, the defects in the inter-
facial region between two parts will be different depending on whether the parts
were co-cured or adhesively bonded. Significant differences in the fatigue lives of
co-cured vs. bonded joints have been reported [24].
In recent years, many methods have been developed to observe manufacturing
defects in composite materials and structures by nondestructive evaluation, based
mainly on ultrasound and radiography [25] and to some extent on thermal wave
imaging [26]. In the conventional approach these methods are utilized primarily for
quality control of the manufactured product. The premise of the quality control is
often that the presence of defects is undesirable. If defects of more than certain
threshold values are found, then the part is rejected and one strives to improve the
manufacturing process. This inevitably increases the cost and can result in the
composite part not being competitive with a metallic alternative. It is important to
realize that the presence of defects is not undesirable in all cases. In fact, if some defects
are allowed, the part can be produced at a lower cost. Figure 8.9 illustrates the
dependence of strength per dollar of manufacturing cost on the defect density. As
seen in the figure, the strength decreases gradually with defect density for low
densities, and drops rapidly at high densities, while the manufacturing cost increases
rapidly at low defect densities and falls off as more defects are allowed. Thus the
strength of the part achieved per dollar of manufacturing cost increases with defect
density, up to a point, beyond which the benefit of allowing more defects decreases. It
must be remembered, however, that this situation is typical of the static strength. The
dependence of residual strength in long-term loading on the initial defect density may
show different behavior. This aspect has not been investigated sufficiently.
Figure 8.9 also suggests that we should get away from the accept/reject
approach and advance to what may be called defect engineering. More specifically,
we should engineer the components to have a certain amount of defects in order to
bring down the manufacturing cost while still having the performance require-
ments satisfied. To achieve this higher level of engineering we need certain cap-
abilities to be in place. Referring again to Figure 8.8, the connection between
manufacturing (box at top) and the material state achieved (box at left) requires a
capability to predict the defect structure along with the material composition
attained from the employed manufacturing process. Some attempts in this direc-
tion have been made. As an example, see references [27–31] for prediction of voids
in a liquid compression molding process.
Our efforts are focused on the connection between defect structures and the
mechanical properties, as well as their degradation in service. This type of activity
may be viewed as an extension of damage mechanics, which in its conventional
form deals with initiation and evolution of damage and the consequent changes in
mechanical properties. Thus our starting point in the extended damage mechanics
8.3 Cost-effective manufacturing and defect damage mechanics 291

Strength/dollar

Strength

Manufacturing
cost

Defect density

Figure 8.9. Illustration of the dependence on defect density of strength, manufacturing


cost, and strength per unit cost.

is not a homogenized continuum, but a composite with fibers and matrix as


constituents, and in addition, defects. The defects in our analyses are real-life
defects with their geometry and distribution as given by actual observations.
A broader strategy for durability assessment that includes analysis of defects
was coined as defect damage mechanics [32]. It is discussed next.

8.3.2 Defect damage mechanics


To illustrate the mechanics of damage incorporating defects, we take two
examples in the following.

8.3.2.1 Autoclave processing voids


The first example deals with voids in composite laminates manufactured by
autoclave molding where we describe the observed characteristics of voids and
their incorporation in the modeling (for more details, see [33]).
Figure 8.10 (upper part) shows two cross-sectional views, parallel and normal to
fibers, of a unidirectional carbon/epoxy composite made by the autoclave process.
The voids seen are generally not spherical and are largely trapped between the
prepreg layers. In the lower part of the figure are two cross sections, a short
distance (1.2 mm) apart, showing the voids more closely.
Figure 8.11 summarizes numerous observations and measurements reported in
the literature [35–37] concerning voids in composites made by autoclave molding.
The shape can be described as elongated cylinders of elliptical cross section capped
at the ends. The volume fraction of the voids is found to be less than 3% for a well-
controlled autoclave process, which is safely below the 5% value taken for rejec-
tion of parts in the aerospace industry.
The process by which voids form suggests that the voids must displace the fibers
around them as they settle down in their equilibrium positions. Most continuum
models that homogenize the composite and “embed” voids do not account for this
fact. Such models essentially “replace” fibers, not “displace” them. Reference [33]
accounted for the fiber displacement as schematically illustrated by Figure 8.12.
292 Future directions

