+ All Categories
Home > Documents > COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami,...

COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami,...

Date post: 17-Jul-2020
Category:
Upload: others
View: 0 times
Download: 0 times
Share this document with a friend
159
COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN VISCOELASTIC-VISCOPLASTIC COMPOSITE MATERIALS by Thomas John Berton A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Department of Materials Science and Engineering University of Toronto © Copyright by Thomas John Berton (2019)
Transcript
Page 1: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE

IN VISCOELASTIC-VISCOPLASTIC COMPOSITE MATERIALS

by

Thomas John Berton

A thesis submitted in conformity with the requirements

for the degree of

Doctor of Philosophy

Department of Materials Science and Engineering

University of Toronto

© Copyright by Thomas John Berton (2019)

Page 2: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

ii

Computational modelling of progressive damage in viscoelastic-viscoplastic composite materials

Thomas John Berton

Doctor of Philosophy

Department of Materials Science and Engineering

University of Toronto

2019

ABSTRACT

Current governmental requirements for greater fuel efficiency are pushing automotive

manufacturers to find innovative ways of reducing car weight, while still maintaining mechanical

performance to ensure passenger safety. Composite materials would allow for significant weight

reduction, while still maintaining a safe environment for the passengers. However, composite

materials are known to exhibit significant time-dependent behaviour, as well as progressive damage

which affect structural performance. Bio-composites in particular are known to undergo complex

damage evolution during their lifetime. In this thesis, a multi-scale computational approach has been

adopted to take these effects into account.

A new modelling approach based on Synergistic Damage Mechanics to understand damage

evolution in composite materials undergoing time-dependent deformation is developed. The model is

applied to matrix micro-cracking in laminates. Computational micro-damage mechanics is combined

with a continuum level description of stiffness degradation to predict the evolution of micro-cracking

while requiring minimal experimental calibration. The time-dependent behaviour is modelled using

Schapery’s theory of viscoelasticity and viscoplasticity, and the evolution of matrix micro-cracking

during creep is predicted. The predictions of the model for different stacking sequences, ply

thicknesses and working temperatures show that viscoelasticity and viscoplasticity have a significant

effect on the long-term response of laminates undergoing matrix micro-cracking.

Page 3: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

iii

The multi-scale modelling approach is extended to short fiber bio-composites by constructing a

micro-damage model of fiber-matrix debonding using a Cohesive Zone Model. The effect of strain

rate on fiber-matrix debonding is analyzed by implementing a non-linear viscoelastic model for the

matrix, and it is shown that viscoelastic behaviour leads to a competition between void nucleation and

debonding. The effects of fiber stiffness, dimensions and interfacial shear strength are evaluated.

Following the micro-damage analyses, the thesis focuses on damage evolution at the structural

scale by implementing the damage model in Finite Element Analysis software to analyze the dynamic

response of a car bumper. The rate-dependent evolution of damage is predicted for different stacking

sequences and material systems.

Page 4: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

iv

To my parents, and brothers

Page 5: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

v

ACKNOWELDGEMENTS

Many people have helped me in different ways through this challenging research journey. First and

foremost, I wish to thank my parents and my brothers for their unwavering emotional support during the

past five years.

I wish to thank my supervisor, Prof. Chandra Veer Singh, for his insightful guidance throughout my

entire research work. His vision for the potential of computational modelling of damage in composite

materials has allowed me to learn and make contributions to this field that I never thought would be

possible at the start of my PhD. I also would like to acknowledge my other PhD committee members,

Prof. Hani Naguib and Prof. Mohini Sain, who provided me with valuable support and advice throughout

my studies. Their encouraging attitude and willingness to review my work in detail on a regular basis have

allowed me to finish my thesis in the best conditions possible. I also am grateful for the support of my

external reviewer, Prof. Xinran Xiao, as well as Prof. Ben Hatton and Prof. Yu Zou who provided excellent

feedback to improve my final thesis.

I wish to thank all the members of the Computational Materials Engineering lab at the University of

Toronto, as well as the members of the Centre for Bio-composites and Bio-materials Processing. I have

felt at home in this research environment, and much of my ability to conduct research for the past five

years was made possible by the support of these two research groups.

I received technical help from several people. In particular, I want to thank Prof. John Montesano,

Yangjie, Dr. Sandip Haldar and Farzin for their technical support. I have learned a tremendous amount

about finite element modelling and solid mechanics from them, and they made my PhD research possible.

Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit,

Shwetank, Kulbir, Nav, Sean, Ashok, Gurjot, Meysam, and Hao. We supported each other through the

difficulties all of us had to face at some point, and cheered each other on through the many successes and

victories. They made Toronto my home.

Page 6: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

vi

TABLE OF CONTENTS

ABSTRACT ......................................................................................................................................ii

Dedication ..................................................................................................................................... iv

ACKNOWELDGEMENTS ......................................................................................................................... v

Chapter 1. Introduction .................................................................................................................. 1

1.1 Lightweighting in the transportation sector ................................................................................. 1

1.2 Modelling the mechanical performance of composites ............................................................... 5

1.3 Thesis objectives .......................................................................................................................... 6

Chapter 2. Literature review .......................................................................................................... 9

2.1 Composite materials .................................................................................................................... 9

2.2 Damage mechanisms in composites .......................................................................................... 10

2.2.1 Matrix micro-cracking in laminates ....................................................................................... 11

2.2.2 Damage mechanisms in short fiber composites ..................................................................... 13

2.2.3 Interaction between viscoelasticity, viscoplasticity and damage evolution in composites .... 14

2.3 Models to predict matrix micro-cracking in laminates .............................................................. 16

2.4 Models to predict the combined effects of damage, viscoelasticity and viscoplasticity in

composites ........................................................................................................................................... 20

2.5 Structural analysis of composite structures under low-velocity impact..................................... 22

2.6 Micro-mechanical models to predict damage evolution in short fiber composites.................... 24

2.7 Gaps in the literature .................................................................................................................. 26

2.8 Thesis methodology ................................................................................................................... 26

Chapter 3. Methodology .............................................................................................................. 28

3.1 Prediction of damage in viscoelastic-viscoplastic laminates under multi-axial creep loading .. 28

3.1.1 Synergistic Damage Mechanics (SDM) model ...................................................................... 29

3.1.2 FE micro-damage model ........................................................................................................ 30

3.1.3 Energy-based crack multiplication model .............................................................................. 31

3.1.4 Viscoelasticity and viscoplasticity models ............................................................................. 33

3.1.5 Combined viscoelastic-viscoplastic-damage model ............................................................... 34

3.1.6 Numerical implementation ..................................................................................................... 35

3.1.7 Detailed implementation of the model into MATLAB .......................................................... 36

3.1.8 Development of an RVE to evaluate the effect of damage on laminate stiffness .................. 44

3.1.9 Assumptions of the model and range of applicability ............................................................ 47

3.2 Micro-damage model for a short fibre composite ...................................................................... 48

3.2.1 Geometry ................................................................................................................................ 48

3.2.2 Material models ...................................................................................................................... 48

3.2.3 Validation of the implementation ........................................................................................... 52

Page 7: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

vii

3.3 Structural analysis of a composite car bumper .......................................................................... 54

Chapter 4. Development of an SDM-based viscoelastic creep damage model ............................ 57

4.1 Multidirectional laminates undergoing cracking in multiple plies. ........................................... 57

4.2 Cross-ply laminates undergoing cracking in transverse plies. ................................................... 61

4.3 Predictions for [±60/902]s and [±75/902]s CFRP laminates. ....................................................... 64

Chapter 5. Time-dependent bi-axial stiffness degradation envelopes for structural composites . 70

5.1 Model validation ........................................................................................................................ 70

5.1.1 Validation of the viscoelastic-viscoplastic implementation ..................................................... 71

5.1.2 Validation of the crack multiplication model ........................................................................... 73

5.1.3 Validation of the SDM model .................................................................................................. 73

5.2 Model predictions ...................................................................................................................... 74

5.2.1 GFRP ........................................................................................................................................ 74

