Sie sind auf Seite 1von 157

PROGRESSIVE D A M A G E M O D E L I N G OF COMPOSITE MATERIALS

UNDER COMPRESSIVE LOADS

by

Navid Zobeiry

B.Sc. (Civil Engineering), University of Tehran, 2002

A THESIS SUBMITTED IN PARTIAL FULFILMENT OF


THE REQUIREMENTS FOR THE DEGREE OF

Master of Applied Science

in

THE FACULTY OF GRADUATE STUDIES

Department of Civil Engineering

THE UNIVERSITY OF BRITISH COLUMBIA

December 2004

Navid Zobeiry, 2004


ABSTRACT

The in-plane compressive strength of fibre reinforced composite materials is known to be


less than their corresponding tensile strength. There are a multitude of compression
damage mechanisms that occur in composites, the form of which depends on the
properties of the constituents and fibre lay-up. These mechanisms primarily consist of a
combination of matrix cracking (yielding), localized buckling of fibres or kinking, and
delamination. Whether the interest is to assess the structural integrity of composite
materials in-service or to quantify their energy absorption capability under axial crushing,
it is crucial to have predictive analysis tools that capture the physics of the damage
mechanisms and their propagation under compressive loads.

In this study, a constitutive model is formulated for the complete in-plane response of
composite materials within the framework of a previously developed continuum damage
mechanics model CODAM (Williams, 1998; Williams et al., 2003; Floyd, 2004). While
the previous CODAM formulation was limited to simulating the progression of damage
under tensile loading, the current formulation accounts for the initiation and propagation
of damage under compression, tension and load reversals in each mode of loading. The
model is implemented in the commercial finite element code, LS-DYNA, and combined
with a modified crack band model originally developed by Bazant (Bazant and Planas,
1998) to overcome the mesh sensitivity problems that plague all strain-softening type
constitutive models.

The new model is validated against two sets of experimental data available in the
literature, namely, eccentric compression loading of notched sandwich panels of various
sizes (Bayldon, 2003a, 2003b), and axial compression of composite panels with central
open holes of various panel and hole sizes (Soutis et al., 1993, 2002). It is shown that for
these loading applications the predictions of the compressive strengths and the degree of
size effect are in good agreement with the measured experimental results. Since the

ii
Abstract

formulation of the model and its calibration are based entirely on the fundamental physics
of the damage mechanisms, these successful validations instil confidence in exercising
the model for predicting the response of composite structures of various sizes under a
variety of in-plane loading applications involving compression, tension and load
reversals.

iii
TABLE OF CONTENTS

ABSTRACT II

TABLE OF CONTENTS IV

LIST OF TABLES VIII

LIST OF FIGURES IX

NOMENCLATURE XV

ACKNOWLEDGEMENTS XVIII

CHAPTER 1 : INTRODUCTION , 1

1.1. Background 1

1.2. Motivation 4

1.3. An Outline of This Study 6

CHAPTER 2 : COMPRESSIVE FAILURE OF FIBRE COMPOSITES 7

2.1. Introduction 7

2.2. Failure Mechanisms in Unidirectional Composites 8


2.2.1. Elastic Microbuckling 10

2.2.2. Fibre Crushing 10

2.2.3. Splitting 10

2.2.4. Buckle Delamination 11

2.2.5. Shear Band Formation 11

2.2.6. Plastic Microbuckling or Kinking 12


Table of Contents

2.2.6.1. H i s t o r y o f K i n k B a n d A n a l y s i s 12

2.2.6.2. K i n k i n g F o r m a t i o n 16

2.3. Failure Mechanisms in Angle-ply Composites 19

2.4. Summary 21

CHAPTER 3 : ANALOG MODEL 22

3.1. Introduction 22

3.2. Elements : 25
3.2.1. R i g i d L i n k 25

3.2.2. S p r i n g and T / C S p r i n g 25

3.2.3. S l i d e r 27

3.2.4. Fuse and T / C Fuse 27

3.2.5. G a p 29

3.2.6. L o c k 29

3.3. Collection of Elements or Boxes 31


3.3.1. L a m i n a t e B o x 31

3.3.1.1. L a m i n a t e B o x E q u a t i o n s 33

3.3.2. R u b b l e B o x 42

3.4. Analog Model 46


3.4.1. M o d e l C o n s t r u c t i o n 46

3.4.2. M o d e l F o r m u l a t i o n 48

3.4.2.1. L o a d R e v e r s a l : C o m p r e s s i o n - U n l o a d i n g B e f o r e B a n d B r o a d e n i n g - T e n s i o n 49

3.4.2.2. L o a d R e v e r s a l : C o m p r e s s i o n - U n l o a d i n g A f t e r B a n d B r o a d e n i n g - T e n s i o n 51

3.4.2.3. L o a d R e v e r s a l f o r m T e n s i o n to C o m p r e s s i o n 53

3.5. Further Remarks on the Analog Model 55

CHAPTER 4 : DAMAGE GROWTH, ENERGY CONCEPT AND SCALING LAW


56

4.1. Damage Growth 57

4.2. Compressive Fracture Energy 61


Table of Contents

4.2.1. Main Fracture Energy 62

4.2.2. Band Broadening Fracture Energy 63

4.3. Compressive Fracture Energy in the Damage Propagation Process 65


4.3.1. Compressive Damage Propagation 65

4.3.2. Compressive Fracture Energy 67

4.3.2.1. Linear Damage Height Function 68

4.3.2.2. Generalized Damage Height Function 69

4.3.3. Summary of Fracture Energy Functions 71

4.3.4. Comparison with Experimental Results 71

4.4. Scaling Law 73


4.4.1. Crack Band Method 73

4.4.2. Modified Crack Band Method 74

4.4.3. Scaling Factor in Compression 74

4.4.3.1. Scaling Factor for an Element with Peak Stress at the Damage Initiation Strain 75

4.4.3.2. Scaling Factor for an Element with Peak Stress not at the Damage Initiation Strain 76

4.4.4. Summary of Scaling Factor Equations 77

CHAPTER 5 : MODEL VERIFICATION AND VALIDATION 78

5.1. Model Verification 79

5.2. Simulations of Notched Sandwich Panels under Eccentric Compression 82


5.2.1. Sandwich Panel Failure 83

5.2.2. Test Geometry 85


5.2.3. Failure Mechanisms of Woven Glass Laminates 88

5.2.4. Material ' 91

5.2.5. Finite Element Model 96

5.2.5.1. Experimental Results .' 96

5.2.5.2. Damage Initiation Strain 97

5.2.5.3. Scaling Factor 101

5.2.6. Simulations 103

5.2.7. Summary 107

5.3. Simulation of Open Hole Plates under Compression 108


5.3.1. Introduction 108

5.3.2. Test Geometry 110

vi
Table of Contents

5.3.3. Material 112

5.3.4. Stress-Strain Curve Scaling 116

5.3.4.1. Analytical Solution . 116

5.3.4.2. Stress Contour 117

5.3.4.3. Scaling 118

5.3.5. Simulation Results 121

5.3.5.1. Notch Size Effect 122

5.3.5.2. Specimen Sire Effect 124

5.3.6. Summary of Observations 126

CHAPTER 6 : CONCLUSIONS AND FUTURE WORK . 127

6.1. Conclusions 127

6.2. Future Work 130

REFERENCES 131

APPENDIX A: TAYLOR SERIES FOR DAMAGE HEIGHT FUNCTION 139


LIST OF TABLES

Table 4-1 Damage height and fracture energy functions for different values of n in
Equation 4-26 ; 71

Table 4-2 Material parameters used to verify fracture energy functions 71


Table 5-1 Compressive damage parameters for AS4/PEEK unidirectional panels.. 79
Table 5-2 Tensile damage parameters for AS4/PEEK unidirectional panels 80
Table 5-3 Material parameters for AS4/PEEK unidirectional panels 80
Table 5-4 Test geometries of notch size effect experiments (ECC). All numbers are
in mm 86
Table 5-5 Test geometries of specimen size effect experiments (CDC). All numbers
are in mm 86
Table 5-6 Properties of the foam core. All properties are in the loading direction... 91
Table 5-7 Properties of the face sheets. Properties in both directions are the same.. 91
Table 5-8 Damage parameters for woven glass-epoxy laminate 95
Table 5-9 Damage initiation strains for different CDC experiments based on their
elastic analysis 101
Table 5-10 Scaling factors for notched sandwich panel under eccentric compression...
102
Table 5-11 Damage parameters for woven glass-epoxy laminate which give the best
result for CDC simulations 106
Table 5-12 Specimen sizes with different notch sizes for Soutis et al. experiment
(2002) 111
Table 5-13 Material parameters for the carbon-epoxy laminate in the compressive
load direction 112

Table 5-14 Damage parameters for [(45/0/90) 3] carbon-epoxy laminate


s 115

Table 5-15 Modified parameters for the simulation of 50*50 mm specimens 120

viii
LIST OF FIGURES

Figure 2-1 The main compressive failure mechanisms of unidirectional composites.


(a) Elastic microbuckling, (b) Kinking, (c) Fibre crushing, (d) Splitting, (e)
Buckle delamination, (f) Shear band formation (Fleck, 1997) 9
Figure 2-2 Kink band formation under compressive load for a unidirectional laminate.
(a) Before failure, (b) After failure 13
Figure 2-3 Kink band Formation in a unidirectional laminate under compressive load.
(a) Out-of-plane kink band, (b) In-plane kink band 15
Figure 2-4 Overall behaviour of unidirectional fibre composites under compressive
loads in the fibre direction when kinking occurs (Moran et al., 1995) 16
Figure 2-5 Initial damage height in kinking process (Fleck et al., 1995). Fibres rotate
under compressive load 17
Figure 2-6 Schematic showing the behaviour of a single fibre during the early stage
of kink band formation (Moran et al., 1995) 18
Figure 2-7 (a) Overview of arrested damage zone, (b) Band broadening in progress.
(c) A close-up view of the microbuckle band, (d) 45 degree off-axis
microbuckle band linking the 0/45 degree delaminated interfaces, (e) Out-
of-plane 0 degree microbuckle band extending into the 45 degree off axis
layer. (Sivashanker, 2001) 20
Figure 3-1 The behaviour of a rigid link element under compressive and tensile
loads 25
Figure 3-2 The symbol and behaviour of a spring element under compressive and
tensile loads 26
Figure 3-3 The symbol and behaviour of a T/C spring element under compressive and
tensile loads 26
Figure 3-4 The symbol and behaviour of a slider element under compressive and
tensile loads '. 27

IX
List of Figures

Figure 3-5 The symbol and behaviour of a fuse element under compressive and
tensile loads 28
Figure 3-6 The symbol and behaviour of a T/C fuse element under compressive and
tensile loads 28
Figure 3-7 The symbol and behaviour of a gap element under compressive and tensile
loads 29
Figure 3-8 The symbol and behaviour of a lock element under compressive and
tensile loads 30
Figure 3-9 Laminate box and the arrangement of basic units. All the elements are T/C
springs and T/C fuses 32
Figure 3-10 Basic unit as used in the laminate box. "m" represents the response of
matrix and 'J/" represents the response of fibre in a representative volume
element 32
Figure 3-11 Force-displacement response of a basic element in the laminate box 33
Figure 3-12 Grouping of matrix and fibre elements in the laminate box under tensile
loads 34
Figure 3-13 The response of group "w" under tensile loads in which the group consists
of four elements 35
Figure 3-14 The response of group "tn" under tensile loads in which group consists of
an infinite number of elements 36
Figure 3-15 Force-displacement response of the laminate box under tensile loads, m
and/represent the response of matrix and fibre respectively 37
Figure 3-16 Grouping of elements in the laminate box under tensile loads..: : 38
Figure 3-17 Individual responses of Group m and Group / of elements in the laminate
box under compressive loads. Group m represents the response of matrix
while Group/represents the response of fibres. 39
Figure 3-18 Force-Displacement response of Group m and Group / while in parallel
under compressive loads 40
Figure 3-19 A schematic of the Rubble Box and the basic elements inside it 42
Figure 3-20 Force-displacement and stiffness-displacement responses for an infinite
number of units in the rubble box 44
List of Figures

Figure 3-21 A schematic of the analog model which shows arrangement of elements
and boxes. 46
Figure 3-22 Force-displacement response of Spring/Lock elements in series 48
Figure 3-23 Force-displacement response of the analog model. First compressed, then
unloaded and finally loaded in tension to complete failure 51
Figure 3-24 Force-displacement response of the analog model (Unloading after the
yielding) 52
Figure 3-25 Force-displacement response of the analog model. First pulled in tension,
then unloaded and finally loaded in compression to complete failure 54
Figure 4-1 Damage growth-potential function) diagrams for tensile and compressive
loading 59
Figure 4-2 (Modulus reduction-damage growth) diagrams for compressive and tensile
loading. Modulus reduction due to matrix saturation is more pronounced
in compression comparing to the reduction in tension 60
Figure 4-3 A schematic of stress-strain response of composites under compressive
loads which showing areas associated with formation of kink bands and
damage band broadening 61
Figure 4-4 Calculation of the main fracture energy from the area under the stress-
strain response of the laminate box 62
Figure 4-5 The unloading path after the saturation of each kink band. Only the stress-
strain response of the rubble box is shown here 63
Figure 4-6 Stress contour in front of the notch and Stress-Strain response of one
element 65
Figure 4-7 Stress contour in front of the notch while damage is growing. Damage
height grows from h to hi and finally to hui- The maximum damage
c

height, JJUL, is the damage height at the ultimate strain 66


Figure 4-8 Damage height increment due to the damage length increment 67
Figure 4-9 Calculation of the main fracture energy for one element from the stress-
strain response of that element 68
Figure 4-10 Comparison of experimental results (Jackson and Ratcliffe, 2004) and
analytical solutions for critical stress intensity factor functions 72
List of Figures

Figure 4-11 Crack band method. Both the master and the scaled cure are shown here...
73
Figure 4-12 Modified crack band method. Both the master and the scaled cure are
shown here 74
Figure 4-13 Different unloading paths for compressive failure. The left diagram shows
unloading from the peak stress point while the peak stress and damage
initiation coincide. The right diagram shows the unloading path for an
element in which the peak stress and damage initiation do not coincide. 75
Figure 5-1 . Stress-Strain response of one element obtained from the implemented
constitutive model in LS-DYNA. Some of the input parameters are
obtained from Vogler and Kyriakides (1999) while the damage parameters
are estimated 81

Figure 5-2 Failure modes in sandwich columns (Fleck and Sridhar, 2002). From left
to right, Euler macrobuckling, core shear macrobuckling, face sheet
microbuckling and face sheet wrinkling 84
Figure 5-3 Notched sandwich panel under double eccentric compression (Bayldon,
2003a, 2003b) 85
Figure 5-4 Test geometry for notched sandwich panels under double eccentric
compression 87

Figure 5-5 Initiation of kinking with matrix cracking. Photoelastic image for the
notched sandwich panel under eccentric compression (Bayldon, 2003a). 89
Figure 5-6 Initial damage height is equal to 3.5 mm for kinking and delamination.
Photoelastic image for the notched sandwich panel under eccentric
compression (Bayldon, 2003a) 90
Figure 5-7 Propagation of the kink band and delamination in horizontal and vertical
directions. Height of the damaged zone is almost equal to 6 mm.
Photoelastic image for the notched sandwich panel under eccentric
compression (Bayldon, 2003a) 90

Figure 5-8 Experimental size effect result for CDC sandwich panels under eccentric
loading 97
List of Figures

Figure 5-9 Gauss integration points for CDC8. Only half of the specimen is shown
here 98
Figure 5-10 Stress contour at the notch tip for CDC2. Triangles are the Gauss
integration points of the finite element model. The zero distance point is
extrapolated based on the stress at other points 99
Figure 5-11 Stress contour at the notch tip for CDC4. Triangles are the Gauss
integration points of the finite element model. The zero distance point is
extrapolated based on the stress at other points. 99
Figure 5-12 Stress contour at the notch tip for CDC8. Triangles are the Gauss
integration points of the finite element model. The zero distance point is
extrapolated based on the stress at other points 100
Figure 5-13 Stress contour at the notch tip for CDC16. Triangles are the Gauss
integration points of the finite element model. The zero distance point is
extrapolated based on the stress at other points 100
Figure 5-14 Finite element model for simulating the CDC8 experiment. Only half of
the specimen is modeled 103
Figure 5-15 Simulation of CDC experiments with different fracture energies or with
different saturation strains 104
Figure 5-16 Simulation of CDC experiments with different damage initiation strains
while fracture energy is equal to 20.9 mJ/mm 2
105
Figure 5-17 Test geometry for Soutis et al. experiments (2002). "b" is the length of the
tab section 110
Figure 5-18 Modified ICSTM compression test fixture 111
Figure 5-19 Open hole panel under tensile loads 116
Figure 5-20 Finite element model for the half of the 50x50 mm specimen with the hole
diameter equal to the 10 mm 117
Figure 5-21 Stress contour at different load stages for the open hole specimen in Figure
5-20. As damage progresses, the notch tip stress drops after the remote
stress reaches 236 MPa 118

Figure 5-zZ Modifying the Stress-Strain curve based on the element width 119

Xlll
List of Figures

Figure 5-23: Modifying the Stress-Strain curve based on the element height. The
< scaling is shown for an element height less than h c 119
Figure 5-2f Analytical and numerical strength predictions of 50x50 mm open hole
panels under uniaxial compressive loads 121
Figure 5-25 Notch size effect results for different fracture energy values 122
Figure 5-26 LS-DYNA strength predictions for different saturation strains in 50x50
mm panels. Fracture energy is equal to 25.4 mJ/mm 2
123
Figure 5-2f. Specimen size effect for different fracture energy values. (Hole diameter/
Specimen width = 0.2). 124
Figure 5-2$ Specimen size effect for different damage saturation strains. (Hole
diameter/ Specimen width = 0.2) 125

xiv
NOMENCLATURE

Superscript

C Compression
T Tension

Subscript

00 Relating to the remote stress


1,2,..,/,.., Relating to the 1 , 2 , ith, nth, etc. element
ST nd

app Relating to the applied force

c Relating to the characteristic material parameter


core Relating to the sandwich panel core
e Element
F Final

f In C O D A M , fibre

f Relating to the fracture energy

face Relating to the sandwich panel face

fb Relating to the broadening fracture energy

fm Relating to the main fracture energy

g In C O D A M , gap

8 Relating to the remote stress

/ Initiation
inc Relating to the interval strains in kink band formation
j Relating to the force of j active elements
Nomenclature

L Laminate box
m In CODAM, matrix
max Relating to the maximum stress
M Relating to the modified parameter
new Related to the modified Young's modulus
P Plastic
peak Relating to the peak stress
plateau Relating to the plateau stress
R Rubble box
S Saturation
tot Relating to the total stiffness
UL Ultimate
UN Unloading
V Relating to the lock element
Relating to the lock element gap
y Relating to the slider yielding force

Latin Symbols

a Modification factor for Young's modulus in element width effect


b Width of the laminate
Cfr Length of the failure process zone
d Deformation
E Young's modulus
F Force
F(E) Damage potential function (effective strain)
Fj., F M Strength of fuse element
q Number of failed elements
G Fracture energy
Nomenclature

h Damage height
j Number of active elements
K,L,S,T,U In CODAM, strain interaction constants
k While used without index, Scaling factor
k While used with index, Stiffness
k 0
Inter-laminar shear strength
L Length of the damage band
n In CODAM, the ration of characteristic height to the element height
n In compressive fracture energy, the power of the damage height function
P Applied force
RE In CODAM, Ratio of remaining modulus to the initial modulus
T Time
U Strain energy

Greek Symbols

a Fibre rotation during the kink band formation


P Kink band inclination
7 Fracture energy density
To Composite yield strain
A Increment
5 Displacement
s Strain
a Stress
Initial misalignment of the fibres
CO Current damage parameter

xvii
ACKNOWLEDGEMENTS

First, I would like to thank my family members for their unconditional love,
encouragement and support. I would like to thank my dear parents, Til la and Mohammad
for providing me a life full of joy and inspiration. I would like to thank my sister, Nazli,
for loving me so much. I would like to express my gratitude to my compassionate
brother, Nima, for his important suggestions, remarks and countless patient hours spent in
reviewing this thesis which aided me to improve it significantly. I would also like to
thank my Yalda, who supported me in all means during the last two years.

I would like to sincerely and especially thank my supervisors, Dr. Reza Vaziri and Dr.
Anoush Poursartip, for their guidance, encouragement and constructive criticisms. Their
valuable suggestions and constant support have made this thesis a better product.

I would like to express my deep appreciation to all my friends in Composite Group


specially Carla McGregor who assist me to accomplish such a wonderful joint study. I
would not have made it without your support.

My biggest personal thanks goes to my friends in Vancouver as they have provided me


such a wonderful surrounding to concentrate on my studies. I appreciate you all for
listening to my constant thesis stress and strain. Let God bless you all.

Finally, my best thanks of all goes to God, who has always been there for me. I thank you
for all of your inspirations that guided me to learn a tiny part of the science.

xviii
Chapter 1: Introduction

1.1. Background

Composite materials are widely being used in industrial applications, e.g. aerospace and
automotive industries. The growing popularity of composite materials is mainly due to
their cost-effective manufacturing of large scale structures, light weight, high specific
stiffness and energy absorbing capabilities under high intensity impact loading. They
have also been identified as a suitable class of materials for use in naval and offshore
structures where they can replace more conventional metallic materials. Their use
eliminates the maintenance costs associated with corrosion protection required for metal
structures in deep-water applications. Owing to their relatively small compressive
strength compared to their tensile strength, designers have been understandably cautious
about using these materials in such applications where compression is the dominant mode
of loading.