1.2 mm

2.5 mm 2.5 mm

Figure 8.10. Observed voids in unidirectional carbon/epoxy composites made by autoclave


process. Upper pictures [34]: cross section parallel to fibers (left) and across fibers (right).
Lower pictures [35]: two cross sections 1.2 mm apart showing voids. Upper pictures
reprinted, with kind permission, from K. J. Bowles and S. Frimpong, J Compos Mater,
Vol. 26, pp. 1487–509, copyright # 1992 by Sage Publications. Lower pictures reprinted,
with kind permission, from Springer Science+Business Media: Review of Progress in
Quantitative Nondestructive Evaluation, A morphological study of porosity defects in
graphite-epoxy composite, Vol. 6B, 1986, pp. 1175–84, D. K. Hsu and K. M. Uhl.

Figure 8.11. Observed characteristics of voids in carbon/epoxy composites made by autoclave


molding.

The predictions of the elastic moduli by the Huang–Talreja procedure [33] are
compared with experimental data in Figure 8.13. Note the large change in the
through-thickness modulus (Ezz) due to the voids.

8.3.2.2 Interlaminar voids


We now consider the effect of the presence of voids in an interlaminar plane
(layer) on the growth of a crack in that plane. As described above, most voids in
8.3 Cost-effective manufacturing and defect damage mechanics 293

Figure 8.12. Modeling of voids accounting for displacement of fibers. Reprinted, with
kind permission, from Compos Sci Technol, Vol. 65, H. Huang and R. Talreja, Effects
of void geometry on elastic properties of unidirectional fiber-reinforced composites,
pp. 1964–81, copyright Elsevier (2005).

Figure 8.13. Elastic moduli predicted by [33] compared with experimental results. Data for
Ex and Ey are from [37] and for Ez from [36]. Reprinted, with kind permission, from
Compos Sci Technol, Vol. 57, C. A. Wood and W. L. Bradley, Determination of the effect
of seawater on the interfacial strength of an interlayer E-glass/graphite/epoxy composite
by in situ observation of transverse cracking in an environmental SEM, pp. 1033–43,
copyright Elsevier (1997).
294 Future directions

layered composites tend to place themselves between layers when manufactured


with a compression molding process. These voids can have different shapes, sizes,
and spacing in the interlaminar plane. This plane is also a plane that is prone to
cracking under service or may have pre-existing flaws due to insufficient adhesion.
A convenient way to make assessment of the effects of voids on interlaminar
fracture is to consider the geometry used for evaluation of interlaminar fracture
toughness. Ricotta et al. [38] conducted a systematic study of the voids’ influence
on crack growth by considering this geometry. In the following some results from
that study are discussed to illustrate the effects.
Figure 8.14 shows a woven fabric composite where voids are found in the resin-
rich regions between the fiber bundles. These regions are likely to develop cracks
under service environment such as fatigue or fail under in-plane compression,
leading to delamination. A representation of the effect of such voids on crack
growth is illustrated in the figure where a double cantilever beam (DCB) specimen
with voids ahead of the crack tip is shown. This geometry for mode-1 crack
growth has been systematically analyzed in [38] considering various parameters
such as void shape (circular and elliptical), void size, and distance of void from of
the crack tip.
The approach described in detail in [38] consists essentially of first validating
an analytical method by finite element analysis and then using the method to
conduct a parametric study of the effects of voids. The method uses a beam-on-
elastic-foundation analysis accounting for shear compliance and material
orthotropic symmetry. The voids are simulated as regions without support
from the elastic foundation. The strain energy release rate with voids present
(GI,v) is calculated for different cases. Figure 8.15 shows the effect of a single
circular void of different radius R placed at different distance D from the crack
tip. The GI,v normalized by the value without void (GI) shows increasing
enhancement as the void approaches the crack tip, and this enhancement
increases with increasing void radius. Figure 8.16 shows a similar effect for
elliptical voids.
The effects of multiple circular voids on the energy release rate are shown in
Figures 8.17 and 8.18. Figure 8.17 shows the effects of two and three voids of fixed
radius and fixed mutual spacing located at different distances from the crack tip.
In Figure 8.18 an interesting effect of void interaction is shown. As seen there, for
multiple circular voids where the nearest void is kept at a fixed distance from the
crack tip, while the mutual void spacing is varied, the energy release rate does not
show a monotonic dependence on the void spacing. Instead, the void interaction
increases the energy release rate up to a certain void spacing, beyond which the
effect of having multiple voids decreases.
Finally, Figure 8.19 illustrates how the energy release rate increases with
crack propagation when a void exists ahead of the crack tip. The increase of the
energy release rate with crack length when no void exists is plotted for refer-
ence. Thus the enhancement of the energy release rate is seen as the crack tip
approaches the void.
8.3 Cost-effective manufacturing and defect damage mechanics 295