5.2.2 CFRP ........................................................................................................................................ 79

5.2.3 Effect of temperature on the stiffness degradation envelopes .................................................. 81

5.3 Discussion ................................................................................................................................... 83

Chapter 6. Strain rate-dependent damage model for a short fiber composite .............................. 87

6.1 Model validation ......................................................................................................................... 87

6.2 Effect of strain rate on damage evolution ................................................................................... 87

Chapter 7. Structural analysis using a rate-dependent SDM model ........................................... 102

7.1 Implementation validation ........................................................................................................ 102

7.2 Model validation ....................................................................................................................... 103

7.3 Parametric studies ..................................................................................................................... 104

Chapter 8. Conclusions and future work .................................................................................... 115

8.1. Conclusions .............................................................................................................................. 115

8.2 Future work ............................................................................................................................... 117

Page 8: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

viii

LIST OF TABLES

Table 3-1. Elastic properties of the constituents of the RVE. .................................................................... 51

Table 3-2. Parameters of the Cohesive Zone Model used to describe the fiber-matrix interface. ............. 51

Table 4-1. Elastic ply properties for the [±45/902]s CFRP laminate studied by Asadi [26] ....................... 58

Table 4-2. Parameters for the non-linear creep model described by Eq. (3.14), which were obtained using

the experimental data in Asadi [26]. .......................................................................................................... 58

Table 4-3. Parameters used for the creep compliance model, Eq. (3.14), as obtained by fitting the

experimental creep curve for 90° unidirectional plies, at a stress level of 510 MPa [121] ........................ 63

Table 5-1. Parameters for the viscoelastic part of the model [187] ........................................................... 72

Table 5-2. Elastic properties of the GFRP plies, obtained from [187]. ...................................................... 72

Table 5-3. Damage model parameters for the glass fiber/epoxy composite .............................................. 73

Table 5-4. Parameters for the viscoelastic part of the model [24] ............................................................. 80

Table 5-5. Elastic properties of the plies, measured at 93 °C, obtained from Tuttle et al [24] .................. 80

Table 5-6. Parameters for the viscoplastic part of the model [24] ............................................................. 80

Page 9: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

ix

Table 5-7. Damage model parameters for IM7/5260 ................................................................................. 80

Table 6-1. Fiber dimensions studied in this thesis ..................................................................................... 95

Table 7-1. Material properties of CFRP and GFRP unidirectional lamina .............................................. 106

Table 7-2. Maximum degradation of material properties of laminates with different stacking sequences due

to low-velocity impact. ............................................................................................................................. 110

Page 10: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

x

LIST OF FIGURES

Figure 1-1. U.S. greenhouse gas emissions breakdown by economic sector [10] ---------------------- 2

Figure 1-2. Proportion of materials, by weight, making up the structure of the modern car[19] ------------ 3

Figure 1-3. CO2 emissions of natural (NMT) and glass fibre (GMT) composites over their lifetime [16].

From “Biocomposites reinforced with natural fibers: 2000-2010” by Faruk O. et al., 2010 Prog. Polym.

Sci., Vol. 37, p. 1570. Copyright (2012) by Elsevier. Reprinted with permission. ----------------------------- 4

Figure 1-4. Market growth for natural fibre composites [21] ------------------------------------------------------- 5

Figure 1-5. Improvement of NF-DLFT hybrid bio-composite impact strength with glass fiber hybridization

[40]. From “Injection-Molded Short Hemp Fiber / Glass Fiber- Reinforced Polypropylene Hybrid

Composites — Mechanical , Water Absorption and Thermal Properties”, by Panthapulakkal S. and Sain

M., 2006 J. Appl. Polym. Sci., Vol. 103, p. 2438. Copyright (2006) by Wiley Periodical, Inc. Reprinted

with permission. ------------------------------------------------------------------------------------------------------------- 6

Figure 1-6. Creep behavior of NF-DLFT for different amounts of glass and hemp fibres[39]. -------------- 8

Figure 2-1. Schematic explaining the structure of a multi-directional laminate [44]. ------------------------- 10

Figure 2-2. Schematic explaining the difference between a short fiber composite and a laminate [43]. -- 10

Figure 2-3. Schematic showing matrix micro-cracking in a cross-ply laminate [47]. ------------------------ 11

Page 11: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

xi

Figure 2-4. Crack density evolution versus applied stress in the 90° ply of a glass/epoxy cross-ply. From

“On matrix crack growth in quasi-isotropic laminates - I. Experimental investigation”, by Tong J. et al,

1997 Compos. Sci. Technol., Vol. 57, p. 1529. Copyright (1997) by Elsevier. Reprinted with permission.

--------------------------------------------------------------------------------------------------------------------------------- 13

Figure 2-5. Damage mechanisms in a short fiber composite. (a) Fiber-matrix debonding mechanism [66].

(b) Fiber pull-out [67]. (c) Fiber fracture [68]. (d) Matrix micro-cracking initiating from fiber tips [61].

Last figure from “Microfailure behaviour of randomly dispersed short fiber reinforced thermoplastic

composites obtained by direct SEM observation.”, by Sato N. et al, 1991 J. Mater. Sci., Vol. 26, p. 3892.

Copyright (1991) by Chapman and Hall Ltd. Reprinted with permission. -------------------------------------- 14

Figure 2-6. Effect of multiple creep-recovery tests on the time-dependent strain response of a bio-

composite material. From “Characterization and modeling of performance of Polymer Composites

Reinforced with Highly Non-Linear Cellulosic Fibers” by Rozite L. et al, 2012 IOP Conf. Ser. Mater. Sci.

Eng. , Vol. 31, p. 4. Copyright (2012) by IOP Publishing Ltd. Reprinted with permission. ---------------- 15

Figure 2-7. Evolution of damage quantified at the microscopic level (a) and at the macroscopic level (b),

for a sheet molding compound versus applied strain for different strain rates. From “Multi-scales

modelling of dynamic behaviour for discontinuous fibre SMC composites” by Jendli et al., 2009 Compos.

Sci. Technol., Vol. 69, p. 101. Copyright (2008) by Elsevier. Reprinted with permission.------------------ 16

Figure 3-1. Schematic showing the different modes of ply micro-cracking in a multi-directional laminate.

VE refers to the viscoelastic properties of the plies; VP refers to the viscoplastic behaviour; QS is for quasi-

static loading. The top right graph shows the loading scheme used in the current thesis. -------------------- 29

Page 12: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

xii

Figure 3-2. Flowchart explaining the MATLAB program used to performed the progressive damage

simulations under viscoelastic-viscoplastic creep. ------------------------------------------------------------------- 36

Figure 3-3. (a) SDM micro-damage FE model for a [0/90/∓45]s quasi-isotropic laminate, with micro-

cracks in the 0, 90 and -45 plies and periodic in-plane boundaries. (b) Stress contour (𝜎𝑦𝑦) in the RVE

under uniaxial strain (휀𝑥𝑥 = 0.5%, 휀𝑦𝑦 = 0, 𝛾𝑥𝑦 = 0) with a single crack in the 90° ply. ----------------- 45

Figure 3-4 . Model development for the short fiber RVE. (a) Short fiber RVE used in this thesis to study

micro-damage mechanisms in short fiber composites. (b) Top face of the RVE. (c) Side face of the RVE.