Often used as face sheets in sandwich components, composite materials are steadily
gaining popularity as spacecraft and marine construction materials. These composite
components are used in aircraft primary and secondary structures, such as primary wing
and fuselage structures of Boeing aircraft, floor panels for military aircraft, structure of
racing yachts, superstructures for fast passenger ferries, control surfaces of submarines
and internal ship hull stiffeners where they undergo axial compression as the main
loading condition.

Because of the multitude of damage mechanisms involved, composite materials have


superior specific energy absorption properties when compared with common engineering
metallic materials. Composites are therefore the preferred materials to be used in impact-
related applications in which high specific energy absorption is a critical design criterion.

1
Chapter 1: Introduction

There are several issues related to the behaviour of composite materials that need to be
addressed. One of the important issues is the prediction of failure and damage
propagation from an imperfection under compression. Composite components are likely
to contain many voids (stress raiser sites) introduced either intentionally, e.g. cut-outs and
fastener holes, or unintentionally, e.g. fabrication defects and imperfections, mismatch
between different fibre orientations, misalignment of fibres and accidental damage. These
defects can act as potential sites for development and propagation of cracks under
compressive loading. Despite these problems, the reduction of the mechanical properties
of composite components caused by defects and the subsequent reduction in their
compressive load carrying capacity have received little attention in the literature.

It is well known that the tensile strength of composite materials stem from the high
tensile strength of the fibres. In compression, however, the matrix and fibre-matrix
interface play important roles because they must provide lateral support for the fibres and
prevent them from undergoing microbuckling. Furthermore, under compressive loads or
impact loads, composite materials can deform and fail in combinations of several
interacting failure modes such as delamination and fibre microbuckling that emanate
from voids. While much research work has been conducted on the tensile failure and
damage behaviour of composite panels containing holes, relatively little work has been
done on the effects of holes and voids on the compressive behaviour of composite
components. Therefore there is a need for a more in-depth study of the nonlinear
behaviour of composite materials under compressive loading in order to gain a better
understanding of their load carrying capability in such loading scenarios. This
understanding can then guide the formulation of physically realistic constitutive models
that when implemented in finite element codes can be used to simulate the response of
composite structures under compressive loads.

As stated before, one of the main advantages of using composite materials is their high
specific energy absorption capability. For example, tubular structures made of composite
materials are being considered as frame structures in automotive applications. A key
requirement in such cases is to absorb the impact energy of accidental collisions through
axial crushing of the tubes. Under such loading scenarios, a significant volume of the

2
Chapter I: Introduction

composite component undergoes a full range of compressive stress-strain response and


fails completely. Failure associated with impact usually appears in the form of one or
several combined failure mechanisms such as fibre fracture, matrix cracking, debonding
and delamination. No matter which mechanism prevails depending on the specific
application, damage always results in a reduction in strength and stiffness of the
structure. To assess impact energy absorption capability or damage tolerance of a
structural component, the micromechanical behaviour of failure process under
compression needs to be well understood and models need to be developed that capture
this behaviour in numerical analyses of such structures.

3
Chapter 1: Introduction

1.2. Motivation

During the past forty years, many studies have been focused to improve our
understanding of compressive failure in composite materials. It is well recognized that
the compressive failure of these materials is mainly caused by a combination of localized
buckling of fibres, kinking, and delamination (Sivashanker, 1998).

Recently, various micromechanical aspects of the kinking failure as the main


compressive failure mechanism have appeared in the literature. Based on the studies of
Sutcliffe and Fleck (1994) and Fleck et al (1997), a good understanding of the fracture
mechanics aspects of kinking has been gained. These studies revealed the crack- like
behaviour of the kink band which was confirmed by Moran et al (1995) who discovered
the phenomenon of band broadening.

A brief review of literature, reveals the lack of a consistent description for composite
materials failure under compression and tension. Although the micromechanical
behaviour of composite materials under tensile and compressive loading induced damage
have been studied thoroughly, little has been done to relate the damage propagation
behaviour of these materials in tension to damage propagation in compression. Also little
work has been reported on models that take into account the micromechanical behaviour
of composites under load reversals from tension to compression and vice versa.

Recently, a plane-stress continuum damage mechanics based model for composite


materials (CODAM) was developed at UBC ( Williams, 1998; Williams et al., 2003;
Floyd, 2004). This model was implemented in the non-linear finite element code LS-
DYNA and its predictive capability for size effect was evaluated for tensile loading
applications (Floyd, 2004). It has been shown that CODAM predicts a size effect for the
tension specimen geometries.that is in good agreement with the experimentally observed
size effect.

The goal of this thesis is to construct a physically meaningful model within the
framework of CODAM, capable of describing the micromechanical damage propagation
behaviour of composites under different loading conditions from initiation of damage all
4
Chapter I: Introduction

the way to its fully saturated damage state. It is also desired to relate the damage effect in
one loading condition, such as axial compression, to the response of material in other
loading conditions. This research is mainly focused on studying the response of
composites under tension, compression and reversal loadings.

The practical importance of the model is that it will enable the designers to predict the
performance of large structures by doing experiments on small composite parts rather
than performing the costly full-scale experiments. Besides the cost effective method of
predicting the performance, we can obtain the desired safety level on large structures by
changing and improving the parameters that matter most, rather than improving
everything, which could be a costly proposition.

5
Chapter 1: Introduction

1.3. An Outline of This Study

An outline of this thesis is briefly presented here. In Chapter 2, a literature review of


composite materials related to the compressive failure is presented. Both unidirectional
and multidirectional composite failures are considered. This chapter is concluded with a
detailed micromechanical description of the behaviour of composite materials during the
compressive failure.

In Chapter 3, a physically meaningful damage mechanics model, namely an analog


model, to simulate the response of composite materials under compression, tension, and
cyclic loads (load reversals) is constructed. Based on the response of various elements
used to construct the model appropriate formulations are derived to relate the
characteristic parameters of material to the response of composites.

In Chapter 4, some issues, which need to be considered before implementing the analog
model in LS-DYNA, are discussed. These issues are focused on damage growth, energy
concept, and scaling law for the compressive response of composites.

Chapter 5 is dedicated to verification and validation of the analog model. Verification


consists of simulating the response of a laminate to check the model implemented in LS-
DYNA. The successful verification is obtained by studying the stress-strain response of
one element.

Validation also consists of comparing the predictions the experimental results obtained
from the literature consisting of sandwich panels under eccentric compression (Bayldon,
2003a, 2003b) and open hole panels under compression (Soutis et al., 2002).

Finally, Chapter 6 presents conclusions and recommendations for future work.

6
Chapter 2: Compressive Failure of Fibre Composites

2.1. Introduction

Compressive failure of composite materials has been an area of considerable research


over the past forty years. This is mainly due to the fact that compressive failure is a
design-limiting feature of fibre composite materials. For instance, the compressive
strengths of unidirectional carbon-epoxy laminates are often less than 60% of their tensile
strengths as a result of compressive failure mechanisms that are rather sensitive to fibre
misalignment.

It is acknowledged that the compressive failure of composites is usually caused by


microbuckling of fibres, also known as kinking. This failure mode is particularly
sensitive to the shear properties of the matrix and also the degree of misalignment of the
fibres (Argon, 1972; Budiansky, 1983; Fleck, 1997). Fibre misalignment usually causes
the kinking to localize in areas of imperfection and therefore reduces the compressive
strength of composite materials compared to their tensile strength.

In this chapter, the main compressive failure mechanisms of composites for both
unidirectional and multidirectional laminates are summarized. Emphasis is placed on
studying the kinking failure as the main failure mechanism of unidirectional fibre
composites. Micromechanics of kinking propagation are also described in the following
section. For multidirectional composites the main failure mechanism is usually a
combination of kinking, delamination and off-axis matrix cracking.

7
Chapter 2: Compressive Failure of Fibre Composites

2.2. Failure Mechanisms in Unidirectional Composites

Long fibre composites are usually designed taking advantage of their high axial stiffness
and strength. For this reason, the fibres are usually made from strong and stiff materials,
such as graphite or silica glass. At the same time, to provide the composite with adequate
in-plane strength and ductility, the matrix needs to possess high toughness properties.
Also as described before, the axial compressive strength of composites is relatively low
due to weak matrix properties and misalignment of fibres. These two sources are
responsible for most of the individual mechanisms of compressive failure.

In 1993, Budiansky and Fleck presented a study on compressive failure of fibre


composites. They focused on studying kinking as the main failure mechanism in
unidirectional laminates. Subsequently, Fleck (1997) presented a comprehensive study
on compressive failure mechanisms of unidirectional and notched multidirectional
composite laminates. In his study, he categorized the main mechanisms of compressive
failure for unidirectional composites as follows (see Figure 2-1).

1- Elastic microbuckling. This is a shear buckling instability, in which the matrix


deforms in simple shear.

2- Fibre crushing. This failure is due to the waviness of the fibres embedded in a soft
matrix and occurs at the fibre level. The compliant matrix provides inefficient
lateral support for the fibre, which leads to fibre buckling.

3- Splitting. In this mechanism, for matrix materials with low toughness, matrix
cracks form parallel to the main fibre direction.

4- Buckle delamination. This is actually debonding and buckling of a surface layer


from the sub-surface. This failure mechanism often occurs due to low matrix
toughness and a large surface flaw. Also, post-impact compressive loading can
induce buckle-delamination growth due to large debonding after the impact.

8
Chapter 2: Compressive Failure of Fibre Composites

5- Shear band formation. In this failure, matrix cracking or yielding occurs in a band
oriented at about 45 to the loading axis.

6- Plastic microbuckling or kinking. This mechanism is a shear instability, which


occurs when matrix deforms nbnlinearly. After cracking or yielding of the matrix,
instability of fibres leads to the fibre rotation and kink band formation.

Each of these mechanisms is reviewed in the following sections.

single
fiber

(a)

0
fiber -

(d) ()

shear
planes

Figure 2-1 T h e main compressive failure mechanisms of unidirectional composites, (a) Elastic
microbuckling, (b) Kinking, (c) Fibre crushing, (d) Splitting, (e) Buckle
delamination, (f) Shear band formation (Fleck, 1997).

9
Chapter 2: Compressive Failure of Fibre Composites

2.2.1. Elastic M i c r o b u c k l i n g

Rosen (1965) pioneered an analysis to model elastic microbuckling by assuming elastic


bending of the fibres and elastic shearing of the matrix. Rosen assumed that elastic
microbuckling occurs in two possible modes: Either a transverse buckling mode, where
matrix undergoes strains transverse to the axial direction or a shear buckling mode, where
the matrix shears in the axial direction (see Figure 2-1).

In practice, with fibre volume fraction greater than 0.3, the shear mode gives lower
failure loads. Assuming there is no misalignment in fibres and no fibre bending stiffness,
Rosen calculated the critical load at which the fibres deflect into a sinusoidal shape. This
formulation was later confirmed by Jelf and Fleck (1992) by modeling composite
materials made from spaghetti rods in a silicone matrix.

Rosen's analysis (1965) has been used and modified in many studies, e.g. Hahn and
Williams, 1986; Johnson and Ellen, 1974, 1975a, b, 1976.

2.2.2. F i b r e C r u s h i n g

Fibre crushing occurs when the matrix is sufficiently stiff and strong to inhibit fibre
microbuckling. This mechanism occurs when the uniaxial strain in the composite equals
the crushing strain, which depends on the material properties and the geometry. Typical
value of crushing strain for Kevlar and carbon fibres is 0.5% (Fleck, 1997). Fibre
crushing in the form of kinking can also occur within fibres of small width (Piggott and
Harris, 1980, Gibson and Ashby, 1988).

2.2.3. Splitting

Under uniaxial loading, this mechanism occurs when the stiffness of the matrix is greater
than the stiffness of the fibres and when the composite has a low toughness and high
porosity (Kaute et al., 1996). It is a tensile cracking mode developed from imperfections
such as voids or inclined flaws. This can happen in compression as well, where under
increasing compressive loads tensile stresses around imperfections increase and the
specimen fails along a macroscopic shear band.

10
Chapter 2: Compressive Failure of Fibre Composites

2.2.4. Buckle Delamination

Delamination in unidirectional laminates mainly occurs when a surface layer is debonded


over a certain length. Debonding is due to one of the following reasons:

1. The manufacturing process

2. Surface impact

3. Out of plane loading by waviness of the fibres

Unlike multidirectional laminates, where delamination can occur frequently due to the
lay-up, in unidirectional laminates with no imperfections the occurrence of this failure
mode is rare. In fact, studies on delamination have primarily focused on the effect of
imperfections or the effect of impact on development of delamination in unidirectional
laminates (see for example Kim and Sham, 2000). Many studies also have been focused
on characterizing the post-buckling behaviour of delaminated laminates (Kardomateas
and Schmueser, 1988; Kardomateas, 1993; Gu and Chattopadhyay, 1995).

2.2.5. Shear Band Formation

This failure mechanism may occur for polymer-matrix fibre composites with very low
fibre volume fraction. Fried (1963) showed that failure occurs in a band oriented at about
45 with respect to the loading direction (Figure 2-1). This failure mode essentially
occurs for unreinforced matrix material and is not expected to happen for composites
with practical fibre volume fractions.

11
Chapter 2: Compressive Failure of Fibre Composites

2.2.6. Plastic Microbuckling or Kinking

This is the dominant failure mode for unidirectional composite laminates. The
compressive strength is controlled by plastic shear deformation in the matrix and fibre
misalignment. During the kinking process the supporting matrix material undergoes
plastic deformation. As a result, this failure mechanism is also called plastic
microbuckling.

In this section, a review of the studies on kinking process done in the literature are
presented. Also, micromechanics of various stages during kink band formation for both
thermosets and thermoplastics are discussed.

2.2.6.1. History of Kink Band Analysis

Kinking as one of the main compressive failure mechanisms has been extensively studied
in the literature in the past forty years. As mentioned before, Rosen (1965) presented a
simple formulation to calculate the compressive strength of elastic fibres embedded in an
elastic matrix, while neglecting the bending stiffness of the fibres. Argon (1972) extended
Rosen's formula by considering plastic yielding and misalignment of the fibres. In this
study, he derived the following formula for the compressive strength of composites
failing by kinking mode:

a = k /j
0 0 (2-1)

where, ko is the inter-laminar shear strength and ^ is the initial misalignment of the
0

fibres. Budiansky (1983) extended Argon's formula by assuming an elastic-perfectly


plastic material and proposed the following formula for kinking critical stress:

a = - (2-2)

in which, yo is the composite yield strain. Budiansky assumed that kink band inclination,
(3, is equal to zero (Figure 2-2). Based on this work, several studies were carried out to
measure the misalignment in the fibres. For example Yurgatis (1987) found that most of

12
Chapter 2: Compressive Failure of Fibre Composites

the fibres in a carbon fibre-PEEK unidirectional composite were oriented within 3 of


the mean fibre direction.

Figure 2-2 Kink band formation under compressive load for a unidirectional laminate, (a)
Before failure, (b) After failure.

Various improvements were made by subsequent studies to analyse the post-buckling


behaviour of the fibres, inelastic behaviour of the matrix, and imperfections. These works
provided an understanding of the factors governing the kink band initiation, inclination of
the kink band, height of the kink band, fibre rotation, shape of the assumed yield surface,
kink band propagation, and fracture toughness of this failure mechanism (Fleck and
Budiansky, 1993; Soutis and Fleck, 1993; Fleck et al., 1995; Kyriakides et al., 1995;
Moran and Shih, 1995, 1996, 1998; Sivashanker and Fleck, 1996; Sutcliffe and Fleck,
1994, 1997; Sutcliffe et al., 1996; Shu and Fleck, 1997; Vogler and Kyriakides, 1997,
1999, 2001; Budiansky et al., 1998; Sivashanker, 1998; Soutis et al., 1999; Bazant et al.,
1999; Berbinau et al., 1999; Niu and Talerja, 2000).

Several studies have been carried out to understand the micromechanics of this failure
mechanism. Guynn and Bradley (1989) compared the damage zone to a crack with a
plastic zone. They applied the Dugdale model (1960) to predict the size of the buckled

13
Chapter 2: Compressive Failure of Fibre Composites

region as a function of the compressive load for both carbon-epoxy and carbon-PEEK
laminates. They concluded that the Dugdale model, in which a constant stress was
assumed over the damage zone, is unable to accurately predict the compressive failure
stress. Following this study, Soutis et al. (1991) presented a cohesive zone model to
analyse the damage propagation and failure of the orthotropic laminates due to kink band
formation. In their study, the damage zone was represented by a line crack loaded on its
faces with a traction, which varies linearly with the crack displacement.

The experimental studies of Sutcliffe and Fleck (1994), Sivashanker et al. (1996) and
Fleck et al. (1997), based on strain measurement of the kink band, showed that the axial
stress across the kink band decreases with the distance from the band front and reaches a
plateau equal to about 50 percent of the maximum stress. Such a crack-like behaviour of
the band was then confirmed by Moran and Shih (1995). They also discovered band
broadening; whereby the damage zone grows under a constant remote stress.

Budiansky, Fleck and Amazigo (1997) presented a formula for the kink band critical
stress based on the kink band fracture energy by analysing the propagation of an out-of-
plane kink band (Figure 2-3). Bazant et al. (1999) showed that this formula does not
account for the compressive strength size effect of the specimen. In this study, they
calculated the fracture energy from the area under the stress-strain curve and above the
plateau stress based on the analysis of Palmer and Rice (1973). In fact in the Bazant et al.
study (1999), the formation of kink band was assumed to be the result of shear cracking
of the matrix, while neglecting the effect of the bending stiffness of the fibres. Finally,
they derived a formula to account for the specimen size effect on the compressive
strength.

Niu and Talreja (2000), reviewed the classical results of kink band compressive strength.
They reached the previous results by using a generalized Timoshenko beam model that
was developed for a two-dimensional periodic matrix-fibre-matrix laminate. They
presented a matrix shear instability based mechanism for kink band formation. They also
modified the Argon-Budiansky kinking formula to account for the combination of the
axial compressive stress and transverse shear perturbation.

14
Chapter 2: Compressive Failure of Fibre Composites
Chapter 2: Compressive Failure of Fibre Composites

2.2.6.2. Kinking Formation


Several researchers, such as Moran et al. (1995,1998), Liu et al. (1996), Sivashanker et
al. (1996), and Sivashanker (1998), have studied the micromechanical stages of kink
band formation process in unidirectional thermoset and thermoplastic matrix composites.
Various stages in the kink process were recorded by video camera, which is depicted in
Figure 2-4 for the formation of an out-of-plane kink band in the Moran and Shih study
(1995).

Displacement (d)

Figure 2-4 Overall behaviour of unidirectional fibre composites under compressive loads in the
fibre direction when kinking occurs (Moran et al., 1995)

The initial stress-strain response is linear elastic. However, at some point the matrix in
the highly stressed zone in front of the notch begins to fail and loading curve begins to

16
Chapter 2: Compressive Failure of Fibre Composites

show some non-linearity. The failure mechanism in the matrix depends on the type of
material, i.e. thermoset or thermoplastic. For thermoset materials, such as epoxy, matrix
cracking occurs under a combination of axial and shear loading. On the other hand, for
thermoplastics such as PEEK, matrix begins to flow plastically.

Moran et al. (1995) named this early inelastic deformation as "incipient kinking". The
incipient kinking is followed by the early stage of kink band formation, which involves
progressive fibre bending/rotation and plastic shearing or cracking of the matrix within a
narrow band. Fleck et al. (1995) showed that this initial damage height (Figure 2-5) is of
the order of 10-15 times a fibre diameter. This was also confirmed by Moran et al.

Damage Height

Figure 2-5 Initial damage height in kinking process (Fleck et al., 1995). Fibres rotate under
compressive load.

During these early stages, the peak or critical stress is reached. Experimental observation
(Moran et al., 1995, 1998; Liu et al., 1996) showed that the peak stress is typically no
more than 5 or 10% higher than the incipient kinking stress. This is because after the
matrix cracks, the fibres cannot carry any more loads due to instability. This leads to the
peak stress being reached just after the matrix cracks or yields.

The progressive matrix failure in the form of cracking or yielding and fibre bending leads
to a strain-softening response of the material within the band and the load subsequently
begins to drop. During this stage, at some point, due to the stiffening of the composite

17
Chapter 2: Compressive Failure of Fibre Composites

shearing response at large shear strains, the rotation of the fibres stops. This phenomenon
is known as "fibre lock-up", shown in Figure 2-6.

Time T< T T, Ti2 Ti ,


1
A

Rapid fibre i
Fibre rotation terminated by Gradual fibre bending and
rotation caused 1

matrix strain hardening rotation


by geometric (

softening i

Figure 2-6 Schematic showing the behaviour of a single fibre during the early stages of kink
band formation (Moran et al., 1995).

After the lock-up stage, the kink band starts to broaden. Band broadening is accompanied
by the rotation of the fibres, and matrix cracking or yielding moving towards the
undamaged interior material. During this stage, the kink band orientation remains fixed.

The final height of the kink band is set when the fibres snap (Sivashanker et. al., 1996).
Moran et al. reported a width of approximately 30-50 times the fibre diameter for the
final band height, while Sivashanker et al. (1996) observed a factor of around 100.

18
Chapter 2: Compressive Failure of Fibre Composites

2.3. Failure Mechanisms in Angle-ply Composites

Some studies have been done on the micromechanics of damage propagation in


multidirectional composites in compression (Soutis et al., 1993, 2002; Sivashanker and
Bag, 2001; Sivashanker 2001).

Sivashanker (2001) presented a detailed explanation of micromechanical behaviour of


three different multidirectional lay-ups of carbon-fibre epoxy composites. He identified
the compressive failure process as out-of-plane microbuckling of the [0], off-axis layer
damage and interface delamination between layers.