Figure 8.14. Voids in resin-rich areas between bundles in a woven fabric laminate and a DCB
specimen representation of crack growth in the presence of the voids.

1.3

1.2
R = 0.2 mm
GI,v /GI

R = 0.1 mm
R = 0.08 mm
R = 0.05 mm
1.1

1
0 5 10 15 20
Distance from crack tip D (mm)

Figure 8.15. Effects of a circular void of radius R and of distance D from crack tip on the
energy release rate.

1.9
b = 0.1 mm
1.8
D
1.7

1.6
a/b = 4
2b
G I,v /GI

1.5 a/b = 2
1.4 a/b = 1.5 2a
a/b = 1
1.3

1.2

1.1

1
0 2 4 6
Distance from crack tip D (mm)

Figure 8.16. Effects of an elliptical void of different aspect ratio a/b, b = 0.1 mm, and of
distance D from crack tip on the energy release rate [38]. Reprinted, with kind permission,
from Compos Sci Technol, Vol. 68, M. Ricotta, M. Quaresimin and R. Talreja,
Mode-I strain energy release rate in composite laminates in the presence of voids,
pp. 2616–23, copyright Elsevier (2008).
296 Future directions

1.3

1.2
G I,v /GI

3 voids R = 0.1 mm
2 voids R = 0.1 mm
1 void R = 0.1 mm
1.1

1
0 1 2 3 4
Distance from crack tip D (mm)

Figure 8.17. Effect on the energy release rate of circular voids of radius R = 0.1 mm and of
2.0 mm mutual spacing placed ahead of the crack tip for varying distance D from the
crack tip.

1.11

1.10
G I,v /GI

1.09

3 voids R = 0.1 mm
2 voids R = 0.1 mm

1.08
0 2 4 6 8 10
Distance from first void c (mm)

Figure 8.18. Effect on the energy release rate of multiple voids of fixed radius R = 0.1 mm of
varying mutual spacing c placed at a fixed distance from the crack tip.

8.4 Final remarks

The analyses and methods presented in this book have been mostly directed at
composite materials having continuous fiber reinforcement in individual layers
that are stacked to form laminates. These material systems with polymers as
matrix materials and with stiff fibers such as carbon have spurred the development
of lightweight structures in the aeronautics industry. Today, new aircraft such as
Boeing 787 and Airbus 380 are products of those developments. It is arguable,
however, how much of the advancement in damage modeling presented in this
book is embedded in the design of these aircraft. While this situation is under-
standable due to the stringent and costly airworthiness certification requirements,
it is hoped that eventually the output of research efforts in damage and failure will
transfer to designing safer and more cost-effective structures.
8.4 Final remarks 297

540

520
D –a
500
GI (J/m2)

480

460
Circular void, R = 0.1 mm
440
Without void

420
0 1 2 3 4
Crack propagation (mm)

Figure 8.19. Increase in the energy release rate as the crack tip approaches the void is
illustrated. The lower curve shows the energy release rate for comparison when no void
exists.