Image from “Effect of the matrix behavior on the damage of ethylene-propylene glass fiber reinforced

composite subjected to high strain rate tension”, by Fitoussi J. et al, 2013 Composites Part B, Vol. 45(1),

p. 1183. Copyright (2012) by Elsevier. Reprinted with permission. --------------------------------------------- 50

Figure 3-5. FE model of the short fiber composite. ------------------------------------------------------------------ 50

Figure 3-6. (a) Validation of the implementation of the non-linear viscoelasticity model for HDPE into

Abaqus in tabular form. (b) Yield stress versus applied strain rate. --------------------------------------------- 52

Figure 3-7. Effect of interface stiffness on composite RVE modulus -------------------------------------------- 53

Figure 3-8. (a) Effect of mesh size on stress-strain response of composite RVE with linear elastic

components. The legend refers to the element size in the central region of the RVE around the fiber tips ;

when the mesh size was constant throughout the RVE, the term ‘uniform’ is used. (b) Effect of mesh size

on stress-strain response of composite RVE with a non-linear viscoelastic matrix under a strain rate of 10-5

s-1. ------------------------------------------------------------------------------------------------------------------------------ 54

Page 13: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

xiii

Figure 3-9. Synergistic Damage Mechanics (SDM) RVE, showing the main features of the model,

including uniform ply crack spacing, cracks parallel to the fibers in each layer, cracking in multiple off-

axis orientations and multi-axial loading scenarios. ----------------------------------------------------------------- 56

Figure 3-10. General framework for the numerical implementation of the hierarchical multi-scale

methodology. ---------------------------------------------------------------------------------------------------------------- 56

Figure 4-1. SDM model results and comparison to experimental data [26] for a quasi-static test performed

on a [±45/902]s carbon-fibre/epoxy laminate: a) crack density evolution in different plies; b) stiffness

degradation; c) stress-strain curve. The SDM model used here does not include viscoelastic effects. ---- 59

Figure 4-2. Results of the extended SDM model for creep strain (a) and damage evolution in the 90° ply

(b) for a [±45/902]s laminate subjected to a constant creep stress of 45 MPa, and comparison to

experimental data [32]. Note that the initial experimental creep strain in a) at time t = 0 s was adjusted so

that the predicted results and the experimental data coincided at the onset of creep. ------------------------- 60

Figure 4-3. Predicted crack density evolution for a [03/903]s cross-ply laminate for different levels of strain

rate (low, medium and high) using the SDM methodology, along with experimental data from Nguyen and

Gamby [89]. ----------------------------------------------------------------------------------------------------------------- 62

Figure 4-4. Predicted and experimental crack density evolution during quasi-static loading for a [0/903]s

cross-ply laminate; the prediction assumes no viscoelastic effects. ---------------------------------------------- 62

Figure 4-5. a) Predicted and experimental [121] 90° ply crack density evolution for a [0/903]s cross-ply

laminate subjected to a constant stress of 530 MPa and b) corresponding laminate creep strain. ---------- 64

Page 14: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

xiv

Figure 4-6. Predicted creep strain evolution for [±60/902]s and [±75/902]s CFRP laminates, with and

without the effects of damage, subjected to a constant creep stress of 19 MPa. ------------------------------- 65

Figure 4-7. The predicted evolution of compliance is shown here for the [±75/902]s laminate at a creep

stress of 19 MPa: a) different components of compliance (creep and elastic), b) stiffness degradation with

crack density in the 90° ply, and c) evolution of 90° ply crack density for creep and quasi-static loading.

--------------------------------------------------------------------------------------------------------------------------------- 66

Figure 4-8. Predicted crack density evolution in a [±75/902]s laminate in the angle plies, under creep at 19

MPa and quasi-static loading. The figure shows that cracks evolve more slowly during creep than during

quasi-static loading at low strain levels. ------------------------------------------------------------------------------- 67

Figure 4-9. Predicted COD evolution for a 90° ply in a [±75/902]s laminate undergoing creep at 19 MPa,

showing a clear increase linked to creep strain evolution. The increased in COD affects crack density

evolution through its effect on the energy release rate WI. --------------------------------------------------------- 68

Figure 4-10. a) Plot showing the predicted evolution of the elastic and time-dependent (creep) components

of the [±75/902]s laminate compliance, at two different stress levels, demonstrating the impact of damage

on compliance increase. b) Plot showing the predicted evolution of the different components (creep and

elastic) of the compliance for the two different laminates, at a stress of 19 MPa. ----------------------------- 68

Figure 5-1. Schematic showing the different modes of ply micro-cracking in a multi-directional laminate.

VE refers to the viscoelastic properties of the plies; VP refers to the viscoplastic behaviour; QS is for quasi-

static loading. The top right graph shows the loading scheme used in the current chapter. ----------------- 71

Page 15: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

xv

Figure 5-2. Flowchart explaining the MATLAB program used to perform the progressive damage

simulations under viscoelastic-viscoplastic creep. ------------------------------------------------------------------- 72

Figure 5-3. a) Viscoelastic-viscoplastic creep strain prediction for a [±45]s GFRP laminate under a creep

stress of 50 MPa, and comparison to a previous model from the literature [187]. b) Evolution of crack

density under quasi-static loading for a cross-ply GFRP composite, and comparison to experimental data

[188] c) Time-dependent crack density evolution in each layer of a [±45/902]s CFRP laminate predicted by

the model, and comparison to experimental data under a creep stress of 45 MPa [26]. (d) Predictions of

stiffness degradation with respect to crack density in a [0/90/∓45]s GFPR laminate using the SDM model

(Eqs. 3.1-3.3) and independent FEA simulations in APDL -------------------------------------------------------- 75

Figure 5-4. Crack density evolutions versus simulation time for a [0/90/∓45]s GFRP laminate loaded to an

axial load of 210 MPa. a) 90° ply crack density. b) 45° ply crack density. ------------------------------------- 76

Figure 5-5. Crack density versus axial stress with different levels of transverse stress in the 90° and 45°

plies of the quasi-isotropic GFRP [0/90/∓45]s CFRP laminate before (t=0) and after (t= tend) creep

deformation. ----------------------------------------------------------------------------------------------------------------- 76

Figure 5-6. Evolution of crack density and normalized stiffness for the GFRP composite under different

levels of transverse loading: (a) Quasi-isotropic [0/90/∓45]s laminate under an axial load of 150 MPa, and

(b) cross-ply laminate under an axial load of 150 MPa ------------------------------------------------------------- 77

Figure 5-7. Axial modulus degradation versus applied axial stress for (a) [0/90/∓45]s and (b) [0/90]s GFRP

laminates, at the start of the creep tests, and at the end, for different biaxial loads. --------------------------- 78

Page 16: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

xvi

Figure 5-8. The lines in the stiffness degradation contour map correspond to constant levels of stiffness

loss after time-dependent deformation under biaxial loading of (a) [0/90/∓45]s and (b) [0/90]s GFRP

laminate. ---------------------------------------------------------------------------------------------------------------------- 79

Figure 5-9. Stiffness degradation contour lines. The lines in the stiffness degradation contour map

correspond to constant levels of stiffness loss after viscoplastic deformation under biaxial loading of (a)

[0/90/∓45]s and (b) [0/90]s CFRP laminate. -------------------------------------------------------------------------- 82

Figure 5-10. Stiffness degradation contour lines for different temperatures. a) Stiffness degradation contour

lines at the end of the creep simulations (T = tend) at two different temperatures for (a) [0/90/∓45]s and (b)

[0/90]s IM7/5260 laminates. This material has a glass transition temperature of 260 °C. ------------------- 83

Figure 5-11. (a) Stress-strain curves for a [0/90/∓45]s GFRP laminate under uniaxial loading using

different material models. (b) Stress-strain curves for a [0/90/∓45]s CFRP laminate. ----------------------- 84

Figure 5-12. (a) Time-dependent evolution of transverse ply stress for the [0/90/∓45]s under uniaxial

constant load of 300 MPa, during ramp-up and subsequent creep deformation. (b) Effect of including

viscoelastic-viscoplastic stress relaxation on crack multiplication predictions for a [0/90/∓45]s GFRP

laminate. ---------------------------------------------------------------------------------------------------------------------- 85

Figure 6-1. (a)Von Mises stress contour for the short fiber composite RVE undergoing longitudinal

deformation with inset showing the stress in the neighborhood of the fiber tips. Stresses are expressed in

MPa. (b) Longitudinal displacement contour plot of the RVE under an applied strain of 0.1 %.

Displacements are expressed in μm. (c) Local damage variable for a fiber Df = 40 μm and Lf = 200 μm.