He observed that microbuckling in multidirectional laminates is accompanied by


delamination in the vicinity of the microbuckle zone. He reported the main feature of the
fractured multidirectional laminates as a zigzag pattern in off-axis plies, which he
classified as parallel splits along the fibre direction and out-of-plane kink band normal to
the splits.

For the band broadening stage, he reported a growth of delamination combined with the
growth of kink band height. In his study, microbuckling and delamination are never
presented independently, but always as parts of a damage unit. It is further proposed that
delamination is not a cause for microbuckling, but in fact is the consequence of it.

He also observed and measured the constant plateau stress in the band broadening stage.
He mentioned that this constant bridging stress is mainly associated with steady-state
band broadening in [0] layers and steady-state delamination crack growth under a
constant remote stress.

Microscopic images of the Sivashanker's study are presented in Figure 2-7.

19
Chapter 2: Compressive Failure of Fibre Composites

Figure 2-7 (a) Overview of arrested damage zone, (b) Band broadening in progress, (c) A close-
up view of the microbuckle band, (d) 45 degree off-axis microbuckle band linking
the 0/45 degree delaminated interfaces, (e) Out-of-plane 0 degree microbuckle band
extending into the 45 degree off axis layer. (Sivashanker, 2001)

20
Chapter 2: Compressive Failure of Fibre Composites

2.4. Summary

A brief review of studies on compressive failure of composites was presented in this


chapter. While extensive research has been done on the unidirectional laminates failure
under compressive loads to understand their micromechanical behaviour, little has been
done with respect to compressive failure of multidirectional laminates.

The work by Sivashanker (2001) indicates that compressive failure mechanism is a


combination of kinking, delamination and off-axis matrix cracking.

Lack of consistent models for unidirectional and multidirectional composites failure


under compressive loads is clear in the literature. Soutis et al. (1993, 2002) applied his
crack bridging model to analyse the failure of notched multidirectional laminates under
compressive loads. However, there is no detailed explanation on how the model relates to
the micromechanical behaviour.

Band broadening, which is an important stage in the compressive failure has not been
studied thoroughly in the past and there is no physical model to explain and account for
this stage.

The need to construct a model to analyse the compressive failure of multidirectional


laminates that could relate to the observed micromechanical behaviour such as incipient
kinking and kink band broadening is the driving force to build an analog mechanical
model of the response in the next chapter.

21
Chapter 3: Analog Model

3.1. Introduction

This chapter is mainly focused on constructing a physically meaningful "analog model",


that is able to simulate the behaviour of composite materials during the failure process in
tension, compression and under reversal loads. The main advantage of constructing such
a model is that it will enable us to explain and study the true behaviour of composite
materials under various loading conditions.

As mentioned in Chapter 2, a comprehensive model to simulate the micromechanical


behaviour of composite materials during the compressive failure does not exist. The main
reason for this is perhaps that various failure mechanisms exist for different loading
conditions and geometries. Even for a specific geometry and loading condition, the lay up
of the laminate plays an important role in the damage propagation process. All these
effects make it difficult to construct a physically meaningful model capable of capturing
the true behaviour of composite materials.

The method that has been used in this chapter to construct the analog model involves
incorporating the compressive behaviour of composite material during the failure process
into the existing tensile failure model, CODAM (Williams, 1998; Williams et al. 2003;
Floyd, 2004). Data from the literature was used to distinguish between compressive and
tensile failure mechanisms.

There are some primary differences between the failure mechanisms in tension and
compression. Tensile failure is due to fibre breakage and matrix cracking or yielding for
thermoset and thermoplastic matrix materials respectively (Floyd, 2004). Often however,
depending on the lay-up of a multidirectional laminate, the failure is a combination of
delamination and fibre breakage.

22
Chapter 3: Analog Model

For the compressive failure on the other hand, the primary failure mechanism is kink
band formation which is the formation of matrix cracking or yielding and fibre rotation
(McGregor, 2005). Like tensile failure, sometimes a combination of different failure
mechanisms such as kinking and delamination lead to the complete failure (Sivashanker,
2001; Sivashanker and Bag, 2001). But unlike tensile failure, as Moran and Shih (1998)
observed, there is little evidence of fibre breakage during the formation of the kink band.
Fibre breaking in compression was observed after the formation of the first kink band
during the band broadening process by Sivashanker et al. (1996) and Vogler and
Kyriakides (1997).

After the formation of the kink band, the damage zone, under further application of
displacement, propagates into the undamaged interior materials. This broadening of the
kink band which combines with fibre breakage and splitting results in a plateau stress
after the softening in the overall stress-strain response of the composite material. This is
another difference between compressive and tensile failure.

Upon unloading the damaged material, there will be a permanent deformation in the
compressive specimen which is not the case in the tensile specimen. The presence of the
constant deformation in the compressive damage process is due to the friction between
the new cracked surfaces and rotation of the fibres in the kink band. After rotation of the
fibres in the kink band, the new cracked surfaces will appear. Upon unloading the friction
between these cracked surfaces leads to plastic deformation.

The construction of the analog model has to be such that it will incorporate all these
behaviours of composites material under compressive loads using basic elements, such as
springs and sliders, which will be chosen based on the behaviour of composite materials
in the damage process. Besides slider and spring elements, we need some other basic
repeating elements in the model to simulate band broadening and plastic deformation as
described in the next section.

The combination of all elements will result in construction of two primary collections or
boxes of elements which can then simulate the overall behaviour of a composite material
under tension, compression and reversal loads. Each of these boxes represents a physical

23
Chapter 3: Analog Model

behaviour of the compressive failure such as matrix cracking as described in the


following sections.

After building the boxes, the model will be presented and its response will be examined
under compression, tension and reversal loads. The final section will be dedicated to the
model behaviour explanation.

24
Chapter 3: Analog Model

3.2. Elements

Each of the elements used in constructing the analog model will be described in this
section. It is noteworthy that in addition to the normal elements used in the literature,
such as springs and sliders, introduction of new elements has been necessary in order to
describe the behaviour of composite materials. The response of each element will be
described in detail.

3.2.1. Rigid Link

This element doesn't deform, hence called rigid, and transfers the applied load to the next
adjacent element. The governing equation for this element is given below.

F = F, (3-1)

where, F will be used to denote the applied load. A schematic of the element
behaviour is also shown in Figure 3-1 below. This element will be shown with a bold
solid line in the analog model.

'+00

Figure 3-1 The behaviour of a rigid link element under compressive and tensile loads.

3.2.2. Spring and T/C Spring

Spring element is the classic linear spring, which can be represented with the following
equation:

25
Chapter 3: Analog Model

F = kS (3-2)

where, k represents the spring stiffness. The behaviour of a spring is shown in Figure
3-2.

F
A

Figure 3-2 The symbol and behaviour of a spring element under compressive and tensile loads.

Another form of spring used here is the so-called Tension/Compression (T/C) spring,
which unlike a normal spring has different stiffnesses under tension and compression.
Schematics of element symbols and element behaviours are shown in Figure 3-3.

Figure 3-3 The symbol and behaviour of a T/C spring element under compressive and tensile
loads.

26
Chapter 3: Analog Model

3.2.3. Slider

This element is used to show the state of band broadening in kink band failure or plastic
yielding. In fact in the broadening state, under a constant load, damage propagates into
the interior undamaged area (Moran and Shih, 1996). This element has a rigid-perfectly
plastic behaviour. In other words, it acts as a rigid link until reaching a limit load and
then it yields and exhibits a zero tangential stiffness. This element is formulated as
follows:

F apP KP|<N) 5=0


S>0 (3-3)
-F.. S<0

where, F is the yield load. Schematics of the element symbol and behaviour are shown
y

side-by-side as follows.

Figure 3-4 The symbol and behaviour of a slider element under compressive and tensile loads.

3.2.4. Fuse and T/C Fuse

The fuse element is similar to the slider, but after reaching the limit load it fails and
cannot carry any loads. The definition of this element is necessary to model the matrix
and fibre cracking and breakage.

27
Chapter 3: Analog Model

T/C Fuse has two different values for compressive and tensile failures. This element will
be used to simulate fibre failure. The difference in tension and compression is because of
the fact that in compression a fibre can fail due to a variety of phenomena such as
instability, while in tension the failure is due to the tensile breakage. These two elements
are defined as follows:

Fapp {-Fc<F apP <FJ) 5= 0


0 {F =F )
app T
(3-4)

0 (F a p p =-F ) c 5<0

where, indices T and C denote limit states in tension and compression, respectively.
Schematics of element symbols and element behaviours are shown in Figure 3-5 and
Figure 3-6.
F
t
p_ p
T c

Figure 3-5 T h e symbol and behaviour of a fuse element under compressive and tensile loads.

FT M FT4= FC

-Fc

Figure 3-6 T h e symbol and behaviour of a T / C fuse element under compressive and tensile
loads. ,

28
Chapter 3: Analog Model

3.2.5. Gap

This element represents a distance, shown as 5 , between two elements. A gap element
cannot carry any tensile loads but if that distance is reduced to zero as a result of
compressive deformation, the gap can close and become a rigid link and thereafter can
carry only compressive loads. The definition of this element is necessary to simulate the
behaviour of the damaged material in compression. In compression, after the formation of
a damaged zone the material can carry a specific amount of load due to friction or
yielding in the damaged matrix. The governing equation of this element is as follows:

f0 8 >Se

F
= \F 8 =8 (3
" 5)

A schematic of element symbol and element behaviour is shown in Figure 3-7.

Figure 3-7 T h e symbol and behaviour of a gap element under compressive and tensile loads.

3.2.6. Lock

As mentioned before, one difference between compressive failure and tensile failure is
the presence of plastic deformation in compression in unloading. This requires us to
define the lock element. This element acts as a rigid link when the applied load is less
than a defined tensile load capacity, F . Once this tensile limit has been reached the
T

element splits apart and does not transfer any more load unless it is forced back to zero

29
Chapter 3: Analog Model

deformation by other elements. For this element we also define a gap, S , before the vg

load carrying capacity is reached. This element is defined as follows:

\F (F <F )
apP app v S=5 y

(3-6)
\0(F =F )
app v 5>S,

A schematic of the both element symbol and behaviour is shown below.

I F 4I
V

o
Figure 3-8 The symbol and behaviour of a lock element under compressive and tensile loads.

30
Chapter 3: Analog Model

3.3. Collection of Elements or Boxes

Element boxes are simply a collection or combination of previously defined elements in


parallel or series and are able to model more complex behaviour. Two boxes are currently
defined in the analog model, called the 'laminate box' and the 'rubble box'. Each of these
boxes is used to represent one aspect of the physical behaviour of the composite during
damage; the laminate box models the behaviour of the intact laminate and the rubble box
represents the behaviour of the damaged material in compression. The analog model can
then be constructed as a combination of these two boxes and some other simple elements.
In this section a mathematical representation of the boxes is presented.

It is noteworthy that the laminate box is an analog model which reacts differently under
compressive and tensile loads. The rubble box on the other hand, only reacts under the
compressive loads.

3.3.1. Laminate Box

To model the response of the intact laminate as damage progresses the laminate box is
introduced. As discussed in Chapter 2, as damage progresses in the laminate the overall
stress-strain response of the laminate shows a softening behaviour, which is different
under tension and compression. Under a tensile load, the softening is due to the reduction
in effective area due to fibre and matrix cracking. In compression, on the other hand, this
softening happens due to the matrix cracking or yielding and fibre instability. In addition,
the instability and rotation of the fibres combined with fibre cracking and splitting
contribute to the overall softening behaviour under compression. In fact, in compression,
matrix cracking leads to the instability of fibres which is not the case in tension.

The overall reduction in the laminate tangential stiffness which forms the softening
behaviour has been modeled using T/C fuses and T/C springs. The choice of T/C spring
elements is based on the idea that the composite laminate acts differently under tension
and compression.

The laminate box consists of a collection of infinite number of basic units in parallel as
shown in Figure 3-9. Each basic unit consists of two T/C spring elements and two T/C
31
Chapter 3: Analog Model

fuse elements as shown in Figure 3-10 below. One of the T/C springs and T/C fuses in
series, denoted as " m " , simulates the response of the matrix, while the other series, "f\
represents the behaviour of the fibres. The element notations are similar to what was
presented before.

m f m f m f
777777777777777777777777777777777777777777777777-
1 2 n

Figure 3-9 Laminate box and the arrangement of basic units. A l l the elements are T / C springs
and T / C fuses.

777777

Figure 3-10 Basic unit as used in the laminate box. "w" represents the matrix response and " / '
represents the fibre response in a representative volume element.

32
Chapter 3: Analog Model

The overall response of the basic unit can be easily obtained and is shown as follows.

f fails

Figure 3-11 Force-displacement response of a basic element in the laminate box.

3.3.1.1. Laminate Box Equations

The equations describing the behaviour of the matrix response under tensile loads will
now be derived as an example. Other responses, such as fibre under compressive loads
can also be derived following the same procedure. First, the springs and fuses for the
matrix and the fibre are grouped together and labelled as "w" and "f \ respectively, in the
following figure.

33
Chapter 3: Analog Model

m f

FT

\//777777777777777777777777777777777777777777777T
l\ 2 3 A 1 2 3 n

Figure 3-12 Grouping of matrix and fibre elements in the laminate box under tensile loads.

To derive the appropriate equations for group m, the following simplified notation is
used:

n = number of springs
= k = spring stiffness (m)
T
m

F, = Fl = tensile strength of the first fuse (m)

AF = AF = increment of the tensile fuse strength (m)


m

8, = 5] Lm = deformation at first (initiated) spring failure (ni)

=5 T
SL = deformation at last (saturated) spring failure (m)

The response of the finite number of elements in group m is therefore given as follows:

n(k )Se 5<5,


{n-q\k )S t S,<S<S S (3-7)

0 8 > S..

where, q is the number of failed springs.

34
Chapter 3: Analog Model

As an example, the general response for four elements is shown in Figure 3-13.

F = nke5, + q=l

q=2
q=3

Figure 3-13 The response of group " A M " under tensile loads in which the group consists of four
elements.

We can also replace q by ^'J^g- After substituting this value and further

simplification, we arrive at the following equation for the response of a finite number of
elements in group m.

f r p
5
A S <S<8 (3-8)
F k.8 r S
AF/

The next step is to determine the response of group m as the number of elements
approaches infinity. This can be easily done by taking the limit of the Equation 3-8 when
n approaches infinity. The result is:

F = k 5, <S <5 S (3-9)

where, k represents the total stiffness of all elements.


tot

By defining the following constants, the total response shown in Figure 3-14 can be
attained.

35
Chapter 3: Analog Model

a = nk = k
e m

c = K s, t

c / \ a
\
5

8, 5 S

Figure 3-14 The response of group under tensile loads in which group consists of an infinite
number of elements.

The response of group / is identical and we only need to replace the corresponding
parameters. The overall response of the laminate box in tension can then be found by
taking into account that m and /are in parallel. The constitutive equation of the laminate
box under tension is shown below.

k 8 + k 5
"- m
, L
f

8 + kl S
-si m J
f
8l L -5 ^
(
S T

F= 8 + k] si <s<si, (3-11)
L, -L
s
n
5
-si f J

0 + kj L
L,
, 8, S L
< S < S
S

'lLf
V V S

S L
f

5 > Si

where,
k\ = nk e = Total initial tensile stiffness (m)

36
Chapter 3: Analog Model

8T
U = Tensile deformation of first spring failure (initiation) (m)
8' SI = Tensile deformation of last spring failure (saturation) (m)
k\ L f = nk ^ = Total initial tensile stiffness if)
e

Sj L = Tensile deformation of first spring failure (initiation) if)

S' sl = Tensile deformation of last spring failure (saturation) if)

In the overall response, when the loads are applied the material initially shows a linear
response, which indicates that there is no damage. Then, by increasing the load damage
starts at some point. Usually the matrix cracking in tension occurs before the fibre
cracking. This means that the damage initiation strain for matrix in tension is smaller than
that of the fibre's. Therefore, after the load reaches the matrix damage initiation strain the
overall response becomes a nonlinear function. This nonlinear function is simulated by
using a parabolic function in the analog model as a consequence of our previous
assumptions. After damage initiation, the damage propagation results in decreasing
tangential stiffness and the overall softening behaviour.

For the above equations, considering the fact that matrix damage initiation strain is less
than the corresponding value for fibre and assuming that the matrix damage saturates
before the fibre is damaged, the total response of the laminate box and each element
group under tensile loads (displacement controlled) can be shown schematically in Figure

3-15.

Figure 3-15 Force-displacement response of the laminate box under displacement controlled
tension, m and /represent the response of the matrix and fibre respectively.

37
Chapter 3: Analog Model

To obtain the response of the laminate box under compression the same procedure can be
followed. Therefore, assuming the box under compression shown in Figure 3-16, the
constitutive equations are derived and presented as Equations 3-12.

m f

F c

1 2 3 n 1 2 3 n

Figure 3-16 Grouping of elements in the laminate box under displacement controlled
compression.

kf, S + kn .8 8<
U
SL ,
8 + kf 8 L f <8<
si m J
\ 8% -8
8 -8
( C f

F= U
SL, 8 + k' <8< (3-12)
m J
sfLfJ

(
S c
-5- ^
0 + k, <S< 'SL
f )

'SL <8

The following parameters have been used in the above equations:

kfL = Total initial compressive stiffness (m)

38
Chapter 3: Analog Model

Sf, = Compressive deformation for first spring failure (initiation) (m)

Sg, = Compressive deformation for last spring failure (saturation) (ni)

kf, = Total initial compressive stiffness (J)

5fL = Compressive deformation for first spring failure (initiation) (f)

Sg, = Compressive deformation for last spring failure (saturation) (J)


f

Schematics of the response of each element group and the laminate box under
compressive displacements are shown in Figure 3-17 and Figure 3-18, respectively.
Similar to the response under tension, the overall response is simply the sum o f the
matrix and fibre responses.

It is also noteworthy that in compression, unlike tension, the effect of matrix cracking or
yielding on the overall response is much more pronounced than the effect of fibre
cracking. In other words, the loss of modulus due to matrix cracking is more than the loss
of modulus due to fibre cracking. The reason for the greater role of matrix cracking in
compressive failure is the increased fibre instability after the matrix is cracked, i.e. the
loss of transverse matrix support leads to the rotation and failure of fibres. This is shown
in the following figures.

Figure 3-17 Individual responses of Group m and Group/of elements in the laminate box under
compression. Group m represents the response of the matrix while Group /
represents the response of the fibres.

39
Chapter 3: Analog Model

Figure 3-18 Force-Displacement response of combination of Group m and Group / i n parallel,


and subjected to displacement-controlled compression.

It is also desirable to study the response of the laminate box under reverse loading. In
such cases, the strain-stress response of a damaged specimen would be different in many
aspects from the undamaged specimen.

First difference is the loss of modulus in reverse loading. After one element in the box
fails, the total stiffness of the laminate box decreases and it results in a strain softening
behaviour of the laminate box. Upon unloading and reloading in the reverse direction, the
failed element cannot carry any load. This means that after damage propagation, the
modulus of the laminate box decreases permanently.

Another difference is the change of the damage initiation strain in the damaged specimen.
The laminate behaves linearly up to the previous strain in the cycle at which unloading
took place. After reaching that strain, a new damage area propagates in the laminate and
the response of the laminate follows a strain softening curve. In the laminate box, this
behaviour has been simulated by using T/C fuses. Because in the previous loading cycle,
all the fuses between the damage initiation and the strain at unloading failed; the new
damage initiation strain would coincide with the strain value at the previously unloading
point.

40
Chapter 3: Analog Model

Lastly, the peak stresses are different. Upon re-loading a previously damaged specimen
the peak stress drops. This drop is due to the fact that the stiffness of the damaged
material is less than that of an undamaged material. In fact, the drop of the peak stress
along with the change in the damage initiation and laminate modulus all show that the
new damaged material is a softer material compared to the undamaged material (see
McGregor, 2005 for more details).

41
Chapter 3: Analog Model

3.3.2. Rubble B o x

The rubble box represents the behaviour of the damaged material in compression. Upon
reaching the damage initiation strain under compressive loads, the matrix starts to crack
or yield resulting in the formation of rubble and cracked surfaces. After the damage
propagates under further compressive loads, the friction between the newly cracked
surfaces causes an increase in load carrying capacity of the specimen. This damaged
material can carry compressive loads. This phenomenon, however, only occurs for
damage propagation under compression.

The behaviour of the rubble box can be represented with an analog model consisting of
gap and spring elements to model the compressive load carrying response with an
increasing stiffness. The gap is active before damage initiation. After damage initiates
new surfaces form resulting in an increased stiffness of the damaged material. The rubble
box cannot carry any load in tension as described above. This model is equivalent to an
infinite number of parallel gaps and identical spring elements to represent the stiffening
response of the compressed rubbles as illustrated in Figure 3-19.

///////////

Figure 3-19 A schematic of the Rubble Box and the basic elements inside it. .

For this box, it is assumed that the gaps are different in size, but the difference between
two neighbouring elements, i.e. the interval, is constant. Also, all spring elements are

42
Chapter 3: Analog Model

assumed to have an identical stiffness, k . Therefore, the interval and the total stiffness of
e

the rubble box, equal to the sum of the stiffnesses of all the springs, are given by:

A5 = (S -5 e e )ln
h b j (3-13)

K, = h n

Thus, we can write the following equations for the rubble box, in which j is the number
of springs for which the corresponding gaps have been closed:

7 =0 F,=0

J =l Fj=k AS e

(3-14)
j= 2 F, = k AS + 2k AS
J L

j= n F, = kAS + 2k AS +... + nkAS

where, the corresponding forces, Fj, are effective when each gap is closed. Therefore, in a
generic step we can write down the following equations for the load and the total
displacement:

F =J0 + \)/2xk AS
j t

(3-15)
S=S +JA5

Using Equation 3-13 and Equation 3-15, we can write:

j = (S-S )/AS
g] = n(S-5 )/(5 -S )
g] gn g] (3-16)

Now by combing the above equations, we can write the response of the rubble box for a
finite number of gap and spring elements under compressive loads, as follows:

(3-17)

43
Chapter 3: Analog Model

The constitutive equation of the rubble box for an infinite number of springs can be easily
obtained by taking the limit of the above equation when n approaches infinity. The result
is given below.