The use of composites has over the years expanded beyond the aerospace
applications to other areas of structures. Carbon fiber composites have experi-
enced an explosive growth in recent decades with an annual growth rate ranging
from 10 to 15%. The emerging applications of composite materials in automotive
and wind energy sectors place different challenges on design of these materials
than what has been the case in the aerospace industry. Although the affordability
of aerospace vehicles, even in the defense industry, has been a consideration, cost-
effectiveness is a prime factor in design of automotive and wind turbine structures.
The defect damage mechanics discussed above is bound to be an integral part of
the design process for these structures in the future. Incorporating this approach
in computational design methodologies will be a crucial next step.
For wind energy applications the key factor is long-term durability, other than
low cost. The design life of these structures is currently at 20 years (earlier it was
30 years!). For fatigue this translates to 10 million load cycles, or more. Most
composites fatigue testing has traditionally been done until 106 cycles, motivated
by metal fatigue where steels typically have a fatigue limit, which is revealed by the
S-N curve flattening out before this number of load cycles. For composite mater-
ials the fatigue limit is not as easily determined. As discussed in Chapter 6,
considerations of damage mechanisms are necessary to deduce this property. This
is a significant challenge for a highly complex composite construction in wind
turbine rotor blades. A much greater challenge is to determine the fatigue life at
a large number of cycles under multiaxial loading conditions typical for these
structures.
The field of multiaxial fatigue in composites must be given the support it
deserves. A large-scale activity that is comprehensive in its approach is needed.
It must include testing and evaluation at scales where damage initiates, to
scales of damage progression (crack multiplication), and failure criteria that
properly represent the mechanisms. The activity so far has been limited in
scope and mostly focused on empirical and semi-empirical approaches. Most
298 Future directions

work continues to emulate metal fatigue despite fundamental differences in the


underlying mechanisms [39].
This book has not specifically dealt with nanoscale reinforcements and compos-
ite systems. Although many advances have recently occurred, it is not yet clear
what advantages this area holds for improving durability at low cost in load-
bearing composite structures. The area of multifunctionality of composites, such
as conducting polymers and their composites, has on the other hand shown clear
promise for applications in structural health monitoring.

References

1. O. O. Ochoa and J. N. Reddy, Finite Element Analysis of Composite Laminates.


(Dordretchet, The Netherlands: Kluwer Academic Publishers, 1992).
2. M. J. Hinton and P. D. Soden, Predicting failure in composite laminates: the back-
ground to the exercise. Compos Sci Technol, 58:7 (1998), 1001–10.
3. M. J. Hinton, A. S. Kaddour, and P. D. Soden, Evaluation of failure prediction in
composite laminates: background to “part B” of the exercise. Compos Sci Technol,
62:12–13 (2002), 1481–8.
4. M. J. Hinton, A. S. Kaddour, and P. D. Soden, Evaluation of failure prediction
in composite laminates: background to “part C” of the exercise. Compos Sci Technol,
64:3–4 (2004), 321–7.
5. Y. W. Kwon, D. H. Allen, and R. Talreja, eds., Multiscale Modeling and Simulation of
Composite Materials and Structures. (New York: Springer, 2008).
6. S. Nemat-Nasser and M. Hori, Micromechanics: Overall Properties of Heterogeneous
Materials, 2nd edn. (Amsterdam: North Holland, 1999).
7. R. Talreja, Multi-scale modeling in damage mechanics of composite materials. J Mater
Sci, 41:20 (2006), 6800–12.
8. R. Talreja, On multiscale approaches to composites and heterogeneous solids with
damage. Philos Mag, 90:31–32 (2010), 4333–48.
9. R. Pyrz, K. Anthony, and Z. Carl, Morphological characterization of microstructures.
In Comprehensive Composite Materials. (Oxford: Pergamon, 2000), pp. 465–78.
10. S. Ghosh, Adaptive concurrent multilevel model for multiscale analysis of composite
materials including damage. In Multiscale Modeling and Simulation of Composite Mater-
ials and Structures, ed. Y. W. Kwon, D. H. Allen, and R. Talreja. (New York: Springer,
2008), pp. 83–163.
11. R. M. Jones, Mechanics of Composite Materials, 2nd edn. (Philadelphia, PA: Taylor &
Francis, 1999).
12. L. E. Asp, L. A. Berglund, and R. Talreja, Effects of fiber and interphase on matrix-
initiated transverse failure in polymer composites. Compos Sci Technol, 56:6 (1996),
657–65.
13. L. E. Asp, L. A. Berglund, and R. Talreja, Prediction of matrix-initiated transverse
failure in polymer composites. Compos Sci Technol, 56:9 (1996), 1089–97.
14. L. E. Asp, L. A. Berglund, and R. Talreja, A criterion for crack initiation in glassy
polymers subjected to a composite-like stress state. Compos Sci Technol, 56:11 (1996),
1291–301.
References 299