(d) Global damage variable versus applied strain and secant modulus loss curve ----------------------------- 89

Page 17: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

xvii

Figure 6-2. (a) Axisymmetric short glass fiber/epoxy composite RVE used to validate the model. (b)

Displacement response under 1% strain. (c) Debonded area fraction versus applied strain for a short glass

fiber/epoxy composite, and comparison to the work of Pan and Pelegri [200]. -------------------------------- 90

Figure 6-3. Von Mises stress distribution in the short fiber RVE with Df = 40 µm and Lf = 200 µm under

different strain levels with a linear elastic HDPE matrix. ---------------------------------------------------------- 91

Figure 6-4. Von Mises stress distribution in the short fiber RVE with Df = 40 µm and Lf = 200 µm under

different strain levels for a viscoelastic matrix under a strain rate of 10-5 s-1. ---------------------------------- 92

Figure 6-5. (a) Evolution of the damage variable (fraction of debonded area) with respect to applied strain

for different strain rates in a short glass fiber composite with fiber diameter Df = 40 μm. (b) Stress-strain

curves. ------------------------------------------------------------------------------------------------------------------------ 94

Figure 6-6. (a) Energy dissipation in the RVE of a short glass fiber composite (Df =40 μm) for a viscoelastic

matrix deformed at 10-1 s-1. (b) Inelastic strain evolution under different strain rates. ------------------------ 94

Figure 6-7. Evolution of COD under different strain rates for an interfacial crack at the fiber tips in a short

glass fiber composite with Df = 40 μm. -------------------------------------------------------------------------------- 94

Figure 6-8. (a) Shear stress distribution along the fiber matrix interface at two different applied strains, for

a linearly elastic matrix, and a viscoelastic matrix under a strain rate of 10-5 s-1. (b) Axial stress distribution.

--------------------------------------------------------------------------------------------------------------------------------- 96

Figure 6-9. (a) Damage variable versus applied strain for 3 fiber dimensions with a linearly elastic matrix.

(b) Maximum damage reached for 3 fiber dimensions as a function of strain rate. --------------------------- 96

Page 18: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

xviii

Figure 6-10. Effect of IFSS on strain energy density at 4 % applied strain: (a) 𝜏𝐼𝐹𝑆𝑆 = 11 𝑀𝑃𝑎, (b) 𝜏𝐼𝐹𝑆𝑆 =

20 𝑀𝑃𝑎, (c) 𝜏𝐼𝐹𝑆𝑆 = 30 𝑀𝑃𝑎 for a matrix loaded under a strain rate of 10-1 s-1. The glass fiber diameter is

20 μm. ------------------------------------------------------------------------------------------------------------------------- 98

Figure 6-11. Evolution of maximum strain energy density versus applied strain for three different values

of the interfacial shear strength. The glass fiber diameter is 20 μm. --------------------------------------------- 98

Figure 6-12. (a) Damage evolution for different types of fibers with different mechanical properties with

a linear elastic matrix. (b) Damage variable at 4% strain as a function of strain rate for different fiber types.

In all cases, the fiber diameter is Df = 20 µm. ------------------------------------------------------------------------ 99

Figure 7-1. Validation of the implementation of the SDM model into Abaqus FEA software using a

VUMAT subroutine, as compared to the MATLAB predictions: (a) Modulus degradation. (b) Crack

multiplication in the 90º ply of the [0/90/∓45]s GFRP composite. --------------------------------------------- 103

Figure 7-2. Effect of strain rate on stiffness degradation as predicted by the rate-dependent SDM model

implemented into MATLAB for the [0/90/∓45]s GFRP composite. ------------------------------------------- 103

Figure 7-3. Comparisons of time histories of contact force for a [0/90]6s composite laminate between

simulation and experiments. -------------------------------------------------------------------------------------------- 104

Figure 7-4. Schematic of an automotive bumper subject to low-velocity impact. (a) an example of the mesh

pattern, (b) geometrical dimensions of the bumper and impactor. ---------------------------------------------- 106

Figure 7-5. Effect of the impactor’s initial velocity on time histories of the contact force for a [0/90/∓45]s

GFRP laminate. ----------------------------------------------------------------------------------------------------------- 107

Page 19: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

xix

Figure 7-6. Time history of the evolution of crack density for the different plies of a [0/90/∓45]s GFRP

laminate. Initial velocity of the impactor is equal to : (a) 4.6 m/s, (b) 8.4 m/s, (c) 13.1 m/s. -------------- 107

Figure 7-7. Contour of 0° ply crack density for [0/90/∓45]s GFRP laminate for different initial velocities

of the impactor: (a) v = 4.6 m/s, (b) v = 8.4 m/s, (c) v =13.1 m/s. ---------------------------------------------- 108

Figure 7-8. Effect of laminate stacking sequence on (a) displacement versus contact force and, (b) time

histories of contact force for GFRP laminates under an impactor velocity of 4.6 m/s. --------------------- 108

Figure 7-9. Effect of laminate stacking sequence on (a) displacement versus contact force, and (b) time

histories of contact force for CFRP laminates under an impactor velocity of 4.6 m/s ---------------------- 109

Figure 7-10. Effect of rate-dependency on (a) displacement versus contact force, (b) time histories of

contact force for a CFRP [0/90/∓45]s laminate. -------------------------------------------------------------------- 110

Figure 7-11. Effect of the cross section profile of the bumper on 0º crack density contours for a [0/90/∓60]s

CFRP laminate under an impactor velocity of 8.4 m/s. ----------------------------------------------------------- 111

Figure 7-12. Effect of bumper cross-section shapes on (a) displacement versus contact force and, (b) time

histories of contact force for CFRP laminates for the [0/90/∓60]s laminate. --------------------------------- 111

Figure 7-13. Crack density in the 0° ply of a GFRP [0/90/∓45]s laminate under low-velocity impact at 13.1

m/s : (a) neglecting rate-dependent behavior and (b) including rate-dependent behavior. ----------------- 112

Page 20: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

xx

Figure 7-14. Effect of including rate-dependency on crack density evolution in the 0° ply of the [0/90/∓45]s

GFRP bumper under low velocity impact. -------------------------------------------------------------------------- 112

Page 21: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

1

Chapter 1. Introduction

In this chapter, the main motivations for the research conducted in this PhD thesis are provided. The

overall economic and societal challenges the automotive industry is facing are explained, and the gaps in

scientific knowledge preventing the incorporation of composite materials into the transportation system

are reviewed.

1.1 Lightweighting in the transportation sector

The transportation sector is responsible for an estimated 27 % of total anthropogenic green house gas

emissions (see Figure 1-1) and is therefore a major contributor to global warming. In order to reduce

mankind’s reliance on fossil fuels, while still maintaining the strong economic performance that the

transportation systems affords our societies, the efficiency of vehicles needs to be improved so that

performance is maintained, while fuel consumption is reduced [1]. Given the initial reluctance of car

manufacturers to implement effective strategies for improving the efficiency of their vehicles,

governmental regulations [2] have been enacted in North America and the European Union for CO2

emissions reductions. Countries in the EU are requiring a reduction in CO2 emissions from 140 grams of

CO2 emitted per car per kilometer driven in 2010, to 75 grams of CO2/km in 2025: if a car manufactured

in 2025 does not meet the CO2 emission target set out by the EU, a penalty of 12 350 EUR for every vehicle

produced will be imposed.

The efficiency of the transportation infrastructure can be improved by improving the energy efficiency

of the drivetrain, and reducing air drag, vehicle weight and rolling resistance. It has been estimated that by

2030 fuel economy can be increased between 50 and 100% by using these methods [1]. More

technologically advanced methods would involve the full electrification of the car and truck fleet [2]. An

easier method of improving efficiency is the reduction of the weight of cars, trucks and buses. It has been

estimated that 0.08 g CO2 /km can be saved per kilogram of car weight reduction[2]. A recent MIT study

showed that a weight reduction of 35 % of the entire transportation fleet could be achieved by 2035 [3],

reducing fuel consumption by 12 to 20 %, which would lead to significantly lower GHG emissions. Two

recent Life Cycle Analyses [4], [5] have also shown that lightweighting through material substitution

would lead to significant reductions in CO2 emissions over the entire lifecycle over the vehicles.