0 5 < \8 n

SIR < \ \ < SR


5 5
(3-18)
2
(8 R-8 )
S m

k S
KU
-(S SR +S) IR 5 > S SI

where:
SIR: Initial Rubble Deformation = 5 V
*i

SSR: Saturation of Rubble box = 5


ksR.- Final Rubble Stiffness = k
A schematic of the rubble box behaviour and the manner in which its stiffness increases
as the gaps close is shown in Figure 3-20.
F
->SR HR

Figure 3-20 Force-displacement and stiffness-displacement curves for an infinite number of


units in the rubble box

An issue that needs clarification is the role of the rubble box after saturation. After the
saturation, which represents the maximum damaged area in the specimen being loaded,
upon further loading the damage propagates into the interior undamaged material. In
other words, unlike tension, the damage height starts to grow under compressive loads.

44
Chapter 3: Analog Model

During unloading, the damaged material becomes uncompressed. This leads to a decrease
in the friction force between the cracked surfaces, i.e. some jagged surfaces become
disengaged. This behaviour has been simulated by opening of gaps after unloading.

45
Chapter 3: Analog Model

3.4. Analog Model

In this section, the analog model capable of modeling the behaviour of composite
material under compression, tension and load reversal is represented. This model is
assembled using the elements and boxes that have already been defined. Each element
simulates an aspect of composite micromechanical behaviour during the failure process,
which will be described. We will also introduce the model formulations for various
loading conditions.

3.4.1. Model Construction

The proposed analog model of the composite material is shown below.

Figure 3-21 A schematic of the analog model which shows arrangement of elements and boxes.

In this model, a rubble box and a lock-spring element are placed in parallel, and in turn
both are placed in series with a slider element to form the left part of the model shown in
Figure 3-21. This part of the model is also in parallel with a laminate box.

A summary of the analog model elements is given here. The laminate box simulates the
behaviour of the intact laminate in the process of damage propagation and shows the

46
Chapter 3: Analog Model

strain-softening behaviour of the composite material under tensile and compressive loads.
Group m and Group / of elements inside the laminate box represent the response of
matrix and fibres, respectively. As discussed before, the response of the matrix and fibres
in tension and compression are different and this difference is modeled by T/C spring
elements. Matrix failure ( matrix cracking or yielding) is also simulated by fuse elements.

The rubble box simulates the response of the damaged material under compression. After
damage initiates, further loading causes more new cracked and jagged surfaces, which
results in an increased friction force between the surfaces. This increased force is
modeled by using a combination of gap-spring elements in parallel.

The lock-spring element is used to simulate the plastic deformations under compressive
loads. This deformation occurs due to the rotation of the fibres in the kink band. After the
fibres rotate, the friction between the cracked surfaces doesn't allow the fibres to rotate
back to their original direction upon unloading. This phenomenon only exists under
compressive loading because there is no fibre rotation in tensile failure.

The response of the slider is related to the response of the rubble box because we
assumed that the saturation of the rubble box leads to yielding of the slider. After the
saturation of the first kink band, the kink band propagates into the interior undamaged
material to form additional kink bands. As mentioned in the previous chapter, many
experiments (Moran et al., 1995; Liu et al., 1996; Moran and Shih, 1998) have confirmed
the band broadening phenomenon.

The slider simulates the damage band propagation. After application of compressive load,
an increase of force in the rubble box and spring leads to activation of the slider, since the
yielding strain is assumed to coincide with the saturation strain of the rubble box. This
assumption also means that the saturation of the damaged zone for a band with constant
height, results in the damage propagation into the undamaged interior material and an
increasing damage height for compressive failure. It should also be noted that the left part
of the model cannot carry any tensile loads. In fact upon loading in tension, only the
laminate box is load bearing.

47
Chapter 3: Analog Model

As described before, upon unloading after the propagation of the damage and loading in
the reverse direction, the modulus of the laminate decreases and the damage initiation
strain moves further (see McGregor, 2005 for more details). This behaviour is modelled
by using the T/C fuses.

3.4.2. Model Formulation

Formulations of the rubble box and the laminate box were presented in the previous
section. The formulation for the lock-spring element can be easily obtained and is
presented below. A schematic of the lock-spring response is also shown in Figure 3-22.

(3-19)

3 1

Figure 3-22 Force-displacement response of Spring/Lock elements in series.

48
Chapter 3: Analog Model

For the slider element, the following formulation has been derived, based on the
aforementioned assumption that the saturation of the rubble box and yielding of the slider
coincide.

F =(^f y + k)(S -S ) SR m (3-20)

Having the formulation for all the elements and taking advantage of their arrangement in
parallel or series we can derive the constitutive equations for the analog model. This is
done based on the load applied to the analog model. To simplify the formulation, the
following simplified notation is assumed:

5<
f = l 5
m = l
5
f
=5
*g = 5
I R

5
S = SL,
5
= SR
S

S
'f ><'f
=5

S
U L = S
S L f

kf, = very small

kf = k u

k, =
k'n

k T
- k T

3.4.2.1. Load Reversal: Compression-Unloading Before Band Broadening-Tension


The load can be divided into the following stages, and the corresponding formulations
can be derived accordingly:

1. Loading in compression:

kfS \S\<\5)c

kfS^ 5 + k{S -Sf ^^l )+


k
\8f\<\8\<\S \
UN
( 3
" 2 1 )

49
Chapter 3: Analog Model

2. Unloading:

k 5 k(5-5f)
c
M + + ^-x^^- \sfU\sU\S,
F = 2 (S'-Sf) (3-22)

{k S + k(S-Sf)
c
M Sj<S<S P

where,

k =kf(5f-5 )/(5^-5f)
c
M UN

5 =kit< kS

H~ k
CM

3. Loading in tension:

Based on the modifications which account for failed elements, the final equations for
loading in tension can be written as follows:

k 5 + k(5-5f)
c
M
0<||<|cv|

k
' M f
5 + k
l i f
5 0 < 5 < 51

f
51-5-
F = 5 + k'L 5 Si, <5<8 l
Ui
(3-23)
51-5]
m J

\ f ... \
5' -8 S' -5
6+
s
5' <5<5'
s
k] Mr s

K Sl-5]
\ m J
s
'f J

0 5>5l

After the damage propagation in compression, the total modulus decreases. Therefore
upon further loading in tension, the modified modulus and damage initiation strain of the
model in the above equations are given by:

k T
M =kj {5 -5 )l(5 -5f)
c
s UN
c
s

m m
k
L = J ( ui
f
k
f
s
- UN) 5
1(8 UL - 5f)
k T

<8 -Sj T
S ) (3-24)

J
k

S
' M = S
S - J
J k

50
Chapter 3: Analog Model

From the above formulations, the force-displacement .diagram for the analog model can
be schematically shown in Figure 3-23. For this diagram, it is assumed that the fibre and
the matrix damage initiation strains in compression are identical.

IF

Figure 3-23 Force-displacement response of the analog model. First compressed, then unloaded
and finally loaded in tension to complete failure.

3.4.2.2. Load Reversal: Compression-Unloading After Band Broadening-Tension


After band broadening, the matrix fails completely. Therefore the corresponding
equations can be derived accordingly:

/. Loading in compression:

kfS 8\<8f

F = kf 7 7 8 + k(8-5f) + ^ . x ' \8f\<\8\<\8^\ (3-25)


v y y

I 'I I I I M
r

((8f-8f) 2 (5 -8f) c
s

(^ + k)(8^-8f) \sc\<\S\<\5 UN

2. Unloading:

F = k(8-8 ) ^,-^4-
o P
rxc+
xc\
\5\<W <\5
r f r r n r w i U < " )
3 26

2
(S ~ l )

51
Chapter 3: Analog Model

where,

S
P= UN S
-( S
5
~ f)5

3. Loading in tension:

Equations for the analog model response under compressive loads upon unloading from
tension are derived and presented below. A schematic of the analog model response is
shown in Figure 3-24.

k(S-5 ) P 5 -5,
P <5 < 5 P

0 Q<\6\<\8 P -sy
F 0<8<8' (3-27)

5\ -8
8' Mf <8<S' S

v ' s
'/ J

0 5 > Si

Using the same procedure as in the previous loading case, the modified modulus and the
new damage initiation strain in compression are derived as follows.

k' =0

L,
k
= , (SUL - 8UN )!(8 - 8f) (3-28)
UL

kl
8'L =tf-- (8l-8f
L
)
J k, J
'f

Figure 3-24 Force-displacement response of the analog model (Unloading after the yielding).

52
Chapter 3: Analog Model

3.4.2.3. Load Reversal form Tension to Compression


The load can be divided into the following stages, and the corresponding formulations
can be derived accordingly:
1. Loading in tension:

k] 8 + ki 8 8<5]

F= 8 + kj 5 s; <s<s; (3-29)

51-5
8 + k], 5] <8 <8,j
S -Sj
T
S
\ s
f J

2. Unloading:

F = k' 8 M
0<8<8, UN (3-30)

where,
k' M =k] (5l-5 )l(S -5] UN
T
s )

k](5 -S )l{S -S] )


T
s UN
T
s r
f

kT
kM
1
+
+k K
T

M ,

3. Loading in compression:

Equations for the analog model response under compressive loads upon unloading from
tension are derived and presented below. A schematic of the analog model response is
also shown in Figure 3-25.

k5
M H<K | C

C\2
(5-5)')
k^5 + k(5-5f) + k |<|<5|<|^
c

2 (5 -5f) c
s

(5 -5) C
k (5-5y) C\2
F= _ - , ,^5-5f) + ^-x
s SR
5 <5<5- L
M
(3-31)
(5 -5f)
C
s 2 (5<-Sf)
(^L + k)(8C-Sf) \5s \<5 < 8 UN

0 5 UL < 5

53
Chapter 3: Analog Model

Using the same procedure as in the previous loading case, the modified modulus and the
new damage initiation strain in compression are given by:

Figure 3-25 Force-displacement response of the analog model. First pulled in tension, then
unloaded and finally loaded in compression to complete failure.

54
Chapter 3: Analog Model

3.5. Further Remarks on the Analog Model

1- Band broadening occurs under a constant stress. This so-called plateau stress has been
measured in previous studies in the literature. It is noteworthy that as Gupta et al. (1998)
and Sivashanker et al. (1996) have shown, at a certain strain, ultimate strain, the steady-
state band broadening is superseded by the sliding action of the two halves of the
specimen along one of the kink band fracture planes.

2- For a multidirectional laminate, the failure mechanism is different. As Soutis and


Spearings (2002) showed for a plain weave [45/0/90] laminate the main failure
mechanism is delamination, while for a [0/90] laminate kink band is the main failure
mechanism. The failure of a multidirectional laminate has been studied in recent years
(Soutis et al., 1993, 2002; Sivashanker and Bag, 2001; Sivashanker 2001). These studies
revealed that a combination of kink band and delamination lead to the laminate failure.
The band broadening process is simply the propagation of delamination and kink band
into the interior undamaged material.

3- Differences in the failure mechanisms of thermosets and thermoplastics were


considered in construction of the analog model. For thermoplastic materials, the
formation of kink band is due to the yielding of the matrix and rotation of the fibres. In
thermosets, on the other hand, the matrix cracking leads to the softening and rotation of
the fibres. The response of the thermosets and thermoplastics are also different under
tensile loads. In thermoplastics, unlike thermosets, the saturation of the matrix damage in
tension occurs after the saturation of the fibre damage. This phenomenon is due to the
yielding of the matrix.

4- Finally, it is noteworthy to mention that the analog model presented here has been
developed in collaboration with Carla McGregor. For a complete description of the
analog model and its response, including description of multidirectional laminate failure,
failure in thermosets and thermoplastics, and the formulation of analog model in stress-
strain domain the reader is referred to McGregor (2005).

55
Chapter 4: Damage G r o w t h , Energy Concept and Scaling L a w

To complete the construction of a comprehensive constitutive model, there are some


issues that need to be considered before implementing the analog model in a finite
element code. These issues, which are discussed in this chapter, are as follows:

1 - Compressive damage growth

2- Compressive fracture energy

3- Compressive scaling law

First a brief review of CODAM damage growth theory and energy concept (Williams,
1998; Floyd, 2004, McClennan, 2004) as it pertains to tensile loading is presented. Then,
a compressive constitutive model based on the analog model is developed that predicts
damage propagation under compressive loads.

At the end of this chapter, scaling laws for finite element modeling of compressively
loaded specimen are presented. This topic is explored in an effort to overcome the mesh
sensitivity problem, which can occur due to different element sizes used to model strain-
softening behaviour (Floyd, 2004; McClennan, 2004).

56
Chapter 4: Damage Growth, Energy Concept and Scaling Law

4.1. Damage Growth

One of the main objectives in damage modeling is to relate the damage growth to the
state of strain, using a single mathematical expression. In the analog model, following the
procedure used in CODAM (Williams, 1998; Williams et al. 2003; Floyd, 2004), a linear
relationship is assumed between the potential function, F, which represents the state of
the strain as a potential force for damage growth, and the damage parameters, associated
with matrix cracking or yielding and fibre breakage. The state of damage, which is a
linear function of fibre and matrix damage will then be used to calculate the modulus
reduction in each stage of loading.

It is also important to note that damage is an irreversible process. In other words, the
effect of damage on matrix and fibre cracking and breakage is permanent and on load
reversal, damage will continue to grow from the previous state. As a result, damage
parameters increase monotonically. Therefore, the key to relate the stage of loading in
compression or tension to the reversal loading is to properly transfer damage parameters.

Now, to define damage growth in the analog model, following the CODAM definitions
(Williams, 1998, Floyd, 2004) the potential function as an equivalent strain is given by:

f c \
7 y
F(e) = +
X
(4-1)
K T u
\ ^J V ^J

Therefore in each stage, the damage parameter for the matrix and fibre under
compressive and tensile loading are as follows:

F(s)-F(s) c
I
Matrix damage in compression (4-2)
F(sf,

F(s)-F(s) T
l
m_ Matrix damage in tension (4-3)
F(s) -F(s) ,
T
s
T

57
Chapter 4: Damage Growth, Energy Concept and Scaling Law

F(s)-F(ef, _.. ,
r (4-4)
c
co: = Fibre damage in compression

F(e)-F(e) T
l

co , = 1
Fibre damage in tension (4-5)
1
F(s)' -F(s) ', f
s
T

where in all the equations, the damage parameter is related to the initiation and saturation
potential functions.

In the next step, damage due to fibre and matrix saturation (see Floyd, 2004) is defined as
follows:

m f
(4-6)
as =\-co s
m f

To obtain these parameters, we have to consider the failure mechanism of the laminate.
For example, assume a unidirectional lamina under compressive load in a displacement-
control test. The threshold of damage is either shear cracking (thermosets) or plasticity in
the matrix (thermoplastics), followed by formation of the kink band. At the onset of
saturation of the kink band, the lock-up stage, almost the entire confined matrix in the
kink band is damaged and therefore damage due to the saturation of the matrix will be
significant. On the other hand, considering the same specimen under the tensile load, the
matrix damage in the 0 degree layer is negligible (see Floyd, 2004).

For other lay-ups, the comparison of the role of the matrix and the fibres in compressive
and tensile failure modes specifies the compressive damage saturation parameters. For
example, in a 45 degree layer, under tensile load, the damage due to matrix saturation
would be around 50%. For compression, however, this parameter would depend on the
failure mechanism. For an oblique kink band, the saturation would show a significant
damage in the matrix, while for delamination the'matrix damage is not pronounced.

The damage parameter in the lamina or sub-laminate is a combination of damage in the


fibre and matrix. The percentage of matrix .damage or fibre damage in the final

58
Chapter 4: Damage Growth, Energy Concept and Scaling Law

combination depends on the damage due to fibre and matrix saturation. This can be
shown by a linear equation as follows:

oh = oh ah + ah]' co
s m
c
f For compression (4-7)

of = col o)l + 0)1 o)] For tension

These two parameters as a function of the potential function are depicted in Figure 4-1. It

is also noteworthy that damage due to matrix saturation, oh , in compression is much


m

higher than in tension, col . As explained before, this is due to the fact that in the
m

kinking process in compression after the matrix damage saturates, almost all of the
volume of the laminate is damaged. In other words, the saturation of the matrix damage
coincides with the saturation of the kink band formation while this is not the case in
tension.

Figure 4-1 Damage growth-potential function) diagrams for tensile and compressive loading.

Now, the relation between damage and modulus reduction in compression and tension, or
the effect of damage on elastic constants can be easily obtained. The difference between
tension and compression is that, in compression modulus reduction due to matrix damage
is more pronounced than fibre damage. This is because after matrix damage in
compression (not considering the steady-state loading) there will be no support for fibres
and they will lose their load-carrying capacity. On the other hand, this is not the case in
tension because the tensile strength of the fibre is larger than that of the matrix, therefore
after the saturation of the fibre damage the specimen does not carry any further load. This

59
Chapter 4: Damage Growth, Energy Concept and Scaling Law

is also shown in the following figure where RE is the ratio of remaining modulus to the
initial modulus.

T T C C

Tension Compression

Figure 4-2 (Modulus reduction-damage growth) diagrams for compressive and tensile loading.
Modulus reduction due to matrix saturation is more pronounced in compression
compared to the corresponding reduction in tension.

In summary, the first step in calculating the stresses in a damaged laminate is to find the
potential function. Damage parameters can then be calculated from the potential function.
Finally, modulus reduction due to damage can be obtained from the damage parameters.
The following equation is used to calculate the stress (Williams, 1998; Williams et al.,
2003; Floyd, 2004):

cr = R Es
E (4-8)

in which, the modulus is modified by the modulus reduction factor, RE.

60
Chapter 4: Damage Growth, Energy Concept and Scaling Law

4.2. Compressive Fracture Energy

Defining the fracture energy under compressive loads becomes an important issue in
studying the size effect. Due to the complexity of the damage mechanisms in
compression it is difficult to determine the fracture energy involved in the progression of
damage.

From the previous sections it is clear that damage grows in a bidirectional way. One
direction is associated with the transient damage growth and the other one with the
damage band broadening. In the analog model, this behaviour has been captured by
using the laminate box and the slider. While the former represents the transient damage
growth and formation of the first kink band, the latter represents the damage broadening
normal to the transient damage which will lead to the formation of parallel kink bands.

In other words, there are two forms of fracture energy involved in compression. One is
the main fracture energy corresponding to the formation of the first kink band, which
governs the peak stress and the size effects. The other one is band broadening fracture
energy, which is required to form additional kink bands or to expand the damage band.
This is depicted in Figure 4-3.

o Area associated with the formation of the first kink band


Formation of second kink band
Formation of third kink band

e
Aplastic

Figure 4-3 A schematic of stress-strain response of composites under compressive loads


showing areas associated with formation of kink bands and damage band
broadening.

61
Chapter 4: Damage Growth, Energy Concept and Scaling Law

4.2.1. Main Fracture Energy

As Bazant & Planas (1998) showed, in the strain softening response of a quasi-brittle
material, the strain energy release rate is the area under the stress-strain curve limited to
the unloading path from the peak stress, as illustrated in Figure 4-4. Then, the fracture
energy would be the area multiplied by the damage band height. This height is assumed
to be a constant value before the matrix damage saturation and then starts to grow with
the formation of additional kink bands.

o
A
Gpeak

Aplastic ^initiation ^saturation

Figure 4-4 Calculation of the main fracture energy from the area under the stress-strain
response of the laminate box.

This area is the fracture energy required to saturate the matrix cracking (or plasticity for
thermoplastics) and fibre rotation in compression.. As stated before, this is the required
energy to form the first kink band. Note that the small amount of energy associated with
the fibre splitting and cracking is negligible compared to the matrix cracking energy.

62
Chapter 4: Damage Growth, Energy Concept and Scaling Law

4.2.2. Band Broadening Fracture Energy

In this section, the fracture energy associated with the band broadening state is

considered. As described in the previous sections, no energy is associated with the rubble

box. In fact, the rubble box can be interpreted as the elastic response of the cracked

rubbles. Slider is, however, different and its activation is followed by some plastic

deformation, which is an indication of energy dissipation. One difference with the main

fracture process is the damage band height; the formation of additional kink bands leads

to the growth of the damage height. The calculation of the fracture energy will be

discussed here. Note that the effect of the fibre splitting is assumed to be negligible again.

At each strain after the matrix damage saturation strain, the unloading path is exactly

similar and parallel to the loading path. The unloading path at the saturation of each kink

band is shown in Figure 4-5.

Saturation of the second kink band

Saturation of the i kink band


th

Saturation of the first kink band


Unloading

"plateau

Figure 4-5 The unloading path after the saturation of each kink band. Only the stress-strain
response of the rubble box is shown here.

By assuming that the intervals between kink band saturations are the same and that the

heights of all kink bands are equal, the energy per unit area required to form the i th
kink

band can be calculated as follows:

(s -s )
G, = cr lateau x (s - s )xh = cr
c x U L s
xhc (4-9)
n-l
63
Chapter 4: Damage Growth, Energy Concept and Scaling Law

And the total fracture energy is the sum of all kink band energies which is calculated as
follows:

G =G,x(n-\)
Jh = G p l m m u x (s UL - s )xh s c (4-10)

where, the saturation strain, s , and ultimate strain, e , are the same as what we had in the
s UL

analog model. This energy is required to form additional kink bands, i.e. to broaden the
damaged zone into the interior undamaged material.