15. J. Aveston, G. A. Cooper, and A. Kelly, Single and multiple fracture. In The Properties
of Fibre Composites. (Guildford, Surrey: IPC Science and Technology Press, 1971),
pp. 15–26.
16. R. Talreja, Damage analysis for structural integrity and durability of composite mater-
ials. Fatigue Frac Engng Mater Struc, 29 (2006), 481–506.
17. R. Talreja and C. V. Singh, Multiscale modeling for damage analysis. In Multiscale
Modeling and Simulation of Composite Materials and Structures, eds. Y. W. Kwon,
D. H. Allen, and R. Talreja. (New York: Springer, 2008), pp. 529–78.
18. R. Talreja, A synergistic damage mechanics approach to durability of composite
systems. In Progress in Durability Analysis of Composite Systems, eds. A. M. Cardon
et al. (Rotterdam: A.A. Balkema, 1996), pp. 117–29.
19. J. Varna, R. Joffe, N. V. Akshantala, and R. Talreja, Damage in composite laminates
with off-axis plies. Compos Sci Technol, 59 (1999), 2139–47.
20. J. Varna, R. Joffe, and R. Talreja, A synergistic damage mechanics analysis of trans-
verse cracking in [ y/904]s laminates. Compos Sci Technol, 61 (2001), 657–65.
21. J. Varna, A. Krasnikovs, R. S. Kumar, and R. Talreja, A synergistic damage mechanics
approach to viscoelastic response of cracked cross ply laminates. Int J Damage Mech,
13 (2004), 301–34.
22. C. V. Singh and R. Talreja, Analysis of multiple off-axis cracks in composite laminates.
Int J Solids Struct, 45 (2008), 4574–89.
23. C. V. Singh and R. Talreja, A synergistic damage mechanics approach for composite
laminates with matrix cracks in multiple orientations. Mech Mater, 41 (2009), 954–68.
24. M. Quaresimi and M. Ricotta, Fatigue behaviour of bonded and co-cured joints in
composite materials. In Experimental Techniques and Design in Composite Materials 6,
Extended Abstracts, ed. M. Quaresimin. (Vicenza, Italy: University of Padova, 2003),
pp. 31–2.
25. B. R. Tittmann and R. L. Crane, Ultrasonic inspection of composites. In Comprehen-
sive Composite Materials, Vol. 5, eds. L. Carlsson, R. L. Crane, and K. Uchino; eds.-in-
chief A. Kelly and C. Zweben. (Amsterdam: Elsevier, 2000), pp. 259–320.
26. R. L. Thomas, L. D. Favro, X. Hanand, and Z. Ouyang, Thermal methods used in
composite inspection. In Comprehensive Composite Materials, Vol. 5, eds. L. Carlsson,
R. L. Crane, and K. Uchino; eds.-in-chief A. Kelly and C. Zweben. (Amsterdam:
Elsevier, 2000), pp. 427–46.
27. A. D. Mahale, R. K. Prud’Homme, and L. Rebenfeld, Quantitative measurement of
voids formed during liquid impregnation of nonwoven multifilament glass networks
using an optical visualization technique. Polym Eng Sci, 32 (1992), 319–26.
28. N. Patel and J. L. Lee, Effect of fiber mat architecture on void formation and removal
in liquid composite molding. Polym Composites, 16 (1995), 386–99.
29. N. Patel, V. Rohatgi, and J. L. Lee, Micro scale flow behavior and void formation
mechanism during impregnation through a unidirectional stitched fiberglass mat.
Polym Eng Sci, 35 (1995), 837–51.
30. V. Rohatgi, N. Patel, and J. L. Lee, Experimental investigation of flow induced
microvoids during impregnation of unidirectional stitched fiberglass mat. Polym Com-
posites, 17 (1996), 161–70.
31. S. Roychowdhury, J. W. Gillespie, Jr., and S. G. Advani, Volatile-induced void
formation in amorphous thermoplastic polymeric materials: I. Modeling and paramet-
ric studies. J Compos Mater, 35 (2001), 340–66.
300 Future directions