Lightweighting is achieved by replacing the materials in the vehicle by lighter materials that can

sustain equal mechanical constraints. By improving the mechanical properties of the materials used in a

vehicle, the weight can be reduced while still maintaining a safe environment for the passengers and

satisfactory structural performance, thereby increasing gas mileage. As shown in Figure 1-2, current cars

Page 22: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

2

are mainly made with plastics, aluminum and steel [2]. The production of these materials is highly energy

intensive; they could be replaced by lighter materials, such as High Strength Steels (HSS), aluminum alloys

and carbon fiber composites [2]. Polymer-based composite materials are a candidate of choice to replace

the traditional metallic materials, due to their very high stiffness-to-weight and strength-to-weight ratios.

Composite materials being considered for automotive applications combine a stiff and strong fibrous

material such as carbon or glass fiber, with a weaker polymer which protects the fibers from the

environment, and transfers the loads applied to the composite onto the fibers [6]. Composites are generally

as strong as metals but are less dense and can therefore provide significant weight savings.

Several automotive companies are trying to optimize the material design process to enhance the

properties of the metals in their vehicles. Ford Company, in collaboration with Magna, are reducing car

weight by designing lighter and stronger metals through improved processing [7]. General Motors have

publicized similar plans of weight reduction through optimization of material performance [8]. High-

strength steel, aluminum, magnesium, as well as carbon fiber composites are being considered for reducing

car weight [2]. Among these, carbon fiber composites are the lightest alternative, but also the most

expensive. Market forecasts suggest that the price of carbon fiber will decrease significantly within the

coming years, which together with its much higher performance would make it competitive against

metallic alloys. Moreover, although the production of a car containing a significant fraction of carbon fiber

materials might only be feasible for high-end luxury and sports cars, incorporating them in at least some

strategic components in an average vehicle would also be a useful strategy for lightweighting. Formula 1

racing car chassis are already made of carbon fiber [9] due to its very low density and very high rigidity.

Figure 1-1. U.S. greenhouse gas emissions breakdown by economic sector [10] .

The automotive industry is already using composite materials in some applications. According to

Lucintel [11], in 2018 the total demand for composite materials in the automotive industry reached 4 billion

lbs. These composites would include glass fiber composites and bio-composites, as well as carbon fiber

components. Demand for carbon fiber composites is expected to reach 90 million lbs by 2025 [12]. The

Page 23: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

3

carbon fiber composites can be used in highly stressed regions of the vehicles, such as the chassis and side

panels. Less strong composites such as bio-composites are already being used as seat backs and door

panels.

Significant car weight reduction could be achieved by incorporation of carbon fiber composites. Other

types of composites could also be used to improve the overall sustainability of the transportation sector.

For example, bio-composites are a particular type of composite materials in which at least one of the

constituents is derived from the natural environment [5], [13]. Bio-composites present several advantages

over plastics used traditionally. According to [14], some bio-composites can offer the same mechanical

performance but for a lower weight (the stiffness of bio-based fibres can be comparable to that of glass

fibres [15]), which will reduce vehicle weight. They are made using renewable resources, such as wood,

hemp and natural materials, which can be grown easily, reducing mankind’s reliance on limited fossil fuel

supplies. The manufacturing process is also sustainable as it requires 80 % less energy than glass fiber-

based composites (see Figure 1-3[16]), currently being used as a replacement to metals. They are easy to

recycle. Some bio-composites can also be bio-degradable, and could be buried in the ground without

impacting the environment [14]. Currently, the equivalent of 25 % of the weight of a car cannot be recycled.

Replacing thermosets with bio-composites and biodegradable and/or recyclable plastics will help alleviate

this issue [17]. They can also be incinerated without releasing as much GHGs as synthetic materials, or

they can be composted. Finally, their cost is low compared to synthetic materials [16] such as glass fiber

composites. Automakers also wish to project a good public image to customers, and at a time when

environmental problems are of increasing concern to our societies, investing in more sustainable products

is considered by some as a judicious policy [18] to improve brand reputation. Due to these factors, the

market for natural fiber composites is expected to continue growing in the coming years (see Figure 1-4).

Figure 1-2. Proportion of materials, by weight, making up the structure of the modern car[19]

Page 24: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

4

Figure 1-3. CO2 emissions of natural (NMT) and glass fibre (GMT) composites over their lifetime [16]. From “Biocomposites

reinforced with natural fibers: 2000-2010” by Faruk O. et al., 2010 Prog. Polym. Sci., Vol. 37, p. 1570. Copyright (2012) by

Elsevier. Reprinted with permission.

A significant impediment to the large-scale implementation of bio-composites as car components is

their lower impact strength and their highly non-linear behaviour. The improvement of impact behaviour

is critical for automotive components. Because the natural fibres are short, the energy dissipation

mechanisms occurring upon impact are not sufficient to absorb the collision energy. Natural fibre enhanced

direct-long fibre thermoplastics (NF-DLFT) manufactured by the Center for Bio-composites and Bio-

materials Processing (CBBP) at the University of Toronto are a possible compromise between the

mechanical performance of synthetic glass fibre-based composites and the environmental sustainability of

bio-composites. NF-DLFTs are hybrid bio-composites reinforced with both natural hemp fibres and glass

fibres, embedded in a polypropylene (thermoplastic) matrix. The presence of glass fibres increases the

strength of the material, making them suitable for structural applications, and improves the impact strength

by providing more energy dissipation mechanisms (see Figure 1-5). A new manufacturing process has

recently been developed to further increase the length of the glass fibers and thereby improve impact

performance [20].

Overall, both continuous fiber composites and bio-composites (such as NF-DLFT) can, if properly

implemented into the transportation infrastructure, play a major role in the lightweighting targets of the

automotive industry. Continuous fiber composites can be used in the structurally demanding components

of the vehicle, and bio-composites can serve as minor load-bearing components while weighing less than

traditional materials and increase the overall sustainability of the car. However, in order to incorporate

composite materials in the automotive industry, significant technical challenges still need to be overcome.

Page 25: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

5

Figure 1-4. Market growth for natural fibre composites [21]

1.2 Modelling the mechanical performance of composites

Reducing the weight of vehicles will help reduce greenhouse gas emissions from vehicles. However,

in order to design vehicle with composite components that demonstrate equivalent performance to the

metals used currently, their mechanical behaviour needs to be predictable. Current mathematical models

are not accurate enough to predict the constitutive response of multi-directional composites loaded under

arbitrary multi-axial scenarios and undergoing progressive damage in multiple plies at the same time.

Moreover, time-dependent effects induced by the viscoelastic and viscoplastic properties of the polymer

matrix are still not completely understood. According to the National Research Council [22], the lack of

predictive tools for the evolution of damage in composite materials has led manufacturers to overdesign

their products, preventing the full utilization of the mechanical capabilities of composite materials. This

is an especially important issue in the context of lightweighting.

Composite materials are known to exhibit time-dependent behavior, due to the viscoelasticity and

viscoplasticity of the matrix [23], [24]. Bio-composites are especially prone to time-dependent

deformation, even under relatively low loads [25] due to the shortness of the fibers and the mechanical

properties of the matrix. In addition to causing supplementary deformation which leads to lower composite

stiffness over time, these non-linear properties are known to affect the evolution of damage in laminates,

as well as short fiber composites [26], [27]. Studies on laminates have shown that viscoelasticity and

viscoplasticity can cause the evolution of damage during creep loading [26]. Studies on short fiber

composites have also shown that the rate of loading affects damage processes at the microscopic scale

[28], [29], most likely because of the time-dependent behavior of the matrix (see Figure 1-6 for the creep

behavior of NF-DLFT).