It is also noteworthy that at each strain, by using Equation 4-10, we have the following
equation to calculate the broadening fracture energy:

G =0- x(- )xh


Jb platem s c (4-11)

and we also have:

= (4-12)

where h and s, are the parameters related to a specific point in the stress-strain curve
response. Therefore:

r - (h-h )x(e -e )c UL s _ (h-h )x(n-Y)s


c inc

^ lb ~ ~ plateau* , , X
"c ~ plateau X
, 1 X I
X
K

Ki-K (n-\)h c (4-13)


Gfl = plateau* ^nc X
( h
~ h
c )

which is another way to define broadening fracture energy at each strain. In this equation,
s, is the interval strain between complete kink band formations, depicted in Figure 4-5.
c

64
Chapter 4: Damage Growth, Energy Concept and Scaling Law

4.3. Compressive Fracture Energy in the Damage Propagation Process

In this section, an analytical calculation of the compressive fracture energy as a function


of damage band length is presented. Data from the literature (Jackson and Ratcliffe,
2004) is used to validate the analytical solution. As a result, the fracture energy concept
of the analog model, presented in the previous section, will be validated.

4.3.1. Compressive Damage Propagation

Jackson and Ratcliffe (2004) presented a study to experimentally measure fracture energy
for kink band growth in sandwich specimens with sharp notches. They used woven
carbon-epoxy composite in a [(0-90)(45)] 2 lay-up. A schematic of their test
configuration and the stress-strain response of composite material under compression is
shown below.

Figure 4-6 Stress contour in front of the notch and Stress-Strain response of one element.

For this specimen, under increasing remote stress, the stress intensity factor in front of the
notch increases. Therefore, as a result of an increase in stress intensity factor, the notch
tip stress reaches the critical stress (peak stress). At this moment, compressive damage

65
Chapter 4: Damage Growth, Energy Concept and Scaling Law

initiates. The propagation of the damage under increasing far field displacement, is
divided into three different stages as shown in Figure 4-7.

_ "peak
plateau

Figure 4-7 Stress contour in front of the notch while damage is growing. Damage height grows
from h to h and finally to /);;. The maximum damage height, lt , is the damage
c L vi

height at the ultimate strain.

These stages are:

1- In a failure process zone (FPZ) in front of the notch, the matrix cracks and the
fibres rotate to form the kink band. Stress in this zone decreases, until it reaches
the plateau stress. Damage height also increases form zero in front of the FPZ to a
characteristic damage height h , at the notch tip.
c

2- Damage zone propagates horizontally under the remote displacement. Meanwhile,


damage height also increases as a result of damage band broadening. The failure

66
Chapter 4: Damage Growth, Energy Concept and Scaling Law

band is divided into two zones: FPZ and band boarding zone, in which stress is

equal to the plateau stress a plaleau.

3- After damage height reaches the ultimate damage height, stress drops to zero.

Damage height does not increase in this section. Therefore, three different zones

can be distinguished: FPZ, band broadening zone and zero stress zone, where

damage height is constant.

4.3.2. Compressive Fracture Energy

Here, appropriate equations are derived to calculate the fracture energy required to

propagate the damage zone for the damage propagation mechanism explained above.

Assume f(L) is the damage height function in front of the notch, where L is the length of

the damage zone. The shape of this function depends on loading, geometry and material

properties. For an increment of length dL in the damage zone, according to Figure 4-8,

the strain energy required to propagate damage into new undamaged areas is (shaded in

Figure 4-8):

N L rt H-
dL

Figure 4-8 Damage height increment due to the damage length increment.

AU AUj am tj + AU d broadening
(4-14)
-> AU = AU, +AU 2

where Ui is the strain energy required to grow the damage over an incremental length dL

in front of the damage zone and U2 is the energy needed to expand the damage zone to a

new height.

67
Chapter 4: Damage Growth, Energy Concept and Scaling Law

To calculate Ui, the fracture energy of an infinitesimal element is used as follows:

AU. = G xbxdL = y xh xbxdL


f c c (4-15)

In which b is the laminate width and Gf is the main fracture energy, which can be
calculated from Figure 4-9:

y =G /h
c f c
"peak

"plateau

Figure 4-9 Calculation of the main fracture energy for one element from the stress-strain
response of that element.

To calculate the strain energy required to expand the damage zone, U2, first we have to
calculate the new damaged volume, shaded area, in Figure 4-8. This is calculated for
linear damage height function and also for generalized damage height function. After
finding the strain energy, the fracture energy can then be derived.

4.3.2.1. Linear Damage Height Function


The linear damage height function is derived below:

h(L) = axL
h(c ) = axc =h
h h c (4-16)

-^h(L) = ^xL
c h

where a, the constant of the linear function, is calculated from the boundary condition.
Therefore, the damage height increment, dh, due to the damage length increment, dL, is:

68
Chapter 4: Damage Growth, Energy Concept and Scaling Law

2dh = h(L + dL) - h(L) = ^x(L + dL)-^xL = ^xdL (4-17)

And the new damaged volume is:


h
Avolume = 2dh xLxb = xdLxLxb (4-18)
Cu

Therefore, the strain energy for the damage height increment, based on Equation 4-13,
can be calculated as follows:
h
AE/ = o-plaleau xs x^xbxLxdL (4-19)
c
2 mc

Now we can calculate the total strain energy as:


h
AU = AU +AU = y xh xbxdL
y 2 c c +a hlcm xs x^-xbxLxdL
mc

U(L) =\AU=\(y xh xb + o- c c plalemi xs im x^-xbxL)dL (4-20)

h L 2

U(L) = y xh xbxL c e +a lateau xs inc x^xbx


c
h 2

Therefore the total fracture energy can be derived as:

i i j K L 2

y xh xbxL
f n c c +a laleau xs inc x xbx
G(L) = "W= C
- 2
-
bxL bxL (4-21)
h L L
-> G(L) = Y x h + a c c laleau x^x- = h x(y + a c c xs mc x )
c 2 h 2c
h

4.3.2.2. Generalized Damage Height Function


Assuming that n is the order of the damage height function, a generalized damage height
function can be written as:

69
Chapter 4: Damage Growth, Energy Concept and Scaling Law

h(L) = \xL" (4-22)


c h

Using Taylor series (Appendix A) expansion, the damage height, dh, can be written as:

2dh = h(L + dL) - h(L) = \x(L + dL)" -L-x(L)"


(4-23)
nh,.
A- x [(!)" + n(L)->df,]- A - x A x (Z)-' dL
c,

Therefore, the new damaged volume is

A
,o,ume = \2dhxbxdL=\\ ^xiLy-'ch
(
xbxdL =^.L" xdL (4-24)

As a result, the strain energy increment is:

b x h,.
AU = AU + AU =y xh xbxdL + a ,
i 2 c c u m xs x inc -S-L" x dL (4-25)

And the total fracture energy is

U(L) _ \AU L"


G(L) = = h x (j + a xe ) (4-26)
bx L bxL
c c plaleau inc
{n + \)xc "h

70
Chapter 4: Damage Growth, Energy Concept and Scaling Law

4.3.3. Summary of Fracture Energy Functions

The damage height function can be derived for different values of n. The different
damage height functions and the corresponding fracture energies are summarized as
follows:
Table 4-1 Damage height and fracture energy functions' for different values of n in Equation
4-26.

Damage Height Function Fracture Energy Function


^xL K X
c
b inc

\xl} K (Yc + ^plalea,,Einc


X
X

c
b *?>>
2JL
K (Xc + plal.au ^inc
X U X

fb c

x(L) L"
(y + <j , ,
h c n

n K x x )
c
b V / c plateau (n + \)xc "h

4.3.4. Comparison with Experimental Results

The following parameters are used for comparison of the model with experimental
results. These parameter are taken from the literature for the material and lay-up that are
used in Jackson and Ratcliffe (2004):

Table 4-2 Material parameters used to verify fracture energy functions.

Parameter Symbol Amount Source


Young's Modulus E 57.2 GPa Soutis et al., 2002
Critical Stress Intensity Factor K
IC 35 MPaVm Jackson and Ratcliffe, 2004
Damage Initiation Strain
i -0.01 Soutis etai., 1993
Damage Saturation Strain E
s -0.03 Soutis etai., 1993
Length of failure process zone Cb 5 mm Soutis et al., 2002
Plateau Stress " plateau 20 MPa

The difference between the material used in all other experiments and the one by Jackson
and Ratcliffe (2004) is the lay-up, which is woven in the latter experiment. Because of

71
Chapter 4: Damage Growth, Energy Concept and Scaling Law

this difference and based on the Sivashanker experiment (1998) who measured the
plateau stress to be 100 MPa, here the plateau stress is assumed to be 20 MPa. This
difference is due to misalignment of the fibres in the woven laminate. Also, the thickness
of the laminate in Sivashanker was twice the laminate thickness in Jackson and Wade
study (2004), which is another reason for reducing the plateau stress.

By using the above parameters and assuming the e, = -0.05, we can plot the following
c

diagram for critical stress intensity factor in which a parabolic function, better correlates
with the experimental results .

150

Experiments

100 V

x
OH

50

Linear Strain Field Parabolic Strain Field

-4 1 1 (- -4 1_
10 20 30 40 50 60

Damage Length (mm)

Figure 4-10 Comparison of experimental results (Jackson and Ratcliffe, 2004) and analytical
solutions for critical stress intensity factor as a function of damage length.

72
Chapter 4: Damage Growth, Energy Concept and Scaling Law

4.4. Scaling Law

Following the element height scaling in tension (Floyd, 2004) the scaling law for
compression is constructed here. Two methods are considered for scaling the element
stress-strain curve for different element heights: Crack band method and modified crack
band method, both of which are briefly explained here.

It is noteworthy that, only the response for the first kink band formation is scaled here. In
fact, the change in the height of the element doesn't require scaling of the plateau stress.
Therefore, for different element sizes, only the ultimate strain is scaled to preserve the
area below the plateau stress.

4.4.1. Crack Band Method

This method is based on the approach presented by Bazant (Bazant and Planas, 1998), in
which it is assumed that the work done by both the master and the scaled curves are
equal. This method also leads to an invariant fracture energy for both cases (depicted in
the following diagram) as well as work done at any intermediate value of strain
corresponding to a partially damaged state. The detail for this method is not provided
here as this method will not be used in this thesis.

"peak

Master Curve Scaled Curve

Figure 4-11 C r a c k band method. Both the master and the scaled curves are shown here.

73
Chapter 4: Damage Growth, Energy Concept and Scaling Law

4.4.2. M o d i f i e d C r a c k B a n d M e t h o d

This method is based on assuming the same fracture energy for the master and scaled
stress-strain curves and by scaling the strains after the peak point strain, e peak. This
method is shown schematically in Figure 4-12.

Aplastic Speak Aplastic Speak

Figure 4-12 Modified crack band method. Both the master and the scaled cure are shown here.

The height scaling in tension using this method has been presented by Floyd (2004) who
proposed this method because of its efficient numerical implementation. The details of
the procedure can be found in this reference and will not be presented here.

4.4.3. Scaling F a c t o r in C o m p r e s s i o n

In this section, we will present the scaling law for compression, using the procedure
presented by Floyd (2004), in tandem with the analog model.

In the analog model, the peak point strain is either the same as the initiation strain or half
of the saturation strain, whichever is larger. The difference is that in the former case there
is no permanent deformation upon unloading from the peak stress point as shown in
Figure 4-13.

The scaling factor for these two cases will be obtained separately.

74
Chapter 4: Damage Growth, Energy Concept and Scaling Law

Figure 4-13 Different unloading paths for compressive failure. The left diagram shows
unloading from the peak stress point while the peak stress and damage initiation
coincide. The right diagram shows the unloading path for an element i n which the
peak stress and damage initiation do not coincide.

4.4.3.1. Scaling Factor for an Element with Peak Stress at the Damage Initiation
Strain
Using hr/he = n as the ratio of the characteristic height to the element size, the shaded
area below the stress-strain curve from the above figure can be calculated as:

( _ 3
s, x Es, E
2
2
\ Es- -ds = Es, 12 A
3
(4-27)
s E
E- - + (s -
J
s e, )(2e, + s ) s
I o

The area below the scaled curve is:

e 2
E
y =E^-
e + - k(s -e,){2s,+ E )
s S (4-28)
2 o

Now, by equating the fracture energies for the two curves, we obtain:

Ye K = Yc K -^Y =nxy
x x
e c ->

s E E
E
~Y + -^ (e k
- i )( ,' + s ) =
2s s m (4-29)
s
1 o

75
Chapter 4: Damage Growth, Energy Concept and Scaling Law

Rearranging the above equation, the following equation for calculating the scaling factor,

L is obtained:

x ( f . + , i
2
s +" )-3f /
2
/
2

(4-30)

4.4.3.2. Scaling Factor for an Element with Peak Stress not at the Damage Initiation
Strain
Using the same notation, the area under the stress-strain curve shown as the shaded area

in Figure 4-13 is:

y,. = x ^ + EE d =
2 4( -,) S
I2 S v - , (4-31)

EE S ( /2-E )
S F | E ? ; g . s

8(e -ff/) s e -e, 12


s

And the area under the scaled curve is:

EE^ts^ 12-e ) p , E E.3

y, = + kx x-- (4-32)
S( -,) S -,
S 12

Again, if the fracture energies are equated, the following equations are derived:

y xh =y xh
e e c c -+y e =nxy c -

EE S (E /2- )
S P [ / t _ E , S =

S(E -,)
s e -e,
s 12
(4-33)
EEJ(E^ 12-Ep) E EJ*
nx \ ^ ^ + x-^-
8(e -,) s s -,
s 12

The scaling factor for this case is as follows:

In - 3 3 p

(4-34)
k= - + -(1 - )
4 2 e.

76
Chapter 4: Damage Growth, Energy Concept and Scaling Law

4.4.4. Summary of Scaling Factor Equations

As a summary, in this section the scaling factor equations which will be used to modify
the element stress-strain curves due to the element height effect are given below:

1- While peak stress is at the damage initiation strain:

X(\, 2
+ ,., +, )-3S,
2
2

k= ' s
2 ' - 2 (" )
4 3S

2- While peak stress is not at the damage initiation strain:

In-3 3 e
= r- + - 0 - " )
P
k
(4-36)
4 2 s

These equations will be used to effectively model the damage propagation process with a
finite element code (see Section 5.2. and Section 5.3.).

77
Chapter 5: M o d e l Verification and Validation

The focus of this chapter will be on verifying and validating the presented composite
compressive damage model. This model has been implemented in the commercial finite
element code, LS-DYNA, the predictions of which will be discussed here.

To verify the model, data from the literature will be used to simulate the response of a
unidirectional laminate. The successful simulation of the laminate response by using one
element, will verify the implementation of the model in LS-DYNA.

Subsequently , LS-DYNA predictions for the behaviour of composite components under


compression will be compared with experimental results to validate the model. The
following two experimental studies will be used to validate the model:

1. Notched sandwich panels under eccentric compression (Bayldon, 2003a,


2003b).

2. Open hole plates under compression (Soutis et al., 2002).

The first step to validate the implemented model is to derive appropriate material damage
parameters by studying the damage mechanism in each composite panel. These
parameters will then be used as input to the model. At the end of each simulation, the
experimental results will be compared with the analog model predictions and conclusions
will be drawn.

The successful prediction of composite component responses would validate the analog
model and allows effective use of the model to predict responses of other composite
components under tension, compression and load reversals (see McGregor, 2005, for
other simulations).

78
Chapter 5: Model Verification and Validation

5.1. Model Verification

Vogler and Kyriakides (1999) presented a study on the axial propagation of kink band in
fibre composites under compression. From their study on AS4/PEEK unidirectional
composite panels, material parameters for compression are derived and explained below.
These parameters are listed in Table 5-1.

The modulus loss due to matrix damage in compression is assumed to be 97% in this
case. This assumption is the result of kink band formation in the unidirectional laminates,
which leads to the complete failure of the panel after the saturation of matrix cracking
(see Section 4.1. for more details).

T a b l e 5-1 Compressive damage parameters for A S 4 / P E E K unidirectional panels.

Parameter Symbol Amount


Matrix damage initiation m -0.0082
Fibre damage initiation r -0.0082
Matrix damage saturation m -0.013
Compression Fibre damage saturation -0.05
'h
Damage due to the matrix saturation 'm 0.4
c
Damage due to the fibre saturation 0.6
Modulus loss due to the matrix damage \-R-F 0.97
Modulus loss due to the fibre damage 1- R C
R
0.03
Plateau stress ^ plateau 430 MPa

Other assumptions have also been made for tensile damage parameters as shown in Table
5-2 (see Floyd, 2004, for details on how to estimate tensile parameters).

It is also noteworthy that for both tensile and compressive parameters, the assumed
damage due to the fibre and matrix saturation are derived from fibre volume fraction of
60% in all panels (see Floyd, 2004, and McGregor, 2005, for more details on similar
estimations).

79
Chapter 5: Model Verification and Validation

Table 5-2 Tensile damage parameters for AS4/PEEK unidirectional panels.

Parameter Symbol Amount


Matrix damage initiation m 0.015
Fibre damage initiation 1
f
0.02
Matrix damage saturation m 0.035
Compression Fibre damage saturation 0.035
Damage due to the matrix saturation co' s 0.4
m
Damage due to the fibre saturation 0.6
Modulus loss due to the matrix damage m 0.15
Modulus loss due to the fibre damage \-R T
F
LL 0.85

The undamaged elastic material parameters were also reported for AS4/PEEK panels in
the Vogler study, as follows:

Table 5-3 M a t e r i a l parameters for AS4/PEEK unidirectional panels.

Unidirectional AS4/PEEK Amount


Longitudinal Young's modulus 128 GPa
Transverse Young's modulus 10.57 GPa
Shear modulus 5.79 GPa
Poisson's ratio 0.3

Having all the material and damage parameters, we can model the behaviour of the
laminate by using the constitutive model implemented in LS-DYNA. The response of a
shell element under uniform uniaxial and in-plane displacement in tension and
compression is shown in Figure 5-1.

From this figure, it is obvious that the compressive strength is smaller than the tensile
strength. This is due to misalignment of the fibres, which leads to formation of kink band
and failure of the laminate (Fleck, 1997).

As the shell element response in this figure follows the expected pattern for the material
response, it verifies the implementation of the model in LS-DYNA. This means that by
feeding appropriate parameters into LS-DYNA, one can reproduce the desired response.

80
Chapter 5: Model Verification and Validation

The so-called plateau stress, damage initiation strains and damage saturation strains are
also clearly shown in this figure.

I _ 1 i _ 1 5 0 0

-0.06 -0.05 -0.04 -0.03 -0.02 -0.01 0.01 0.02 0.03 0.04

Strain

Figure 5-1 Stress-Strain response of one element obtained from the implemented constitutive
model in LS-DYNA. Some of the input parameters are obtained from Vogler and
Kyriakides (1999) while the damage parameters are estimated.

81
Chapter 5: Model Verification and Validation

5.2. Simulations of Notched Sandwich Panels under Eccentric


Compression

Experiments on sandwich panels under eccentric compression with various notch and
specimen sizes were carried out at the North-Western University (Bayldon, 2003a,
2003b). The choice of sandwich panel structures and application of double-eccentric
loads were to prevent elastic buckling of specimens while concentrating the stress at the
notch tip.

During all tests, strain gauges were used to measure the strains at the edges of the
specimens, except the notched edge. Photoelastic images were also obtained from the
front face of each sandwich panel to demonstrate the strain field around the notch. Final
measured loads were reported as the specified strength for each specimen. It was
observed that a combination of kinking and delamination, starting from the notch tip,
were the main failure mechanisms in all the experiments.

The main goal of this section is to simulate these experiments using the compressive
damage model developed in this study. For this purpose, after reviewing the compressive
failure in the sandwich panels, damage parameters describing the failure process in these
panels will be derived. Then, the nominal strength of these panels, predicted by the
model implemented in LS-DYNA will be compared with the experimental results.

82
Chapter 5: Model Verification and Validation

5.2.1. Sandwich Panel Failure

The use of sandwich panels in the aeronautical and marine structures has gained
popularity during recent years. Light weight and high bending stiffness are the two main
reasons for the increasing popularity of such panels. Sandwich panels are usually made of
metallic or composite face sheets and a polymeric, metallic, or wood foam core.

Despite the fact that several studies have been performed to evaluate the load carrying
capacity of sandwich structures (Mouritz and Thomson, 1999; Mirazo and Spearing;
2001; Soutis and Spearing, 2002; Fleck and Sridhar, 2002), few predictions are offered
by current theories. One major problem in predicting the load carrying capacity is the
interaction of different failure modes. Sandwich panel failure due to delamination, face
wrinkling, buckling and fracture of skin and core are often combined. As a result the
failure response becomes a mixture of buckling, fracture and damage. Another
unresolved problem of sandwich panels is the role of shear in highly deformable cores.
This problem magnifies considering the high value of the ratio of the elastic modulus of
the face to core (Bazant, 2003).

When subjected to compressive loads, sandwich panel failure can occur by variety of
mechanisms, depending on the material parameters and panel geometry. Fleck and
Sridhar (2002) constructed collapse mechanism maps to study the dependence of failure
mechanism on these specifications.

In their study, the following four distinct failure modes for sandwich panel were
specified. They are schematically shown in Figure 5-2:

1. Face sheet microbuckling.

Face sheet microbuckling, or kink band, is perhaps the most common mode of failure for
defective or notched sandwich panels. For this.failure mechanism, the critical load
required to initiate the failure decreases when the size of the defect or crack length
increases (Mouritz and Thomson, 1999).

83
Chapter 5: Model Verification and Validation

I I I I

t t t t
Figure 5-2 Failure modes in sandwich columns (Fleck and Sridhar, 2002). From left to right,
Euler macrobuckling, core shear macrobuckling, face sheet microbuckling and face
sheet wrinkling.