32. R. Talreja, Defect damage mechanics: broader strategy for performance evaluation of
composites, Plastics Rubber Composites, 38 (2009), 49–54.
33. H. Huang and R. Talreja, Effects of void geometry on elastic properties of unidirec-
tional fiber reinforced composites. Compos Sci Technol, 65 (2005), 1964–81.
34. K. J. Bowles and S. Frimpong, Void effect on the interlaminar shear strength of
unidirectional graphite fiber-reinforced composites. J Compos Mater, 26 (1992),
1487–509.
35. D. K. Hsu and K. M. Uhl, A morphological study of porosity defects in graphite-epoxy
composite. In Review of Progress in Quantitative Nondestructive Evaluation, Vol. 6B, eds.
D. O. Thomson and D. E. Chimenti. (New York: Plenum Press, 1986), pp. 1175–84.
36. Z. Guerdal, A. P. Tamasino, and S. B. Biggers, Effects of processing induced defects on
laminate response: interlaminar tensile strength. SAMPE J, 27 (1991), 3–49.
37. P. Olivier, J. P. Cottu, and B. Ferret, Effects of cure cycle pressure and voids on some
mechanical properties of carbon/epoxy laminates. Composites, 26 (1995), 509–15.
38. M. Ricotta, M. Quaresimin, and R. Talreja, Mode-I strain energy release rate in
composite laminates in the presence of voids, Special issue in honor of R. Talreja’s
60th birthday. Compos Sci Technol, 68:13 (2008), 2616–23.
39. M. Quaresimin, L. Susmel, and R. Talreja, Fatigue behaviour and life assessment of
composite laminates under multiaxial loadings. Int J Fatigue, 32 (2010), 2–16.
Author index

Aboudi 124 Grimes 262


Adkins 149 Gudmundson 111, 113, 114, 117, 155, 208, 209
Adolfsson 192, 208, 209
Ahci 175, 176 Hahn 244, 263
Akshantala 188, 270 Halpin 19
Aoki 70 Hashin 19, 23, 57, 87, 91, 97, 101, 104, 113, 124,
Aveston 3, 4, 49, 58, 66, 74 194, 202, 204, 214, 216, 259
Awerbuch 244 Highsmith 77, 78, 82
Hill 85, 186
Bailey 4, 49, 74, 80, 82, 84, 188, 192, 195, Hoiseth 86, 210
196, 200 Hong 79, 82, 124
Balhi 186 Hori 3, 84, 279
Benthem 115 Hsu 292
Berglund 97 Hu 85, 98, 268
Berthelot 74, 212, 215, 216, 217 Huang 292
Bowles 292
Bradley 40, 293 Ishikawa 70
Broutman 49
Budiansky 44, 85, 249 Jamison 261
Joffe 104, 208, 219
Chaboche 4, 136, 137, 138 Johnson 193
Chang 193 Ju 137
Cooper 49, 58
Cox 74 Kachanov 4, 114, 117, 134, 141
Crocker 186, 193 Katerelos 40, 184
Crossman 206 Kelly 4, 49, 58, 66, 74
Curtis 245 Kim 125, 263
Koiter 115
Daniel 79, 82 Krajcinovic 4, 208
Dharan 239 Kumar 172, 173, 174, 175
Diao 276 Kuriakose 101
Dvorak 76, 82, 85, 200, 201, 202
Lavoie 192
Eshelby 85 Laws 76, 82, 85, 200, 201, 202
Le Corre 212, 215, 216, 217
Fenner 46 Lee 79, 82, 124
Flaggs 80, 82 Lemaitre 4, 136, 138
Frimpong 292 Li 119, 121, 122
Fukunaga 77, 78, 79, 82, 197 Lim 79, 82
Liu 47, 190, 221
Galiotis 40, 46 Lundmark 117, 118, 119, 125, 189, 191
Gamby 124
Gamstedt 38, 247, 248, 249 Mai 276
Garrett 4, 49, 80, 82, 84, 186, 188, 195, 200 Manders 74, 82, 84, 196, 197, 212
302 Author index