While there have been many studies investigating damage mechanisms in composites [30]–[36], either

from a modelling perspective or an experimental approach, most models currently available are only

approximately accurate, or are restricted to very specific types of composites, such as cross-plies. In order

Page 26: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

6

to improve the efficiency of current design processes, a multi-scale computational approach is needed,

which can provide much faster assessment of the performance of a composite than an experimental

approach. In particular, a quantitative understanding of the process of matrix micro-cracking and fiber-

matrix debonding at the microscopic scale will accelerate material selection and the overall design phase

of new composite materials. Moreover, as mentioned above, the time-dependent properties of the matrix,

typically modelled using the theories of viscoelasticity and viscoplasticity, can affect damage progression

in the composite [37]–[39]. To effectively design composite materials which do not fail prematurely due

to the time-dependent properties of the matrix, an understanding of the interaction between viscoelasticity,

viscoplasticity and damage is required.

Figure 1-5. Improvement of NF-DLFT hybrid bio-composite impact strength with glass fiber hybridization [40]. From “Injection-

Molded Short Hemp Fiber / Glass Fiber- Reinforced Polypropylene Hybrid Composites — Mechanical , Water Absorption and

Thermal Properties”, by Panthapulakkal S. and Sain M., 2006 J. Appl. Polym. Sci., Vol. 103, p. 2438. Copyright (2006) by Wiley

Periodical, Inc. Reprinted with permission.

The subject of this thesis is concerned with predicting damage evolution in both continuous fiber

composites and short fiber bio-composites, while taking into account the viscoelastic-viscoplastic

properties of the constituents. Laminates and short fiber composites have different mechanical properties

and undergo damage in fundamentally different ways. However, as will be shown in this thesis, the issues

facing damage modelling in these two types of composites are the same and can be addressed using a

similar methodology. It will be shown the tools that are developed for continuous fiber composites are

applicable to short fiber composites as well.

1.3 Thesis objectives

A proper prediction of the progression of damage and the effects of viscoelasticity and viscoplasticity

is necessary to optimize the performance of composites by tailoring the manufacturing process and the

microstructure (fibre geometry, fibre volume fraction and distribution, thermoplastic mechanical

properties etc.). The motivations behind this thesis are therefore to assist in the design of composites that

permit greater environmental sustainability, and that can perform sufficiently well mechanically to be used

as structural car components. The goal of this work in particular is to provide a comprehensive predictive

Page 27: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

7

methodology capable of assessing the long-term mechanical performance of composite structures, using a

multi-scale modelling methodology to predict damage evolution in viscoelastic-viscoplastic composites.

The multi-scale model developed in this thesis includes the microscopic scale of the fiber-matrix

interface, the mesoscopic level of the homogenized plies, and the structural level of an automotive

component. To achieve the thesis goal, the following objectives have been decided upon:

1. Predict damage evolution in viscoelastic/viscoplastic laminates undergoing creep deformation.

2. Implement the damage model into FEA software and examine the effect of quasi-static loading

and impact on a composite structure. This will provide a predictive tool for the performance of

composites, whose results can be compared to experiments and used to design new structures.

3. Determine the effect of viscoelastic and viscoplastic matrix behavior on the evolution of damage

and the competition between damage mechanisms under varying strain rates in short fiber

composites.

In this thesis, a multi-scale approach based on micro-damage FE modelling of matrix micro-cracking

is combined with Schapery’s equations for calculating creep strain arising from viscoelasticity and

viscoplasticity in multi-directional laminates. The model is applied to different composite structures

(different stacking sequences and ply thicknesses) under arbitrary in-plane loading scenarios. The results

of this analysis are reported in Chapters 4 and 5 of this thesis and provide detailed insights into the

interactions between damage evolution and creep deformation in structural laminates.

In Chapter 6, the multi-scale methodology used for laminates is extended by developing a Cohesive

Zone Model to predict fiber/matrix debonding in a polyethylene-based short fiber composite material, for

potential use in automotive applications. The effect of strain rate is studied, which illustrates and quantifies

the effect of matrix viscoelasticity on competing deformation processes in short fiber composites.

Lastly, in Chapter 7, the multi-scale framework is extended to the structural scale by implementing

the viscoelastic-viscoplastic damage model for laminates as an Abaqus VUMAT. A structural analysis of

a car bumper under low-velocity impact is conducted, and the effects of stacking sequence, material system

and impactor velocity are studied. Rate-dependency is included in the custom material model to represent

the effects of viscoelastic and viscoplastic material properties of the matrix.

Page 28: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

8

Figure 1-6. Creep behavior of NF-DLFT for different amounts of glass and hemp fibres[39].

Page 29: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

9

Chapter 2. Literature review

In this chapter, both the nature and the quantitative evolution of damage in laminates and short fiber

composites which have been observed in experiments are summarized. The different models published in

the literature which have been used to predict damage in viscoelastic-viscoplastic laminates and short fiber

composites are then presented and evaluated. Based on this literature review, the main gaps in the current

understanding of the interaction between damage and time-dependent properties are identified to motivate

the thesis objectives presented in Chapter 1.

2.1 Composite materials

A composite is a material which consists of at least two heterogeneous phases [6] with significantly

different mechanical properties. For the composites studied in this thesis, one phase is fibrous and is the

primary load-bearing component of the composite, while the matrix phase is more compliant and weaker,

and surrounds the fibers. The advantages of composites with a polymer matrix and stiff synthetic fibers

such as glass or carbon relative to ceramics and metallic materials are improved control of the mechanical

properties, and better specific strength and stiffness [41]. Also, their fatigue life is often better than for

aluminum [6] and their thermal stability can also be tailored more easily, which makes them a prime

candidate for aerospace applications.

There are different classes of composite materials, which are distinguished in this thesis by the

configuration and distribution of the fibers. In the case of laminates, shown schematically in Figure 2-1,

the fibers are continuous throughout the length of the component, and are arranged parallel to each other

in a ply; the laminate consists of multiple plies, oriented in different directions, and stacked on top of one

another. A laminate’s properties are in part governed by its stacking sequence, which corresponds to the

orientations of the fibers in the plies with respect to the primary loading axis. In this thesis, the stacking

sequences will all be symmetrical with respect to the mid-plane of the laminate, which allows for the study

of in-plane mechanical behaviour. Three types of laminates will be studied in this work: symmetric cross-

plies ([0n/90m]s) which consist of plies oriented along the loading axis and perpendicular to it, laminates

with [±θ/902]s stacking sequence, and symmetric multi-directional laminates [0/90/∓θ]s, where θ is an

angle between 45 and 75°. Depending on the stacking sequence, the elastic properties, the time-dependent

properties, as well as the damage evolution of the composite, will be affected. The choice of stacking

sequence is therefore an effective design tool for engineers which allows them to tailor the properties of

the composite for specific thermo-mechanical loading scenarios.

Page 30: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

10

In the case of short fiber composites, the fibers are much shorter (on the order of less than a millimeter

to a few centimeters), and are dispersed throughout the matrix [42]. The fibers can be randomly oriented

or aligned, depending on the manufacturing process. The length of the fibers, their orientation, and their

distribution affect the macroscopic behaviour of the short fiber composite. Figure 2-2 [43] shows the

differences between a short fiber composite, and a laminate made with unidirectional fibers.

Figure 2-1. Schematic explaining the structure of a multi-directional laminate [44].

Figure 2-2. Schematic explaining the difference between a short fiber composite and a laminate [43].

2.2 Damage mechanisms in composites

The evolution of damage in composite materials is governed by the structure of the composite [41]

and the mechanical properties of the constituents. The patterns of damage evolution in laminates on the

one hand, and short fiber composites on the other hand, are known to be significantly different. First

damage mechanisms in laminates will be reviewed, concentrating specifically on matrix micro-cracking,

which is the focus of this thesis. Damage mechanisms in short fiber composites will be reviewed next.

Viscoelastic and viscoplastic behavior will be explained, and their interaction with damage evolution

observed experimentally will be discussed.