2. Euler macrobuckling.

The Euler buckling of a sandwich panel is simply a function of bending stiffness and the
length of the panel. For a long panel with constant cross section, the Euler.buckling
becomes a common failure mode. On the other hand, for a sandwich panel with
significant shear deformation the role of core shear modulus will also become important
in defining the buckling load (Engesser, 1889; Haringx, 1942). Such a situation arises for
the continuum built-up columns or for highly orthotropic fibre composite columns
(Bazant, 2003).

3. Shear macrobuckling.

The core shear macrobuckling is a regular failure mode for crack-free sandwich panels
that are unstable under compression. This failure mode is a function of the shear modulus
of the core (Fleck and Sridhar, 2002).

4. Face wrinkling.

Skin wrinkling or short wavelength elastic buckling of the face sheets is also a common
failure mechanism for sandwich panels when there is no defect on the face sheet.

84
Chapter 5: Model Verification and Validation

5.2.2. Test Geometry

In all the experiments, double eccentric compressive loads are applied on notched
sandwich panels. The load is applied by loading plates through the ball bearing on the top
and bottom capture plates. As a result of this configuration, load is applied through two
eccentric pin jointed ends. In order to prevent slipping of these ball bearings, small
detents are arranged. There are also two side plates at both end of the panels to prevent
transverse movement. The test is a displacement controlled experiment in order to study
the strain softening behaviour of the composite material.

Eccentricity is introduced to concentrate the compressive stress at the notch tip and
ensure progressive failure of the material. At the same time, this loading condition will
prevent the panels from buckling or back-face failure. A schematic of the test
arrangement is shown in Figure 5-3 below.

Loading Plate

Loading Ball

Capture Plate

Side Plate

Sandwich Panel

Figure 5-3 Notched sandwich panel under double eccentric compression (Bayldon, 2003a,
2003b).

85
Chapter 5: Model Verification and Validation

There are two sets of experiments on notched sandwich panels denoted as ECC and CDC.

ECC or notch size effect test is a set of experiments on a specific sandwich panel
geometry with three different notch sizes. The face material is a woven glass-epoxy while
the core is a Divinycell foam. The specification of the ECC specimens are shown in
Table 5-4.

Table 5-4 Test geometries of notch size effect experiments (ECC). All numbers are in mm.

Core Face Notch Notch


Specimen Height Width . . . , . .
Thickness Thicknes Diameter Depth
ECC1 80 40 25.4 4.23 6.35 3.175
ECC2 80 40 25.4 4.23 12.7 3.175
ECC3 80 40 25.4 4.23 25.4 3.175

CDC or specimen size effect test is a set of experiments on four different specimen sizes.
Each successive specimen is twice as large in all dimensions as the next smaller
specimen. The smallest specimen face sheets are two-ply laminates of woven glass-
epoxy, while the other three sizes are made of 4, 8 and 16 ply laminates. The
specifications, for CDC tests are shown in Table 5-5.

Table 5-5 Test geometries of specimen size effect experiments (CDC). All numbers are in mm.

.
Specimen
c
Height
TI ,,.,
Width Core Face Notch Notch
Thickness Thicknes Diameter Depth
CDC2 20 10 6.25 0.55 3.175 1
CDC4 40 20 12.5 1.1 6.35 2
CDC8 80 40 25 2.2 12.7 4
CDC 16 160 80 50 4.4 25.4 8

Eccentricity, as shown in Figure 5-4, for ECC tests is 0.1 in both directions, while for
CDC tests is 0.077 in z direction and 0.139 in y direction. The point of load application
can be easily calculated by multiplying the eccentricity by the specimen size in that
direction. A schematic of the eccentricity and other sandwich panel sizes is shown in
Figure 5-4.

86
Chapter 5: Model Verification and Validation

Figure 5-4 Test geometry for notched sandwich panels under double eccentric compression.

87
Chapter 5: Model Verification and Validation

5.2.3. Failure Mechanisms of Woven Glass Laminates

During the past years, there has been increasing demands to use textile reinforcement.
One of the reason for this is perhaps to take advantage of their through thickness
arrangement of fibres which enhances interlaminar strength and toughness.

The other advantage of using woven material is gaining more compression after-impact
strength; the improved interlaminar shear property is considered the main reason for this
improved strength. However, as Yang et al. (2000) showed, a woven laminate may have a
lower total compressive strength compared to a [0, 90] non-crimp laminate as a result of
fibre waviness. In fact, the compressive strength decreases with the increasing fibre
misalignment or waviness (Argon, 1972; Budiansky, 1983).

Despite the fact that substantial research has been done to understand the behaviour of
woven graphite fibre composites, the amount of research on the woven glass composites
is relatively less. While more brittle, the response of woven glass composites to
compressive loads is almost linear up to failure, compared to nonlinear response of
woven carbon (Zhou and Davies, 1995; Fleck et al., 1995).

A woven laminate under compression may fail by various failure mechanisms. In fact,
geometry and material properties, laminate orientation, and laminate imperfections such
as fibre misalignment or fibre yarn waviness contribute to type of failure mechanism.
However, a notched specimen prevented from buckling, would most likely fail by a
combination of kinking and delamination (Khan et al., 2002 )

The difference between the failure of a woven laminate and failure of a [0, 90] non-crimp
laminate is that in the woven one, formation of out of plane kink band is most likely as
opposed to in-plane kinking in the other one. This difference is because of the 90 degree
fibres in woven laminates, which provide the 0 degree fibres with extra in-plane
resistance. The formation of in plane kink band is also a function of thickness and lay-up
of the laminate (see Section 2.2.6.1.).

Out of plane kinking is the main failure mechanism in CDC and ECC experiments.
Although delamination and fibre splitting can also be observed in some cases.

88
Chapter 5: Model Verification and Validation

The failure process starts with matrix cracking from the notch tip. In the case of an un-
notched specimen, the point where waviness and misalignment in the fibre is greatest, is
the starting point of matrix cracking. In tandem with matrix cracks, rotation of fibres
leads to the formation of out of plane kink band. Meanwhile, delamination propagates.
Kink band and delamination propagate in horizontal direction and in the mean time the
failure band expands, moving toward undamaged interior materials. The failure
mechanism is unstable and sudden due to the brittle response of woven glass.

The damage initiation and broadening have been confirmed by photoelastic images,
which show the starting damage height to be around 3.5 mm and then broaden up to
almost 10 mm (Bayldon, 2003a). The photoelastic images are presented in Figure 5-5,
Figure 5-6 and Figure 5-7.

Figure 5-5 Initiation of kinking with matrix cracking. Photoelastic image for the notched
sandwich panel under eccentric compression (Bayldon, 2003a).

89
Chapter 5: Model Verification and Validation

Figure 5-6 Initial damage height is equal to 3.5 mm for kinking and delamination. Photoelastic
image for the notched sandwich panel under eccentric compression (Bayldon,
2003a).

Figure 5-7 Propagation of the kink band and delamination in horizontal and vertical
directions. Height of the damaged zone is almost equal to 6 mm. Photoelastic image
for the notched sandwich panel under eccentric compression (Bayldon, 2003a).

From the photoelastic images it is also clear that out of plane kink band is the main
failure mechanism in notched sandwich panel experiments, because in-plane kink band
does not propagate horizontally (see Figure 2-3). The width of the specimen is
considerably larger than the thickness of the face sheets which provides less resistance in
the direction of out of plane kink band.

oo
Chapter 5: Model Verification and Validation

5.2.4. Material

The sandwich panel face sheets are made of woven glass-epoxy pre-preg (7781 style
glass and Bryte 250 resin). The pre-preg was manufactured by hot melt film coating and
then oven cured under vacuum consolidation (Bayldon, 2003a, 2003b). There was
noticeable porosity reported in some cases, caused by dry plies. The porosity showed no
effect on the elastic properties of the laminate but reduced its compressive strength.

Face sheets were laid up as woven [0/90] layers. Core material was also H250 Divinycell
commercial foam. For the CDC specimens, material properties for face sheets and foam
core are shown in Table 5-6 and Table 5-7.

Table 5-6 Properties of the foam core. A l l properties are in the loading direction.

Young's Shear Modulus Poisson's Density


Core
Modulus (MPa) (MPa) Ratio (Kg/m3)
H250 Divinycell 400 108 0.32 250

Table 5-7 Properties of the face sheets. Properties in both directions are the same.

Young's Shear Modulus Poisson's


Face
Modulus (MPa) (MPa) Ratio
Woven Glass-Epoxy 24300 12000 0.07

There are several damage parameters required to properly model the failure process using
the implemented model. In the following, these parameters will be discussed in detail.

From the photoelastic pictures, we can estimate the damage height. Photoelastic coating
is a thin, transparent plastic of uniform thickness, which when attached to the surface of a
test part and viewed with polarized light shows the strain field in that surface. Therefore,
we can estimate the damage height from Figure 5-5 and Figure 5-6 to be approximately
3.5 mm at the initiation of damage. The damage height at this stage is the characteristic
height parameter, which will be used in the calculation of the main fracture energy. To
calculate the broadening energy, we need to measure the height of the damage zone after

91
Chapter 5: Model Verification and Validation

complete failure. The damage height is much larger at ultimate strain due to damage band
broadening.

For this material, we assume that the peak stress is at the damage initiation strain. In other
words, we postulate a very brittle behaviour for this material according to the
experimental results (there is no nonlinearity before the peak stress). This assumption
leads us to the calculation of the damage initiation strain based on an elastic model. LS-
DYNA is then used to model different CDC specimens with a linear elastic material
model. By comparing the numerical results with the experimental results, we can estimate
the damage initiation, , to be -0.022. The estimation of this parameter will be
described in the next section.

According to this assumption and also based on the formulation of the analog model,

matrix damage saturation should not be greater than twice the initiation or -0.044 (see

Section A A3.2. ). On the other hand, the fact that woven-glass is a brittle material

requires the damage saturation strain to be around the damage initiation strain. So, we

initially choose the damage saturation strain, , to be -0.030 and we will revise this

number later in the following section.

The ultimate strain, in which all the fibres would fail, is chosen to be -0.09 based on other
experiments on the woven lay-up under compression, such as Gupta et al. (1998).

The formation of the first kink band corresponds to the matrix cracking saturation. This
allows us to calculate the damage due to matrix saturation based on the matrix volume
fraction (Floyd, 2004). Because the fibre volume fraction in CDC experiments is
approximately 40%, the matrix volume fraction is 60%. As a result, at the matrix damage
saturation state, the loss of 60% of the surface area is the total damage due to matrix
saturation. In other words, the damage due to matrix saturation is co c
s = 0.6. In the same
m

manner, damage due to fibre saturation, which is the state of ultimate failure, is

co :
c
=0.4.
v

92
Chapter 5: Model Verification and Validation

Two other parameters, required to properly describe the failure process, are modulus loss
due to matrix damage saturation and fibre damage saturation (Floyd, 2004). As described
before, after the matrix damage saturation in compression, there is no lateral support for
fibres and therefore they buckle. This means that after the formation of the kink band, the
laminate loses its entire load carrying capacity. For this reason, the modulus loss due to
matrix damage saturation is 100%. However, this parameter is chosen to be 97% due to
the presence of the bending stiffness of the fibres. Therefore, the modulus loss due to
fibre damage saturation at the ultimate strain is only 3%.

The plateau stress has the least importance among all damage parameters in this analysis.
Because of the brittle response of woven glass sandwich panels and the thin face sheets,
the value of the plateau stress is considered to be negligible. This stress was not reported
in the CDC experiments but from other experiments on the same material and lay-up, for
example Yang et al. (2000), the amount of plateau stress is derived to be 20 MPa. This
small plateau stress compared to the peak stress of approximately 500 MPa, confirms the
negligible role of the plateau stress in compressive failure of thin woven glass-epoxy
laminates. In other words, a small stress is adequate to propagate the damage into the
interior undamaged material after the saturation of matrix cracking. On the other hand,
due to the relation between the fracture energy and plateau stress (Section 4.2. ), a small
amount of energy is required for kink band broadening in the woven glass-epoxy.

This statement on the plateau stress would be altered in the case of multidirectional
laminates. For example, Gupta et al. (1998) showed that for a [+45,-45] woven laminate
under compression, the amount of plateau stress is almost equal to the peak stress. As for
the case of thick laminates, the plateau stress and the required energy to propagate the
damage band will increase. The latter assumption was also confirmed by Khan et al.
(2000), in which the plateau stress was reported to be almost one third of the peak stress.

Now that we have an estimation of all the damage parameters, we can calculate the main
fracture energy due to kinking in sandwich panels. Based on the calculation of fracture
energy for the case that damage initiation and peak stress coincide (Section 4.4.3.1. ) the
fracture energy is derived as follows:

93
Chapter 5: Model Verification and Validation

( 3 2 3 A
E
= ^IEL +f

E s El-^de = Es, / 2 +
2

v 6 2 3
y

, 2

J E 4 - + -(^.-;)(2ff +e .) / A

z o

In this equation the Young's modulus of the laminate is the modulus loss due to matrix
damage saturation (Section 4.4.3.1. ). As explained before, this is equal to 97% of the
laminate modulus. Therefore:-

24.3 x 0.97 x + " X -


(0.03 - 0.022)(2 x 0.022 + 0.03) = 8.03 x 10 KPa
3

So, knowing the damage height and the area under the stress-strain curve, the critical
strain energy release rate can be obtained as follows:

G =y y-h =8.03x10 x3.5 = 28.1 mJ/mm


f c c
3 2

This value for fracture energy changes by altering the assumption for damage saturation
strain. As discussed above, damage saturation strain should be just after damage initiation
strain. On the other hand, lowering the damage saturation strain would decrease the
critical strain energy release rate. The minimum fracture energy, equal to 20 mJ/mm , is 2

obtained by lowering the saturation strain to -0.022, a value equal to initiation strain.

The summary of all parameters is shown in Table 5-8. It should be noted that the damage
initiation in fibres and matrix are the same (see Section 3.4.2.1. for more details).

94
Chapter 5: Model Verification and Validation

Table 5-8 Damage parameters for woven glass-epoxy laminate.

Parameter Symbol Amount


Matrix damage initiation m -0.022
Fibre damage initiation -0.022
Matrix damage saturation -0.030
Fibre damage saturation -0.09
Damage due to the matrix saturation m 0.6
Damage due to the fibre saturation co c
s
0.4
Modulus loss due to the matrix damage m 0.97
Modulus loss due to the fibre damage 1 - R 0.03
C
F

Compressive toughness f G
28.1 mJ/mm2
Damage height K 3.5 mm
Plateau stress ^ plateau 20 MPa

95
Chapter 5: Model Verification and Validation

5.2.5. Finite Element Model

5.2.5.1. Experimental Results


Before modeling C D C specimens with LS-DYNA, we should obtain a representation of

strength versus size diagram for experimental results. This diagram will be used to

compare the numerical C D C size effect results against.

By applying eccentric compression on a sandwich panel, stress on one edge of the panel

will be larger than the stress on the other edges. Therefore, we can choose the maximum

applied stress under the top rigid plate as a measure of the sandwich panel strength. Due

to the linear response of the sandwich panel before complete failure, the maximum strain

instead of the maximum stress can be used. Now, based on Figure 5-4, the maximum

strain can be calculated as follows.

First, we have to calculate the effective modulus of the specimen:

IE r x t, +E xt
~ face face core core ^

E=- J
(5-1)
h
Bending stiffness in both the y and z directions can be obtained as a sum of the core and

the face bending stiffnesses, as followings.

(EI) = 2{E J )
y fa fave + (E J )C0 cnra = 2E _ x
face +Ecorc x t
^L .
(5 2)

{EI), = b x [E face (h - tl,)+ E jl


3
m c ]/12 (5-3)

Assuming that the final load on the sandwich panel is P, the maximum strain is:

P bxte.xP) hx(e x P)
(5-4)
7
= 1 L J
H
max
bhE 2{EI) y 2(EI) Z

96
Chapter 5: Model Verification and Validation

Therefore, by choosing the width of the panels as the size parameter and by calculating
maximum strains in all panels, for non-porous CDC experiments we can plot the
maximum strain as a function of characteristic size of the panel (width) on a logarithmic
scale as shown in Figure 5-8.

-4.2
2.5 3.0 3.5 4.0 4.5

-4.3 CDC Experimental data


Linear fit
-4.4

-4.6 1

J
-4.7

-4.8 1

-4.9 H

-5.0 J

Ln (Width mm)

Figure 5-8 Experimental size effect results for CDC sandwich panels under eccentric loading.

5.2.5.2. Damage Initiation Strain


As mentioned in Section 5.2.4, damage initiation strain can be derived from the elastic
response of the CDC specimens. This assumption is valid because of the linear response
of woven glass laminate. In fact, because we assumed that the damage initiation and the
peak stress coincide we can use the elastic results of CDC experiments to derive the
damage initiation strain from the notch tip strain. In this section, all of the CDC
experiments will be modeled with elastic materials and the load-displacement response
will be compared with the experimental results.

97
Chapter 5: Model Verification and Validation

First, we will obtain the stress contour at the notch tip of the CDC experiments at failure.
By extrapolating from the stresses at the element integration points, we can derive the
notch tip stress. This approximate method has to be used, because there is no closed form
analytical solution available for calculating the stress concentration factor of notched
sandwich panels under eccentric compression. The integration points are shown for
CDC8 in Figure 5-9 below.

"HBHE3

3 4
Y

Figure 5-9 Gauss integration points for CDC8. Only half of the specimen is shown here.

After finding the notch tip stress, we can simply derive the failure strain from an elastic
analysis. Stress contours of CDC specimens are shown in Figure 5-10, Figure 5-11,
Figure 5-12 and Figure 5-13.

98
Chapter 5: Model Verification and Validation

Figure 5-10 Stress contour at the notch tip for CDC2. Triangles are the Gauss integration points
of the finite element model. The zero distance point is extrapolated based on the
stress at other points.

Figure 5-11 Stress contour at the notch tip for CDC4. Triangles are the Gauss integration points
of the finite element model. The zero distance point is extrapolated based on the
stress at other points.

99
Chapter 5: Model Verification and Validation

0.0 2.0 4.0 6.0 8.0 10.0


Distance (mm)

Figure 5-12 Stress contour at the notch tip for CDC8. Triangles are the Gauss integration points
of the finite element model. The zero distance point is extrapolated based on the
stress at other points.

Figure 5-13 Stress contour at the notch tip for CDC16. Triangles are the Gauss integration
points of the finite element model. The zero distance point is extrapolated based on
the stress at other points.

100
Chapter 5: Model Verification and Validation

As a summary, the notch tip strain of the CDC experiments under the failure load is
shown in Table 5-9 below.

Table 5-9 Damage initiation strains for different CDC experiments based on their elastic
analysis.

Damage Initiation
Specimen
Strain
CDC2 -0.0240
CDC4 -0.0227
CDC8 -0.0216
CDC 16 -0.0251

According to this table, we choose the damage initiation strain equal to -0.022 which is
an average value. This strain can be altered later, if needed. However, the maximum
initiation strain should not exceed -0.025.

5.2.5.3. Scaling Factor


Considering both the element width effect and the height effect based on the works of
Floyd (2004) and McClennan (2004), we have to decrease the Young's Modulus and
increase the saturation strain to preserve the same fracture energy for different element
sizes without altering the peak stress at the notch tip (see Section 4.4. for complete
details).

Here, we assume the modified modulus due to element width effect as follows:

E =E/a
ew (5-5)

where a is the modification factor for the element width effect obtained from the
numerical stress concentration factors in Section 5.2.5.2. Fracture energy can also be
evaluated based on the damage propagation formulation, when the peak stress and
damage initiation coincide (Section 4.4.3.1. ). In this case, we assume that there is no
plastic deformation after unloading from the peak stress. This assumption is valid
because there is no damage before the peak stress. Therefore, the strain energy release
density (specific energy) is:

101
Chapter 5: Model Verification and Validation

Y =E -^-
C
C
+ ^{s -e )(2s +e :)
c c
i
c <
(5-6)

Using this equation, we can derive the area under the stress-strain curve of an element as
follows:

Y, = ^ ^ - + ^ k (
c
s - s o i l s ' ; +s )
c
s (5-7)
a 2 6a

And because the fracture energy for different element sizes should be kept constant
(Bazant and Planas, 1998), we have:

Yc K =Ye K
x
x
2
(5-8)

The factor of two is applied on the element size, because we are modeling half of the
CDC specimens. Therefore, the scaling factor is:

h E sf L

y x-S.
c aJ
2h a 2
k
= fpZ . ' (5-9)
-^--<)(2< <) +

Using this equation, the results for the CDC specimens are derived as shown below.

Table 5-10 Scaling factors for notched sandwich panel under eccentric compression.

Modified Young's Scaling


Specimen
Modulus (MPa) Factor
CDC2 20369 4.756
CDC4 19597 1.252
CDC8 22091 0.871
CDC 16 17173 1.822

102
Chapter 5: Model Verification and Validation

5.2.6. Simulations

As described in Section 5.2.4., the least known parameter among the damage parameters
is the damage saturation strain. It should be noted that modifying the saturation strain
while keeping all the other parameters constant changes the fracture energy. It is quite
obvious that by increasing the saturation strain the material becomes tougher and less size
dependent. The strength versus size curve for this material is less steep compared to a
more brittle material. Decreasing the saturation strain, on the other hand, has the effect of
decreasing the fracture energy, which then results in having a more brittle material. As a
result, size dependency would be more pronounced.

In other words, we can control the size dependency of the simulations by changing the
saturation strain. It is also noteworthy that based on the previous discussions we cannot
change the saturation strain to a value higher than twice the damage initiation strain. Also
the lower bound for the damage saturation strain is the damage initiation strain.

As mentioned before, for CDC experiments we will only model half of the specimens.
The finite element mesh for CDC8 is shown in Figure 5-14 as an example.