McCartney 104, 109, 125, 205 Silberschmidt 212


Mori 19, 44, 281 Silwood 58
Murakami 136, 137 Singh 162, 171, 218, 222, 223, 224
Sirivedin 46
Na 124 Sjogren 38
Nairn 83, 85, 97, 98, 110, 125, 185, 187, 202, 204, Smith 186
205, 221, 268 Sørensen 253, 256
Nath 46 Steif 79, 82, 198
Nemat-Nasser 3, 84, 279 Sturgeon 244, 245
Niu 43
Nuismer 79, 80, 81, 82 Tada 115
Talreja 4, 6, 43, 50, 101, 141, 162, 171, 172, 173,
Ogin 79, 82, 186 174, 175, 188, 218, 222, 223, 224, 238, 239, 247,
Ohno 136, 137 250, 253, 270, 292, 295
Ostlund 113 Tan 79, 80, 81, 82
Tanaka 19, 44, 281
Pagano 124 Timoshenko 43
Parvizi 49, 74, 199 Tracey 45
Parvizi-Majidi 67 Tsai 19, 23, 194

Qu 86, 210 Uhl 292


Queresimin 295
Vakulenko 117, 141
Ravichandran 47, 190 Varna 40, 97, 102, 104, 117, 118, 119, 125, 143,
Rebiere 124 174, 189, 191, 208, 221
Reddy 119, 123 Vinogradov 202, 214, 216
Reifsnider 77, 78, 82, 188
Rice 45 Wang 67, 206, 213
Ricotta 294, 295 Weibull 196, 208, 215
Robotnov 135 Wood 40, 293
Rotem 259 Wu 23, 194

Sahu 49 Yaniv 79
Schapery 172 Ye 276
Shoeppner 124 Yokozeki 70, 193
Shtrikman 19
Sicking 172 Zhang 124
Subject index

ACK theory 4, 58, 199 Damage evolution. See damage progression


Acoustic emission 181, 253 Damage evolution curve. See damage progression
curve
Characteristic damage state (CDS) 47, 48, 188 Damage initiation. See Crack initiation
Classical laminate plate theory (CLPT) 24 Damage mechanics 2, 33, 36, 48, 50, 57, 276, 280,
COD based methods 206 290
Complementary energy 12, 92, 93, 94, 101, 203, Damage mechanisms 36, 37, 142, 143, 181,
227 242, 278
Computational methods 119 Damage modes 45, 46, 142–143, 156, 158, 165,
Computational structural analysis 276 167, 175, 179, 193
Constitutive relations/Constitutive response/ Damage mode tensor 142–143, 158
Constitutive relationships/Constitutive Damage progression 46, 139, 179, 180, 184, 187,
equations 11–15, 20, 26, 172 188, 192, 194, 196, 204, 208, 212, 215, 217,
Constraint factor. See Constraint parameter 221, 222, 229, 253, 266, 268, 270, 273, 276,
Constraint parameter 148, 155, 163, 165, 169, 283, 284, 294
174, 287 Curve 189, 190
Constraint. See ply constraint Energy based approaches 198, 210, 277
Continuum damage mechanics (CDM) 4, 57, 134, Strength based approaches 180, 194, 210
137, 144, 154, 161, 174, 176 Damage state 173
Correspondence Principle 172 Damage tensor 151, 157
Cost-effective manufacturing 288 Damage tolerance 265
Crack density 86, 173, 192, 202, 208, 214, 266, Debonding (interfacial) 37–38, 48, 248, 249, 251,
270 256, 283, 284, 287
Normalized crack density 201 Defect damage mechanics 287, 291, 297
Crack density evolution. See damage progression Delamination 33, 39–41, 48, 183, 187, 193, 216,
Crack initiation 180, 185, 201, 273, 276, 278 225, 260, 271, 273, 287
Crack initiation strain 198, 205, 221 Distortional energy (Tsai-Hill) criterion 23
Crack opening displacement (COD) 2, 3, 32, 102, Durability 1, 37, 265, 279, 287
111, 117, 155, 161, 163, 164, 174, 181, 207, Assessment 277
219
Crack progression. See Damage progression Edge replication 181
Crack sliding displacement (CSD) 3, 143, Energy release rate 31, 203, 204, 208, 294
219 Critical 31, 206, 213, 219
Crack spacing 148, 195
Normalized crack spacing 203 Failure 36
Crack size 148 Failure criteria 21, 180
Critical energy release rate. See energy release Fatigue 223, 265, 276
rate: critical Fatigue-life 246, 271
Cyclic loading. See Fatigue Fatigue-life prediction 265–271
Crossply laminates 266–270, 273
Damage 36 General laminates 273
Damage characterization 139 Fatigue-life diagrams 6–7, 237–265, 273
Damage development. See damage progression Polymer matrix composites (PMC) 242, 250,
Damage entity tensor 141 251, 283
304 Subject index