Page 31: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

11

2.2.1 Matrix micro-cracking in laminates

Damage in laminates in the form of matrix micro-cracking initiates at the ply level [41], usually in the

plies that are oriented at more than 50° with respect to the applied load [45]. Under the transverse tensile

stress exerted on the ply, a microscopic crack is nucleated within the soft matrix, which quickly grows

parallel to the fibers and extends throughout the thickness and width of the ply (see Figure 2-3). The process

of crack nucleation in plies occurs through a combination of fiber-matrix debonding, matrix plastic damage

and failure, micro-crack nucleation and subsequent micro-crack growth throughout the thickness of the ply

and the width of the laminate [41]. Due to the constraining effect of the adjacent layers of the laminate,

the matrix micro-crack that has been nucleated will not cause immediate failure of the laminate. Instead,

as the applied load is further increased, additional cracks will nucleate, forming a series of evenly spaced

parallel cracks spanning the width and thickness of the ply. Eventually, at sufficiently high loads, the

micro-cracks can lead to other critical damage modes, such as delamination and fiber fracture [46] due to

the stress concentration induced at the crack tips.

Figure 2-3. Schematic showing matrix micro-cracking in a cross-ply laminate [47].

Many experimental studies have investigated in detail the process of matrix micro-crack initiation in

laminates [35], [48]–[51]. Ply thickness [31] has been shown to play a critical role in the crack nucleation

strain, with thinner plies leading to a larger crack initiation strain. At very small ply thicknesses, ply micro-

cracking can be suppressed entirely. Stacking sequence has also been shown to affect the initiation of

matrix micro-cracking [51]. In the case of laminates with stacking sequence [±30/90n]s and [±60/90n]s, the

orientation of the plies adjacent to the 90° ply was found to affect the ply crack initiation strain in the 90°

ply. Also, the order of the stacking sequence has been found to play a role in crack initiation [52]: when

the laminate is composed of 90° plies on the outer surface of the laminate, the crack initiation strain is

lower than for a 90° ply in the middle of the laminate. The kinetics of crack nucleation and growth depend

on the ply thickness, as well as the orientation of the cracked ply [31], [48].

Page 32: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

12

In experiments, the first micro-cracks typically nucleate at defects such as resin-rich regions, voids,

and fiber tips which induce a stress concentration that enables the opening of a micro-crack [53]. When

the loading is increased furhter, more cracks will nucleate. Because the initial distribution of defects within

the ply is random, the first few micro-cracks will be randomly distributed throughout the length of the

laminate. However, because it is more energetically favorable for a new crack to nucleate in the central

region between two previous cracks, with increasing crack density the distribution becomes more uniform.

The effect of manufacturing defects on the first stages of crack initiation and multiplication has been

studied by several authors [54], [55].

Experimental studies of damage evolution in composites traditionally focus on microscopic defects,

which lead to the occurrence of microscopic cracks which do not cause the fracture of the composite.

Instead, the effect of these micro-cracks can be modelled by defining residual stiffness matrices. On the

other hand, if the crack that is nucleated becomes sufficiently large, a model based on Linear Elastic

Fracture Mechanics, instead of Continuum Damage Mechanics, might be necessary to predict its growth

and the fracture of the part. In this thesis, the focus will be on micro-damage mechanisms, and not on

macroscopic cracks.

Following the nucleation and growth of the first matrix micro-crack, the elastic properties of the

laminate are not significantly affected due to the small size of the crack. However, upon further loading,

supplementary cracks are nucleated [56]. In order to understand matrix micro-cracking quantitatively,

crack density as a function of globally applied stress is typically plotted [57]. As seen in Figure 2-4,

following the formation of the first matrix micro-crack, the crack density increases; after a certain amount

of loading, the curve flattens. It has been found [31] that microstructural defect at the ply level can affect

the shape of the curve at the onset of crack multiplication. The thickness of the ply has been found to also

affect on crack multiplication, with lower ply thickness leading to a larger saturation crack density. The

properties of adjacent plies also affect the rate of damage evolution, and the saturation crack density [51].

Matrix micro-cracking is typically the first mode of damage in structural laminates and dominates at

low stresses. As crack density increases, however, the matrix micro-cracks can lead to delamination [48]

at ply interfaces. This delamination induced by the micro-cracks occurs due to the strong stress

concentration at the tips of the micro-cracks [58]. The cracks are usually straight through the thickness of

the ply, however, at large crack densities, interactions between cracks can cause newly nucleated cracks to

adopt curved shapes, which can affect the delamination process [59]. Other damage modes can also occur

at higher loading, such as fiber breakage [60]. Different modes of damage occur under compressive

loading, but will not be considered in this thesis.

Page 33: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

13

Figure 2-4. Crack density evolution versus applied stress in the 90° ply of a glass/epoxy cross-ply. From “On matrix crack growth

in quasi-isotropic laminates - I. Experimental investigation”, by Tong J. et al, 1997 Compos. Sci. Technol., Vol. 57, p. 1529.

Copyright (1997) by Elsevier. Reprinted with permission.

2.2.2 Damage mechanisms in short fiber composites

Damage mechanisms in short fiber composites include fiber-matrix debonding, matrix micro-cracking,

fiber pull-out and fiber breakage [40], [61]–[65]. The evolution of these mechanisms at the microscopic

scale are very complex to model due to the complex distribution of fibers throughout the part, and possibly

also their random orientation. In the case of polymer composites with a thermoplastic matrix, which can

demonstrate a high level of ductility and deformation prior to cracking, the sequence of damage

mechanisms under uniaxial loading leading to final failure of the material usually consists of the following

features [61]:

- Significant inelastic deformation initiates ahead of the fiber tips due to the mismatch in elastic

properties between the fiber and matrix.

- A void nucleates at the fiber tip following plastic deformation of the matrix and separation between

the fiber tip and the matrix.

- The micro-crack nucleated at the fiber tip propagates along the sides of the fiber leading to full

debonding between the fiber and matrix.

- Significant plastic deformation initiating from the debonded fibers occurs in the matrix in the

regions between the fiber tips.

- Eventually, extensive plastic deformation leads to matrix micro-cracking.

- The micro-cracks coalesce, leading to the growth of matrix macroscopic crack that can propagate

throughout the material.

Page 34: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

14

Figure 2-5. Damage mechanisms in a short fiber composite. (a) Fiber-matrix debonding mechanism [66]. (b) Fiber pull-out [67].

(c) Fiber fracture [68]. (d) Matrix micro-cracking initiating from fiber tips [61]. Last figure from “Microfailure behaviour of

randomly dispersed short fiber reinforced thermoplastic composites obtained by direct SEM observation.”, by Sato N. et al, 1991

J. Mater. Sci., Vol. 26, p. 3892. Copyright (1991) by Chapman and Hall Ltd. Reprinted with permission.

Images of the different mechanisms obtained through experimental observation are shown in Figure

2-5. The damage mechanisms occurring in short fiber composites that will be studied in depth in Chapter

6 of this thesis are fiber-matrix debonding and matrix inelastic deformation.

It has been shown experimentally [69], [70] that the local environment around short fibers can affect

the evolution of damage. Depending on the orientation and size of adjacent fibers as well as their relative

proximity, the local evolution of damage will be affected. It has also been found that the mechanical

properties of the matrix can affect the initiation of matrix micro-cracking and debonding [71]: depending

on the strength of the matrix, the extent of debonding and matrix micro-cracking following fiber breakage

can be different.

2.2.3 Interaction between viscoelasticity, viscoplasticity and damage evolution in composites

The effects of viscoelasticity and viscoplasticity on mechanical behavior are evident during a creep

test, where time-dependent deformation will occur under a constant load (see Figure 2-6). Due to the

(a) (b)

(c) (d)

Page 35: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

15

viscoelastic and viscoplastic properties of the material, when a load is applied, the material will keep

deforming with respect to time. When the load is removed, an immediate elastic response will be recovered.

The strain will then keep decreasing with respect to time due to viscoelasticity. A residual strain will not

be recovered with increasing time due to the viscoplastic material behaviour. A large amount of research

has been conducted on the viscoelasticity and viscoplasticity of composites [24], [72]–[79], and especially

bio-composites, due to the non-linear properties of the fibers and matrix[15], [27], [80]–[82]. While these

studies provide insights into the time-dependent behavior of composites, they do not consider the effects

of damage evolution on the mechanical response of the composites.