Figure 5-14 Finite element model for simulating the CDC8 experiment. Only half of the
specimen is modeled.

103
Chapter 5: Model Verification and Validation

CDC experiments are simulated with different fracture energies. The change of fracture
energy results from changing the damage saturation strain, as explained above. The result
of changing the fracture energy is shown in Figure 5-15.

Ln (Width mm)

F i g u r e 5-15 S i m u l a t i o n o f C D C experiments w i t h different fracture energies o r w i t h different


saturation strains.

From this figure, it is evident that by reducing the fracture energy we get results that are
closer to experiment. In fact, the best result was obtained when saturation strain was
equal to -0.023 corresponding to a fracture energy of 20.9 mJ/mm . This number indicates 2

a very brittle response for the woven glass laminate. This finding also matches our
understanding of the behaviour of this material. It is also noteworthy that for the circular
blunt notch in these panels, the change of fracture energy does not change the size
dependency as much as it does for the sharp notches. McClennan (2004) confirmed this
for sharp and blunt notches.

104
Chapter 5: Model Verification and Validation

On the other hand, the outcome of changing damage initiation strain while keeping the
fracture energy constant will be the same size dependency results for C D C experiments,
but with a lower value of strength. In other words, changing the damage initiation moves
the line of length-displacement parallel to its original position. This is confirmed in
Figure 5-16.

-4.3 1 i i l

0 2.5 3.0 3.5 4.0 4.5 5


-4.4 e, = -0.022
H
-4.5 Experiment
HI ^ v O s ^
s N

a /
-4.6

/ ^ V V ^s. /
-4.7 e, = -0.020 / H
&
e, = -0.021
J -4.; e, = -0.020
Xe, = -0.021
-4.9
Experiment
Ae, = -0.022
-5

Ln (Width mm)

Figure 5-16 Simulation of CDC experiments with different damage initiation strains while
fracture energy is equal to 20.9 mJ/mm . 2

In these simulations, three different damage initiation strains are used with the fracture
energy constant and equal to 20.9 mJ/mm . This choice is based on the previous set of
2

simulations for different fracture energies. It is noteworthy that from the above figure the
best fitted result for simulation can be obtained by using the same brittle material
parameters in the previous section, while changing the fracture energy. So, for both
simulations the best results is obtained with the saturation strain equal to -0.023 and the

105
Chapter 5: Model Verification and Validation

fracture energy equal to 20.9 mJ/mm . As a summary, the material parameters that give
the best results are shown in Table 5-11.

Table 5-11 Damage parameters for woven glass-epoxy laminate which give the best result for
C D C simulations.

Parameter Symbol Amount


Matrix damage initiation m -0.022
Fibre damage initiation -0.022
',
Matrix damage saturation m -0.023
Fibre damage saturation V -0.09
Damage due to the matrix saturation ' m
0.6
Damage due to the fibre saturation 0.4
Modulus loss due to the matrix damage \-R c
F 0.97
Modulus loss due to the fibre damage 0.03
Compressive toughness G
f 20.9 mJ/mm2
Damage height K 3.5 mm
Plateau stress ^plateau 20 MPa

106
Chapter 5: Model Verification and Validation

5.2.7. Summary

Main findings in simulating the notched sandwich panels under eccentric compression
are summarized here:

1- Woven glass-epoxy face sheets fail by a combination of out of plane kinking and
delamination, due to laminate lay-up and thickness. ,

2- Formation of kink band the resulting failure are sudden.

3- Plateau stress is negligible for a woven [0, 90] laminate.

4- Energy required to expand the damage zone is small, compared to the energy
required to form the first kink band.

5- Glass-epoxy used in this experiment is a brittle material.

6- For circular blunt notches, changing the fracture energy in the analog model does
not affect the size dependency of specimens.

7- The analog model successfully showed a size effect for CDC experiments, which
matches the experimental results.

8- Analog model is validated as a result of successful prediction of size dependency


in these specimens.
Chapter 5: Model Verification and Validation

5.3. Simulation of Open Hole Plates under Compression

5.3.1. Introduction

Soutis et al. (1993) investigated the compressive fracture properties of carbon fibre/epoxy
laminates. Compressive strengths of both unnotched and notched carbon-epoxy panels
for a wide rang of lay-ups were reported in their experiments.

They reported that the failure mechanisms in all laminates were due to the microbuckling
in the [0] plies, delamination between off-axis and [0] plies, and plastic deformation of
the off-axis plies. All laminates were symmetrically laid up and consisted of 24 plies with
equal size for both the unnotched and notched experiments.

Unnotched experiments showed that the failure strain (strain at the peak stress) was
independent of lay-up configuration. They concluded that based on this observation a
critical strain to failure or maximum strain criterion can predict the failure point with
adequate precision. For notched specimens, the failure strength, damage zone size at
failure, and notch size effect supported predictions of the cohesive crack model of Soutis
et al. (1991). To use the cohesive crack model, they measured the compressive fracture
toughness of centre-cracked specimens from which they calculated the compressive
toughness. They assumed that the response of each specimen is linear elastic before the
failure load. Then, by using the failure load they measured the compressive fracture
toughness.

Following these works, Bazant et. al. (1999) presented a size effect study on compression
strength of fibre composites failing by kink band propagation. In this work, size effect
law, proposed by Bazant (1983) for quasi-brittle materials, was verified. Bazant
generalized this law to notch-free specimens attaining the maximum load after a stable
growth of kink band and verified this law using Soutis et al. experiments (1993). In his
verification, he used the fracture energy and the length of the fracture process zone
measured solely from the maximum load.

108
Chapter 5: Model Verification and Validation

Soutis et. al. (2002) presented another investigation on the compressive strength of
carbon fibre-epoxy laminates. They used the same material as their previous work (in
1993) but with a [(45/0/90)3] lay-up. All multidirectional laminates were approximately
S

3 mm thick with different specimen and notch sizes. Failure was sudden and occurred
within the gauge section. Failure mechanism was also a combination of kinking and
delamination in all specimens.

In these experiments, they studied the effect of both notch size and specimen size on the
compressive strength of panels. Notched panel strength was predicted successfully as a
function of hole size and width, using Soutis et al. (1991) linear cohesive zone model.
The model predicted the strength by using independently measured laminate parameters
from compressive study of unnotched specimens.

In this section, the more recent experiments by Soutis et al. (2002) will be simulated
using the analog model. To estimate the damage material parameters a procedure similar
to what was used before (Section 5.2.) will be utilized. These estimations then will be
verified by using the implementation of the analog model in LS-DYNA. Fracture energy,
as the main damage parameter, will be derived from the measurement of compressive
fracture toughness in Soutis et al. experiments (1993).

Using these parameters, notch size effect and specimen size effect diagrams will also be
derived and compared with experimental results. It is expected that the results for a blunt
notch will be less dependent on the fracture energy compared to those for a sharp notch
(see McClennan, 2004).

109
Chapter 5: Model Verification and Validation

5.3.2. Test Geometry

Several specimens were cut from carbon epoxy panels, 3 mm thick, to carry out both the
un-notched and notched experiments. There were 4 different specimen sizes, each with
different circular notch sizes drilled at the center of the specimens. At least four
specimens were tested for each configuration. The circular holes were drilled using a
tungsten carbide bit to minimize fibre damage and delamination around the notch
boundary. There were also glass-fibre epoxy reinforcement tabs bonded at both ends of
the specimens for load application. A schematic of the test specimen geometry is shown
in Figure 5-17. Specimen sizes and notch diameters are also shown in Table 5-12.

. 9

4 b ii*Is + h >
-

1 a

cr

Figure 5-17 Test geometry for Soutis et al. experiments (2002). "b" is the length of the tab
section.

All experiments were quasi-static compressive tests with a displacement controlled rate
of 1 mm per minute. Several experiments were stopped during the damage growth, before
complete failure, to study the damage around the notch boundary.

There were also anti-buckling devices employed to prevent the specimens from buckling.
Anti-buckling devices contained a window around the notch tip in order to study the
damage propagation, while restraining the specimen from general bending. Frictional
effects between the anti-buckling devices and specimens were minimized by using Teflon
tapes to line the inner faces of the fixture.

110
Chapter 5: Model Verification and Validation

Table 5-12 Specimen sizes with different notch sizes for Soutis et al. experiment (2002).

Specimen Sizes Hole Diameter


a/W
(mm) (mm)
30x30 1.5 0.05
3 0.1
6 0.2
9 0.3
12 0.4
15 0.5
50x50 2.5 0.05
5 0.1
10 0.2
15 0.3
20 0.4
25 0.5
70x70 7 0.1
14 0.2
21 0.3
28 0.4
35 0.5
90x90 9 0.1
18 0.2
27 0.3
36 0.4

It is also noteworthy that the load introduction to the specimens was mainly done by end
loading using a modified ICSTM fixture as shown below.

Figure 5-18 Modified ICSTM compression test fixture.

111
Chapter 5: Model Verification and Validation

5.3.3. Material

The material used in this experiment was a multidirectional laminate autoclaved from
Toray T300 carbon fibres and embedded in Ciba-Geigy BSL 924C epoxy resin. The pre-
preg tapes were laid up by hand into a [(45/0/90)3] quasi-isotropic lay-up with an
S

approximately 3 mm thickness for all specimens. From the material parameters of the
unidirectional lay-up, we can calculate the laminate parameters in the load direction, as
given in the following table.

Table 5-13 Material parameters for the carbon-epoxy laminate in the compressive load
direction.

[(45/0/90) ]
3 s
Amount
Longitudinal Young's modulus 63 GPa
Transverse Young's modulus 63 GPa
Shear modulus 24 GPa
Poisson's ratio 0.315

We now need to estimate the damage parameters for the failure process. The key to
derive these parameters is first to study the damage mechanism in these laminates. As
mentioned, for all laminates, failure was sudden and occurred within the gauge section
and was a combination of kinking and delamination.

The damage initiation strain is the strain, in which matrix cracking, fibre rotation and
delamination start to propagate. In the experiments by Soutis et al. (2002) on notched
specimens, this strain was measured to be around 80% of the strain at the peak stress
point. The average strain at the peak stress point was reported to be almost 1.0 %.
Therefore, the damage initiation strain can be estimated to be -0.008. This number is the
damage initiation strain for both fibre and matrix. As described before, this is due to the
fact that matrix cracking and fibre rotation initiate at the same strain in the kinking
process.

The saturation of the kink band (before broadening) is related to the strain at the peak
stress. Because damage propagation is simulated as a parabolic function in the analog
model (see Section 3.4.2.) the damage saturation strain is always equal to either twice the

112
Chapter 5: Model Verification and Validation

strain at the peak stress point or the damage initiation strain, whichever is larger. From
this explanation, the damage saturation strain is estimated to be around 2.0%, which is
twice the strain at the peak stress point.

The compressive fracture toughness for a variety of lay-ups for carbon fibres T800
embedded in Ciba-Geigy BSL 924C epoxy resin was measured by Soutis et al. (1993).
The only difference between this and their other experiment (2002) is the T300 carbon
material used in the latter. Now the question arises that whether the measured
compressive fracture toughness can be used for the same lay-up and matrix material, but
with different carbon fibres. It is noted that because the Young's modulus in both cases
are almost the same, in kinking and delamination failure mechanisms where the fracture
energy mainly depends on the matrix cracking, we can use the same amount for
compressive toughness. This amount for [(45/0/90)3] lay-up was reported to be 40 MPa
S

ml/2
(Soutis et al., 1993). Based on this number and by knowing the Young's modulus for
this quasi-isotropic lay-up the fracture energy can be calculated as below:

G = A: / = 40 /62.00 = 25.4
f
2 2
kN/m = 25A mJI mm 2
(5-10)

As explained in the previous section, the two important parameters in describing damage
propagation in notched sandwich panels are damages saturations in the matrix and. the
fibre. As an estimate of damage at the saturation stage, we assume that the entire matrix
in the [0], 50 % in each of the +45 and -45, and almost nothing in the [90] layers will be
damaged and therefore the damage parameter due to the matrix damage saturation would
be 0.5. This estimation is based on the idea that the damage in the [0] layer is due to kink
band formation. In the +45 and -45 layers, a combination of delamination and kinking is
the main damage mechanism. In the [90] layer, we assume delamination is the main
failure mechanism. Accordingly, an equal amount should be assigned to the damage due
to the fibre damage saturation (Floyd, 2004) and therefore this parameter is equal to 0.5
following the same explanation.

To model the behaviour of laminates in the damage propagation process we also need to
derive the modulus loss due to damage saturation in matrix and fibres. At the matrix
damage saturation state, there is 100% damage in [0] layers due to saturation of the kink

113
Chapter 5: Model Verification and Validation

band. As described before, saturation of kink band and saturation of matrix cracking
coincide. This means that after the saturation of damage in matrix, 100% of the modulus
in [0] layer will be lost. In the +45 and -45 degree layers, a combination of delamination
and kinking decreases the modulus of layers to 50% and therefore only 50% of fibres
carry compressive loads at this stage. This is due to the fact that only 50% of the matrix
in these layers cracks and therefore 50% of fibres still have side resistance from buckling.
In [90] layers, there is almost no matrix cracking. This means that the modulus loss due
to the matrix cracking in [90] layers is zero. Now, by using ply discount method (Floyd,
2004) we can derive the remaining modulus at the saturation of matrix damage equal to
13.71 GPa. This modulus is the sum of 50% modulus in +45 and -45 degree layers and
100% modulus in 90 degree layers. The calculation of the modulus loss due to the fibre
and matrix damage saturation is given below:

c = 1 _ ~ E
n g = ] __3_1_ 0 7 8

undamaged
U Z
y
y
(5-11)
l-Rf =1-0.78 = 0.22

Based on these two parameters and by knowing the amount of fracture energy, we can
simply calculate the area under the stress-strain curve to calculate the height of damage
zone. This characteristic height is the height of damage zone at the saturation of kinking
and delamination, before broadening. Assuming that there is no plastic deformation upon
unloading from the peak stress point and by using Equation 4-32, we have the following
equation for strain energy release density:

s e
0.020
y -=7/48x L
^ = 7/48 x (62.99x0.78)
m
s -e
c
s
c
t (0.020-0.008) (5-12)

-> Xc =4.78 MPa = 4.78 xlO 3


KPa

Therefore, the characteristic height of damage is:


h = G ly
c f c = 5.32 mm (5-13)

114
Chapter 5: Model Verification and Validation

There are two more parameters required to capture the behaviour of the damage
propagation in these laminates: plateau stress and ultimate strain. Since there is no data
reported in Soutis et al. (2002) on these two parameters we have to use similar
experiments. Sivashanker (1998) has done an experimental study on the compressive
response of the same lay-up and material as in Soutis et al. (2002), except that they used
T300 carbon. We can safely ignore this difference as the damage mechanism in these
panels mainly depends on the matrix properties. In this work, the plateau stress was
reported to be around 100-133 MPa. Therefore, a plateau stress of 100 MPa will be used
here.

Sivashanker (1996) also presented a study on the response of unidirectional composites


with the same material as in Soutis et al. (2002). In their study, the ultimate strain was
reported to be around 3-4%. This should be used as a lower bound in the current study,
due to the effect of off-axis plies. For this reason, the ultimate stain is chosen to be 5.0%.
As a summary, the following parameters will be used to simulate the damage propagation
in the laminates.

T a b l e 5-14 Damage parameters for |(45/0/90) ] carbon-epoxy laminate.


3 s

Parameter Symbol Amount


Matrix damage initiation 'f.
-0.008
Fibre damage initiation -0.008
'/
Matrix damage saturation m -0.020
Fibre damage saturation -0.05
Damage due to the matrix saturation < 0.5
Damage due to the fibre saturation 0.5
Modulus loss due to the matrix damage \-Rf 0.78
m J

Modulus loss due to the fibre damage I - Rf: 0.22


Compressive toughness
G
r 25.4 mJ/mm2
Damage height K 5.32 mm
Plateau stress ~ plateau 100 MPa

115
Chapter 5: Model Verification and Validation

5.3.4. Stress-Strain Curve Scaling

As described in Section 5.2. , we need to properly scale the stress-strain curve in the
finite element model based on the element height and width.

For each element, the stress-strain response is the response of Gauss point in that
element. Based on the notch tip response from the analytical solution (following section),
we can modify the stress-strain response at the Gauss point to accurately capture the
behaviour at the notch tip (see McClennan, 2004 for more details). In this section, first
we introduce the analytical solution for stress concentration factor at the notch tip and
then by using the stress contours resulting from the finite element model, we can modify
the response of the element. This is in addition to scaling the stress-strain curve of an
element according to the element height (Floyd, 2004).

5.3.4.1. Analytical Solution


To calculate the stress concentration factor at the notch tip, we can use the formulation
for the panel under tension. A schematic of the panel is presented in the Figure 5-19
below.

* = Point of maximum stress

Figure 5-19 Open hole panel under tensile loads.

Based on the Heywood formula (Peterson, 1974), the stress concentration factor, K , is:
g

116
Chapter 5: Model Verification and Validation

2+ 1
W (5-14)
K =
If
w

5.3.4.2. Stress Contour

Using LS-DYNA, we can obtain the stress contour around the notch tip. This stress
contour will be used to calculate the ratio between the notch tip stress and stress at the
nearest Gauss point. This ratio can then be used to modify the stress-strain response of
each element in the analysis (McClennan, 2004). A schematic of the finite element mesh,
for half of the specimen, and the stress contour at different load stages are shown in
Figure 5-20 and Figure 5-21.

I i i I I I I

Symmetry line

Figure 5-20 Finite element model for the half of the 50*50 mm specimen with the hole diameter
equal to the 10 mm.

117
Chapter 5: Model Verification and Validation

In Figure 5-21, the specified points are at the Gauss points in the finite element model. It
is seen that when the far field stress is equal to 195 MPa, there.is no damage in the
laminate. The notch tip stress at this loading stage can be calculated from the above
analytical formulation. At other loading stages, damage starts to propagate and therefore
the stress level at the notch tip drops. In such cases, the notch tip stress is extrapolated
from the stress at other Gauss points.

Remote Stress=
=195 MPa

A Remote Stress==307 MPa

- H Remote Stress==236 MPa


3400

5 10 15 20 25
Distance from the notch tip (mm)

Figure 5-21 Stress contour at different load stages for the open hole specimen in Figure 5-20. As
damage progresses, the notch tip stress drops after the remote stress reaches 236
MPa.

5.3.4.3. Scaling
In order to consider the width effect, we need to lower the Young's modulus and increase
the saturation strain forward to get the exact stress-strain curve at the notch tip, and
preserve the fracture energy in the meantime (Floyd, 2004; McClennan, 2004). This has
been demonstrated in Figure 5-22.

118
Chapter 5: Model Verification and Validation

For the height effect, on the other hand, we have to modify the stress-strain response of
the element after the peak stress according to the scaling law (Section 4.4.). This is also
shown in Figure 5-23 below.

Master Curve

Modified Curve

Figure 5-22 Modifying the Stress-Strain curve based on the element width.

Master Curve
Modified Curve

Figure 5-23 Modifying the Stress-Strain curve based on the element height. The scaling is shown
for an element height less than h .
c

Based on these two figures, the modified Young's modulus and scaling factor for each
model are derived. These parameters are shown for the simulation of 50x50 mm
specimen in Table 5-15 below.

119
Chapter 5: Model Verification and Validation

Table 5-15 Modified parameters for the simulation of 50x50 mm specimens.

Notch Diameter Modified Young's k-scaling


(mm) Kg Modulus (GPa) Factor
2.5 3.0 49.94 4.4
5 3.0 50.0 4.4
10 3.1 56.75 3.8
15 3.3 57.53 3.7
20 ' 3.7 56.95 3.8
25 4.3 65.57 3.8

120
Chapter 5: Model Verification and Validation

5.3.5. Simulation Results

The results of simulations for the specimen strengths are shown in Figure 5-24.

In this figure, there are also two LEFM bounds. One is the "hole insensitive" bound,
which is obtained by multiplying the peak stress by the minimum net area for the notched
specimen. The other one is the hole sensitive bound, which is obtained from the
calculation of stress concentration factor. By calculating the stress concentration factor
from analytical solution we can estimate the far field stress, while at the notch tip stress
reaches the critical stress. In this figure, the LS-DYNA prediction lies between the
minimum and maximum of the experimental results.

0_7 ~ (

Figure 5-24 Analytical and numerical strength predictions of 50*50 mm open hole panels under
uniaxial compressive loads.

121
Chapter 5: Model Verification and Validation

5.3.5.1. Notch Size Effect


For the 50*50 mm specimen, the notch size effect is presented in this section. The
simulation results for different fracture energy values are shown in Figure 5-25.

0.2 -I 1 1 1 1 1 1
0 0.1 0.2 0.3 0.4 0.5 0.6
Hole Diameter/Specimen Width

Figure 5-25 Notch size effect results for different fracture energy values.

It is clear that for different fracture energy values, the resulting curves are almost parallel
to each other. Also as expected, by increasing the fracture energy, the predicted specimen
strength increases. For Gf = 25.4 mJ/mm , the predictions tie between experimental
2

bounds.

We can also obtain the results for different saturation strains, as shown in Figure 5-26.
The results are presented for only two saturation strains. It is also obvious that the best
result is obtained when saturation strain is equal to -0.020. This confirms our first
estimation of the saturation strain. It can also be observed that by increasing the

122
Chapter 5: Model Verification and Validation

saturation strain, the resulting curve moves up but remains almost parallel to the original
curve.