Fatigue-life diagrams (cont.) Ply constraint 49, 187


Angleply laminates 260, 261 Ply cracking 39, 48, 61, 64, 170, 183, 194, 267, 283
Ceramic matrix composites (CMC) 252 Principle of minimum potential energy 16
Crossply laminates 261–262 Principle of minimum complementary energy
Metal matrix composites (MMC) 250, 251 17, 91
Multidirectional laminates 263 Principle of virtual work 16, 92
Quasi-isotropic laminates 263
Unidirectional composite 257 Raman spectroscopy 69, 184
Fatigue limit 237 Randomness in ply cracking 70, 210
Fiber breakage 42, 48, 183, 239, 249, 255 Randomness. See randomness in ply cracking
Fiber failure. See fiber breakage Reference Laminate 220, 221
Fiber microbuckling 42, 43 Representative volume element RVE 2, 5, 69,
Finite element method (FEM) 117, 119, 209, 276 110, 121, 134, 139–148, 161, 278, 280, 281,
Finite strip method 121 282, 287
Fracture 36, 48 Residual stresses 96, 192, 201, 208
Frictional sliding 256, 268 See also Interfacial RVE. See representative volume element
sliding
Self-consistent method. See Self-consistent scheme
Generalized Hooke’s law 12 Self-consistent scheme 84–86
Generalized plain strain analysis 104, 205 Shear lag analysis 198
Shear lag methods. See Shear lag models
Hashin criterion 23 Shear lag models 65, 73–84, 101–104
Shear lag parameter 76–82, 103, 195, 197, 201
Interfacial sliding 39, 41 Shear lag. See Shear lag models
Interlaminar cracking. See delamination Shear lag theory. See Shear lag models
Intralaminar cracking. See ply cracking Stiffness changes. See Stiffness degradation,
Stiffness – damage relationships
Laminate 18 Stiffness degradation 83, 85, 102, 154, 170, 193
Balanced 29 Stiffness-damage relationships 148, 152, 157, 169,
Crossply 29, 285 179, 184
Multidirectional 193, 217 Stiffness reduction 192
Quasi-isotropic 29, 193 Strain energy density 12
Symmetric 29 Strength criteria. See Failure criteria and Damage
Layerwise theory 123 progression: strength based approaches
Length scales of damage 280 Structural integrity 1, 37
Linear elastic fracture mechanics (LEFM) 29, 184 Synergistic damage mechanics (SDM) 5, 155, 156,
164, 170, 171, 174, 176, 184, 286
Macro damage mechanics (MADM) 5, 51, 57,
126, 134, 276 Thermal expansion coefficient 125
Matrix cracking. See ply cracking Thermal residual stress 202, 205
Maximum stress theory 22 Thermal stress. See Thermal residual stress
Maximum strain theory 22 Tsai-Wu criterion 23
Mechanisms of damage. See damage mechanisms
Microbuckling. See Fiber microbuckling Ultrasonic C-scan 183
Microcracking. See ply cracking Unidirectional lamina (UDL) 18
Micro damage mechanics (MIDM) 4, 51, 57, 126,
134, 276 Variational analysis. See variational methods
Micromechanics 3, 17, 44, 57, 144, 161, 176 Variational methods 87, 97, 202, 226
Multiple cracking 63 Virtual work. See Principle of virtual work
Multiple matrix cracking. See ply cracking Viscoelastic composites 170
Multiscale modeling 278 Viscoelastic response 173, 174
Hierarchical 279, 282, 286 Viscoelasticity 170
Void growth 44
Particle cleavage 44
Periodic boundary conditions 120 X-ray radiography 181, 261

You might also like