150 H 1 1 1 1 r 1
0 5 10 15 20 25 30
Hole Diameter (mm)

Figure 5-26 LS-DYNA strength predictions for different saturation strains in 50*50 mm panels.
Fracture energy is equal to 25.4 mJ/mm .
2

123
Chapter 5: Model Verification and Validation

5.3.5.2. Specimen Size Effect


In this section, the effect of specimen size on its compressive strength is studied.
Different specimen sizes are examined for a constant hole diameter to specimen width
ratio of 0.2. Results for two different fracture energy levels are shown in Figure 5-27. For
fracture energy equal to 25.4 mJ/mm , the predictions fit better with the experimental
2

results. This number is the estimate that is obtained from the study on the failure
mechanism and measurement of compressive toughness (Section 5.3.3.).

450
-tt-Gj-25.4 mJ/mm 2

Experimental Results Min. Experiment


400
-X Max. Experiment
X Gj-= 30 mJ/mm 2

350
CS

& 300
C
a
250
Gf = 25.4 mJ/mm 2
Gj- = 30 mJ/mm 2

200

150 i

20 30 40 50 60 70 80 90 100
Specimen Width (mm)

Figure 5-27 Specimen size effect for different fracture energy values. (Hole diameter/ Specimen
width = 0.2).

We can obtain the results for different damage saturation strains. This has been depicted
in Figure 5-28 for two different saturation stains, but with a constant fracture energy of
25.4 mJ/mm . Also it is shown that for saturation strain equal to -0.020 the predictions are
2

closer to the experimental results.

124
Chapter 5: Model Verification and Validation

It is noteworthy that for both the specimen size effect and the notch size effect the best
results are obtained by using damage parameters derived in Section 5.3.3..

450
Min. Experiment
Experimental Results -Hr- Max. Experiment
400 e, = -0.020
- X - E = -0.022
S

350

rfi 300
c

250
e = -0.022
s

Es = -0.020
200

150
20 30 40 50 60 70 80 90 100
Specimen Width (mm)

Figure 5-28 Specimen size effect for different damage saturation strains. (Hole diameter/
Specimen width = 0.2).

125
Chapter 5: Model Verification and Validation

5.3.6. Summary of Observations

In summary, the main findings in simulation of the open hole panels under compression
are presented in point form here:

1- The experimentally observed specimen size effect and notch size effect in
[(45/0/90)3] carbon-epoxy panels were simulated effectively using the
constitutive model.

2- The results obtained from the LS-DYNA simulations showed that the estimate of
damage parameters obtained from the study of failure mechanism can adequately
predict the experimental results.

3- The model was validated by successfully simulating the damage process in angle-
ply laminates that fail by kinking, delamination and off-axis matrix cracking.

4- It was confirmed that the damage-based constitutive model is capable of


predicting both specimen size effect and notch size effect in all similar panels.

126
Chapter 6: Conclusions and Future W o r k

6.1. Conclusions

Through reviewing the literature, the main differences between the tensile and
compressive failure mechanisms were distinguished. One important difference between
the two modes of loading is the wide range of failure mechanisms in compression which
depends on geometry and material parameters. Even for the same geometry and material,
the laminate lay up plays an important role in the compressive damage propagation
process.

Another main difference between the tensile and compressive failure mechanisms is the
band broadening state in compression. After saturation of the damage zone in
compression, under a constant remote stress, or plateau stress, the damage band
propagates into the interior of the undamaged material. This phenomenon suggests a 2D
damage propagation in compression, consisting of transient and steady-state damage
growth.

By taking all these differences into account, an analog model was constructed based on
the experimental studies by Moran et al. (1995). This mechanical analog model is made
of basic elements, such as springs and fuses, to simulate the behaviour of a composite
material during the damage process caused by uniaxial compression, tension and load
reversals.

By grouping the basic elements, two distinct systems (boxes of elements) were
constructed. Each box represents an aspect of the physical behaviour in the failure
process. The laminate box represents the behaviour of the intact laminate in the damage
propagation process, while the rubble box represents the elastic response of the damaged
material.

127
Chapter 6: Conclusions and Future Work

The more challenging issue in constructing the model was to develop an energy concept
in the compressive failure process. The fracture energy is divided into two parts: the main
fracture energy and the broadening fracture energy. The main fracture energy was used to
modify the stress-strain response of the element following the crack band theory (Bazant
and Planas, 1998; Floyd, 2004).

The successful implementation of the model in the finite element code, LS-DYNA,
allowed using the model to predict the response of composite components undergoing
compressive failure process. Comparison of the numerical results and experimental
results from two different studies (Soutis et al., 1993, 2002; Bayldon, 2003a, 2003b), was
used to validate the model. These comparisons served to validate not only the predictions
of the compressive strength but also the influence of the specimen size on these strength
values.

As a result of this work, the implemented constitutive model is proposed as an effective


tool to predict the response of composite structures under compression, tension and load
reversals. The successful prediction of the size effect in the validation of the model show
its usefulness for predicting the response of large composite components.

The conclusions can be summarized as follows:

1- Extensive data on the failure of unidirectional composite laminates under


compressive loads is available in the literature.

2- Few studies have been done with respect to the compressive failure of
multidirectional laminates.

3- In thermoset matrix composites, the formation of kink band under compressive


loads is due to matrix cracking, which leads to softening and rotation of the fibres.

4- In thermoplastic matrix composites, the formation of kink band under


compressive loads is due to the yielding of the matrix, which leads to rotation of
the fibres.

128
Chapter 6: Conclusions and Future Work

5- There are two fracture energies involved in compression failure: main fracture
energy and broadening fracture energy. The main fracture energy is associated
with the formation of the first kink band. The band broadening fracture energy is
the energy required to form additional kink bands or to expand damage into the
interior undamaged material.

6- The model is capable of simulating the damage process in multidirectional


laminates that fail by kinking, delamination and off-axis matrix cracking.

7- The model is capable of predicting both the specimen size effect and notch size
effect in multidirectional composite panels. In other words, the model can be used
to predict the response of large composite structures undergoing damage
propagation in compression or tension.

129
Chapter 6: Conclusions and Future Work

6.2. Future Work

The nature of the load application on composite components requires the study of the
damage and failure process under a 3D state of stress. We cannot apply the ID analog
model to predict the response of a specimen under tri-axial loading condition due to the
effects such as Poisson's ratio or interaction of tension and compression. This reveals the
need to construct a mutiaxial stress-strain model to simulate the 3D response of
composite components under different loading conditions.

A real structure undergoes a variety of stresses. Upon impact, fatigue, off-axis loading
and other loading applications the role of shear and bending stresses become important in
determining the response of composite components. At present, there are no constitutive
models capable of accounting for all different loading conditions, including a
combination of compression, tension, bending and shear.

To achieve this goal, a comprehensive literature review is required to identify failure


mechanisms under shear and bending loads. By studying the differences and similarities
between these failure mechanisms and compressive and tensile failure mechanisms a 3D
damage model could be constructed.

It is also desirable to implement this enhanced model in a finite element code in order to
provide the ability of predicting the response of composite structures in the failure
process.

The future work can be summarized as:

1- Extending the analog model to 3D.

2- Constructing a model to account for compressive, tensile, shear and bending


failure of multidirectional composites.

3- Implementing the model in a finite element code and validating the model.

130
References

Argon, A., (1972). "Fracture of composites", Treatise on Materials Science and


Technology, Vol. 1, pp. 79-114..

Bayldon, J., (2003a). "Effect of defects in GA composite: size effect testing", North
Western University, Presentation.

Bayldon, J., (2003b). "Effect of defects on GA aircrafts", Center for Intelligent


Processing of Composite Materials, Presentation.

Bazant, Z.P., (1983). "Fracture in concrete and reinforced concrete", IUTAM Prager
Symposium on Mechanics of Geomaterilas: Rocks, Concrete, Soil., pp. 281-316.

Bazant, Z.P., (2003). "Shear buckling of sandwich, fibre composite and lattice columns,
bearings, and helical springs: Paradox resolved", Journal of Applied Mechanics,
Transactions ASME, Vol. 70, No. 1, pp. 75-83.

Bazant, Z.P., Kim, JJ.H., Daniel, I.M., Becq-Giraudon, E. and Zi, G., (1999). "Size effect
on compression strength of fibre composites failing by kink band propagation",
International Journal of Fracture, Vol. 95, pp. 103-141.

Bazant, Z.P. and Planas, J., (1998)."Fracture and Size Effect". CRC Press LLC.

Berbinau, P., Soutis, C , Guz, I.A., (1999). "Compressive failure of 0 degree


unidirectional carbon-fibre-reinforced plastic(CFRP) laminates by fibre
microbuckling", Composites Science and Technology, Vol. 59, pp. 1451-1455.

Budiansky, B., (1983). "Micromechanics", Computers and Structures, Vol. 16, pp. 3-12.

Budiansky, B., Fleck, N.A., (1993). "Compressive failure offibrecomposites", Journal of


Mechanics and Physics of Solids, Vol. 41, No. 1, pp. 183-211.

131
References

Budiansky, B., Fleck, N.A., Amazigo, J.C.O, (1998). "On kink-band propagation in fibre
composites", Journal of the Mechanics and Physics of Solids, Vol.46, No.9, pp.
1637-1653.

Dow, J.F., Gruntfest, I.J, (1960). "Determination of most needed, potentially possible
improvements in material for ballistic and space vehicles", General Electric Co.
Report, No. TIS R60SD389.

Dugdale, D.S., (1960). "Yielding of steel sheets containing slits", Journal of Mechanics
and Physics of Solids, pp. 100-104.

Engesser, F., (1889). "Die Knickfestigkeit gerader Sf'abe.", Zentralblatt des


Bauverwaltung, Vol. 11, pp. 483.-486.

Fleck, N.A., (1997). "Compressive failure of fibre composites", Advances in Applied


Mechanics, Vol. 33, pp. 43-113.

Fleck, N.A., Deng, L., Budiansky, B., (1995). "Prediction of kink width in compressed
fibre composites", Journal of Applied Mechanics, Transactions ASME, Vol. 62,
No. 2, pp. 329-337.

Fleck, N.A., Jelf, P.M., Curtis, P.T., (1995)."Compressive failure of laminated and
woven composites", Journal of Composites Technology & Research, Vol. 17, No.
3, pp. 212-220.

Fleck, N.A., Sridhar, I., (2002). "End compression of sandwich columns", Composites -
Part A: Applied Science and Manufacturing, Vol. 33, No. 3, pp. 353-359.

Floyd, A.M., (2004). "An engineering approach to the simulation of gross damage
development in composites laminates", Ph.D. Thesis, Department of Civil
Engineering, The University of British Columbia.

Fried, N., (1963), "The compressive strength of parallel filament reinforced plastics- the
role of the resin", Proceedings of 18 Annual Meeting of the Reinforced Plastics
th

Division, Society of Plastic Industry, Section 9-A, pp. 1-10.

132
References

Gibson, L.J., Ashby, M.F., (1988). "Cellular solids: structure and properties", Pergamon,
New York.

Gu, H., Chattopadhyay, A., (1995). "Delamination buckling and postbuckling of


composite cylindrical shells", Collection of Technical Papers
AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics & Materials
Conference, No. 5, pp. 3095-3105.

Gupta, V. , Anand, K., Grape, J., (1998)." Failure of woven carbon-polyimide laminates
under off-axis compression loading", Acta Materialia, Vol. 46, No. 2, pp. 711-
718.

Guynn, E.G., Bradley, W.L., (1989). "Detailed investigation of the micromechanisms of


compressive failure in open hole composite laminates", Journal of Composite
Materials, Vol. 23, No. 5, pp. 479-504.

Hahn, H.T., Williams, J.G., (1986). "Compression failure mechanisms in unidirectional


composites", ASTM Special Technical Publication, pp. 115-139.

Haringx, J.A., (1942). "On the buckling and lateral rigidity of helical spring", Proc,
Konink. Ned. Akad.Wetenschap., Vol. 45, pp. 533.

Jackson, W.C., Ratcliffe, J.G., (2004). "Measurement of fracture energy for kink-band
growth in sandwich specimens", Proceedings of Composite Testing and Model
Identification, Paper No. 24.

Jelf, P.M., Fleck, N.A., (1992). "Compression failure mechanisms in unidirectional


composites", Journal of Composite Materials, Vol. 26, No. 18, pp. 2706-2726.

Johnson, A.M., Ellen, S.D. (1974). "A theory of concentric kink. And sinusoidal folding
and of monoclinal flexuring of compressible, elastic multilayers. I introduction",
Tectonophysics, Vol. 21, pp. 301-339.

133
References

Johnson, A.M., Ellen, S.D. (1975a). "A theory of concentric kink. And sinusoidal folding
and of monoclinal flexuring of compressible, elastic multilayers. II initial stress
and nonlinear equations of equilibrium", Tectonophysics, Vol. 25, pp.261-280.

Johnson, A.M., Ellen, S.D. (1975a). "A theory of concentric kink. And sinusoidal folding
and of monoclinal flexuring of compressible, elastic multilayers. Ill transition
from sinusoidal to concentric-like chevron folds", Tectonophysics, Vol. 27, pp.l-
38.

Johnson, A.M., Ellen, S.D. (1976). "A theory of concentric kink. And sinusoidal folding
and of monoclinal flexuring of compressible, elastic multilayers. IV development
of sinusoidal and kink folds in multilayers confined by rigid boundaries",
Tectonophysics, Vol. 30, pp. 197-239.

Kardomateas, G.A., (1993). "Initial post-buckling and growth behaviour of internal


delaminations in composite plates", Journal of Applied Mechanics, Transactions
ASME, Vol. 60, No. 4, Dec, pp. 903-910.

Kardomateas, G.A., Schmueser, D.W., (1988). "Buckling and postbuckling of


delaminated composites under compressive loads including transverse shear
effects", AIAA Journal, Vol. 26, No. 3, pp. 337-343.

Khan, A.S., Colak, O.U., Centala, P., (2002). "Compressive failure strengths and modes
of woven S2-glass reinforced polyester due to quasi-static and dynamic loading",
International Journal of Plasticity, Vol. 18, No. 10, pp. 1337-1357.

Kaute, D., Ashby, M.F., Fleck N.A., (1996). "Compressive failure in ceramic matrix
composites". Advances in Applied Mechanics, Vol. .33, pp. 43-117.

Kim, J.K., Sham, M.L., (2000)." Impact and delamination failure of woven-fabric
composites", Composites Science and Technology, Vol. 60, No. 5, pp. 745-761.

Liu, X.H., Moran, P.M., Shih, C.F., (1996). "Mechanics of compressive kinking in
unidirectional fibre reinforced ductile matrix composites", Composites Part B:
Engineering, Vol. 27, No. 6, pp. 553-560.

134
References

McClennan, S.A., (2004). "Crack growth and damage modeling of fibre reinforced
polymer composites", MASc. Thesis, Department of Materials Engineering, The
University of British Columbia.

McGregor, C.J., (2005). MASc. Thesis, Department of Civil Engineering, The University
of British Columbia, in preparation.

Mirazo, J.M., Spearing, S.M., (2001). "Damage modeling of notched graphite/epoxy


sandwich panels in compression", Applied Composite Materials, Vol. 8, pp. 191-
216.

Moran, P.M., Liu, X.H., Shih, C.F., (1995). "Kink band formation and band broadening
in fibre composites under compressive loading", Acta Metallurgica et Materialia,
Vol. 43, No. 8, pp. 2943-2958.

Moran, P.M., Shih, C.F., (1998). "Kink band propagation and broadening in ductile
matrix fibre composites: Experiments and analysis", International Journal of
Solids and Structures, Vol. 35, No. 15, pp. 1709-1722.

Mouritz, A.P., Thomson, R.S., (1999). "Compression, flexure and shear properties of a
sandwich composite containing defects", Composite Structures, Vol. 44, No. 4,
pp.263-278.

Niu, K., Talreja, R., (2000). "Modeling of compressive failure in fibre reinforced
composites", International Journal of Solids and Structures, Vol. 37, No. 17, pp.
2405-2428.

Palmer, A., Rice, J.R. (1973). "The growth of slip surfaces in the progressive failure of
over-consolidated clay", Proceedings of the Royal Society of London, A332, pp.
527-548.

Peterson, R.E., (1974)."Stress Concentration Factors", John Wiley & Sons.

135
References

Piggott, M R . , Harris, B., (1980). "Compression strength of carbon, glass and kevlar-49
fibre reinforced polyester resins", Journal of Materials Science, Vol. 15, No. 10,
pp. 2523-2538.

Rosen, V.W., (1965). "Mechanics of composite strengthening", Fibre Composite


Materials. American Society of Metals, Metals Park, Ohio, pp. 37-75.

Shu, J.Y., Fleck, N.A., (1997). "Microbuckle initiation in fibre composites under
multiaxial loading", Proceedings of the Royal Society of London, Series A:
Mathematical, Physical and Engineering Sciences, Vol. 453, No. 1965, pp. 2063-
2083.

Sivashanker, S., (1998). "Damage growth in carbon fibre-PEEK unidirectional


composites under compression", Proceedings of the 1997 Symposium on
Integrated Experimental-Computational Modeling of Advanced Materials,
McNu'97, Vol. A249, No. 1 -2, pp. 259-276.

Sivashanker, S., (2001). "Damage propagation in multidirectional composites subjected


to compressive loading", Metallurgical and Materials Transactions A: Physical
Metallurgy and Materials Science Physical Metallurgy and Materials Science,
Vol. 32, No. l,pp. 171-182.

Sivashanker, S., Bag, A., (2001). "Kink-band propagation in a multidirectional carbon


fibre-polymer composite", Metallurgical and Materials Transactions A: Physical
Metallurgy and Materials Science Physical, Vol. 32, No. 12, pp. 3157-3160.

Sivashanker, S., Fleck, N.A., Sutcliffe, M.P.F., (1996). "Microbuckle propagation in a


unidirectional carbon fibre-epoxy matrix composite", Acta Materialia, Vol. 44,
No. 7, pp. 2581-2590.

Soutis, C , Curtis, P.T., Fleck, N.A., (1993). "Compressive failure of notched carbon-
fibre composites", Proceedings of the Royal Society of London, Series A:
Mathematical, Physical and Engineering Sciences, Vol. 440, No. 1909, pp. 241-
256.

136
References

Soutis, C, Fleck, N.A., Smith, P.A., (1991). "Failure prediction technique for
compression loaded carbon fibre-epoxy laminate with open holes", Journal of
Composite Materials, Vol. 25, No. 11, pp. 1476-1498.

Soutis, C, Lee, J., Kong, C, (2002). "Size effect on compressive strength of T300/924C
carbon fibre-epoxy laminates", Plastics, Rubber and Composites, Vol. 31, No. 8,
pp. 364-370.

Soutis, C, Spearing, S.M., (2002). "Compressive response of notched, woven fabric, face
sheet honeycomb sandwich panels", Plastics, Rubber and Composites, Vol. 31,
No. 9, pp. 392-397.

Sutcliffe, M.P.F., Fleck, N.A., (1994). "Microbuckle propagation in carbon fibre-epoxy


composites", Acta Metallurgica et Materialia, Vol. 42, No. 7, pp. 2219-2231

Sutcliffe, M.P.F., Fleck, N.A.,' (1997). "Microbuckle propagation in fibre composites",


Acta Materialia, Vol. 45, No. 3, Mar, 1997, pp. 921-932.

Sutcliffe, M.P.F., Fleck, N.A., Xin, X.J., (1996). "Prediction of compressive toughness
for fibre composites", Proceedings of the Royal Society of London, Series A:
Mathematical, Physical and Engineering Sciences, Vol. 452, No. 1954, pp. 2443-
2465.

Vogler, T.J., Kyriakides, S., (1997). "Initiation and axial propagation of kink bands in
fibre composites", Acta Materialia, Vol. 45, No. 6, pp. 2443-2454.

Vogler, T.J., Kyriakides, S., (1999). "On the axial propagation of kink bands in fibre
composites: Part I experiments", International Journal of Solids and Structures,
Vol. 36, No. 4, pp. 557-574.

Vogler, T.J., Kyriakides, S., (2001). "On the initiation and growth of kink bands in fibre
composites: Part I. Experiments", International Journal of Solids and Structures,
Vol. 38, No. 15, pp. 2639-2651.

137
References

Williams, K.V., (1998). "A physically-based continuum damage mechanics model for
numerical prediction of damage growth", Ph.D. Thesis, Department of Materials
Engineering, The University of British Columbia.

Williams, K.V., Vaziri, R., Poursartip, A., (2003). "A physically-based continuum
damage mechanics model for thin laminated composite structures", International
Journal of Solids and Structures, Vol. 40, pp. 2267-2300.

Yang, B., Kozey, V., Adanur, S., Kumar, S., (2000). "Bending, compression, and shear
behaviour of woven glass fibre-epoxy composites", Composites Part B:
Engineering, Vol. 31, No. 8, pp. 715-721.

Yurgatis, S.W., (1987). "Measurement of small angle misalignments in continuous fibre


composites", Composites Science and Technology, Vol. 30, pp. 279-293.

Zhou, G., Davies, G.A.O., (1995). "Characterization of thick glass woven


roving/polyester laminates: 1. Tension, compression and shear", Composites, Vol.
26, No. 8, pp. 579-586.

138
Appendix A : Taylor Series for Damage Height Function

Using Taylor series, we will simplify the following function to a polynomial:

f(x) = (x + dy)" (A-l)

where, dy is a very small number and n is a constant value. Taylor series is given below:

f ( \
n

f(x) = f(a) + f'(a)x(x-a) + ^-^-(x-a) 2


+ ... (A-2)

Now, by expanding the function about a = x - dy, we have:


/ (x) = (x - dy + dy)" + n(x -dy + dy)"~ x(x-x
]
+ dy) +

nx(n-\)(x-dy + dy) ~ x
n 2 { x
~ X
+ d y
^ +... ^

Now considering the fact that dy is very small, we can derive:

fix) = (x + dy) n
(x) + n(x)"~ x dy
n ]
(A-4)

139

Das könnte Ihnen auch gefallen