Figure 2-6. Effect of multiple creep-recovery tests on the time-dependent strain response of a bio-composite material. From

“Characterization and modeling of performance of Polymer Composites Reinforced with Highly Non-Linear Cellulosic Fibers”

by Rozite L. et al, 2012 IOP Conf. Ser. Mater. Sci. Eng. , Vol. 31, p. 4. Copyright (2012) by IOP Publishing Ltd. Reprinted with

permission.

The viscoelastic and viscoplastic properties of the composite can have an effect on its long-term

durability and its susceptibility to undergo progressive damage. According to a study by Guedes [83], creep

rupture can occur in laminates at loads lower than the quasi-static strength. This was attributed to the time-

dependent deformation of the composites, in which the matrix undergoes a change in structure with respect

to time, causing fracture to occur at relatively lower loads. Asadi and Raghavan [32] showed that damage

can evolve during creep loading of a CFRP laminate. Similarly, Darabi et al [84] demonstrated that rate-

dependent material properties can affect damage evolution in polymer-based composites. Practically, this

study demonstrated that for a given applied load, the damage evolution during a creep test would evolve

with respect to time. Early studies by Dillard and his collaborators [23], [85], [86] showed that under creep

loading, time-dependent cracking can occur, due to the viscoelastic properties of the laminates. Birur and

his co-workers [87], [88] similarly observed damage evolution in multi-directional laminates subject to

creep loading. Nguyen and Gamby [89] found that the matrix micro-crack multiplication in cross-ply

CFRP laminates loaded at 120 °C was affected by the loading rate, with lower loading rates leading to

higher crack density for the same amount of strain. Fitoussi et al [90] found that the rate-dependent

Page 36: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

16

behavior of the matrix affected damage evolution in Sheet Moulding Compounds (see Figure 2-7). These

studies show that damage evolution in composites is affected by the viscoelastic and viscoplastic properties

of the matrix.

Figure 2-7. Evolution of damage quantified at the microscopic level (a) and at the macroscopic level (b), for a sheet molding

compound versus applied strain for different strain rates. From “Multi-scales modelling of dynamic behaviour for discontinuous

fibre SMC composites” by Jendli et al., 2009 Compos. Sci. Technol., Vol. 69, p. 101. Copyright (2008) by Elsevier. Reprinted

with permission.

2.3 Models to predict matrix micro-cracking in laminates

Many models have been developed to attempt to predict the progression of matrix micro-cracking in

laminates [45], [91]–[96], however, to date, no model can predict progression of damage in viscoelastic-

viscoplastic multi-directional laminates under multi-axial loading. The earlier models [91], [92] are

analytical, and use the properties of the laminate (ply thicknesses and orientations, elastic properties) to

develop closed-form solutions for stiffness loss as a function of crack density. These models can be applied

to any material with the same stacking sequence, however their accuracy is limited because of the

assumptions that are made in the stress field description within the laminate. Moreover, their extension to

stacking sequences other than cross-plies and to cracking in multiple plies at the same time is not trivial,

which severely limits their predictive abilities [46]. In the case of shear-lag models [91], the effect of a

crack on the distribution of stresses within the cracked ply is calculated by solving the differential equation

for stress equilibrium, and by assuming a functional form of the displacement profile in the different layers

of the cracked laminate. From the stress profile occurring within the layers of the laminate at different

crack spacings, the stiffness can be calculated. In the case of variational methods [92], equations describing

the relationship between applied stress and internal stress distribution are developed by assuming a

particular functional description of the different components of the stress tensor in each layer of the

laminate, from which the stored elastic energy can be calculated; the principle of minimum complementary

energy is then invoked, from which an approximate stress field can be derived using variational calculus.

Page 37: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

17

Because this type of model is based on an approximate representation of the stress field, it demonstrates

limited accuracy.

In order to overcome the complexities involved when attempting to predict the stress field within a

multi-directional laminate containing cracks in multiple plies, the more recent models for predicting

progressive damage in laminates are based on Continuum Damage Mechanics (CDM), which take

advantage of the improving computational capabilities available to researchers. In CDM, the stiffness can

be predicted without first deriving the stress distribution within the cracked laminate. Instead, the geometry

of the cracks is used for predicting stiffness loss. In one of the first such models, Talreja [97] predicted

stiffness loss in different laminates as a function of crack density by defining a damage tensor in terms of

the size of the cracks (width and length) as well as their volume fraction. Following this work, Talreja [98]

developed a Continuum Damage Model which expressed the stiffness of a damaged composite material in

terms of the geometry of the cracks (width, thickness and volume fraction).

In the most common CDM approaches [36], [45], [94], [99], a Finite Element RVE is created in

commercial software, where each layer of a laminate is modelled as homogeneous, with transversely

isotropic elastic properties of individual plies obtained from experimental data. The orientation of the

layers is accounted for in the software by defining local coordinate systems for each layer. Once the meso-

scale model of the laminate has been created, a micro-crack is defined in the layer of interest by

disconnecting nodes along a given surface across the thickness and width of the ply. Using this RVE of a

cracked laminate, the effect of micro-cracking on stiffness can be obtained, without needing to derive

analytical expressions for stress distributions [30], [91]. However, one of the main weaknesses of such

CDM approaches is their inability to make predictions for materials with different elastic properties, with

different crack densities, and different ply thicknesses. Each time a parameter is changed, a new simulation

must be conducted.

Gudmundson and his collaborators [93], [99], [100] proposed a model in which the stiffness of a

laminate, as well as its bending stiffness, was related to the Crack Surface Displacement (Crack Opening

Displacement (COD), and Crack Sliding Displacement (CSD)) of the crack. They developed exact

expressions for stiffness reduction in terms of this parameter. This model was effective at predicting

stiffness degradations at low crack densities, when the stress fields induced by the cracks do not

significantly overlap. However, the method used to obtain the Crack Surface Displacement did not fully

account for the stacking sequence of the laminate. In a realistic application, the stiffness of adjacent layers

as well as their thickness, and the thickness of the cracked layer, will affect the possible opening of the

crack. Therefore, building on the work of Gudmundson and his co-workers, Lundmark and Varna [36],

[101] developed analytical expressions relating the COD of a matrix crack to the properties of adjacent

layers. Varna and his collaborators [102], [103] subsequently developed an expression for the evolution of

Page 38: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

18

COD with increasing crack density, by accounting for the interaction between the cracks in a given layer.

Joffe et al [104] developed an approach to relate the COD of cracks in the 90° layer of a [±θ/904]s laminate

to the elastic properties of the constraining plies. However, the effect of cracking in multiple plies at the

same time was not predicted.

Singh and Talreja [41], [45], [105] conducted a computational implementation of Synergistic Damage

Mechanics (SDM), which combines the strengths of micro-mechanical models of stiffness degradation,

and CDM. Using experimental stiffness degradation data for a cross-ply laminate, they made predictions

of stiffness degradation for laminates with different stacking sequences. Their approach relied on the

calculation of the COD in the different layers of the laminate. By comparing their results to experimental

data, they showed that the COD is a key parameter related to stiffness degradation. They also showed that

stiffness degradation consisted of material-dependent effects, and stacking sequence-dependent effects.

The latter was accounted for through the COD. However, in these works, the interaction between adjacent

micro-cracks, and the effect of this interaction on stiffness degradation was not modelled. Lastly, the effect

of multi-axial in-plane loading scenarios was not investigated.

Montesano and Singh [94], [95], [106] developed FE micro-damage models to predict the effect of

multiple micro-cracking in multi-directional laminates. They took into account the effect of intra-ply crack

interactions, which causes the normalized COD to decrease as the crack density increases. They accounted

for multi-axial loading by defining the COD in terms of an effective strain and by using Periodic Boundary

Conditions in their model. They also showed that cracking in adjacent plies could affect the COD due to

inter-ply interactions. They proved that the results of the SDM model predictions, based on COD evolution,

were in agreement with experimental data, and independent FE calculations. Singh extended the SDM

model [107] to include non-linear relationships between the damage tensor and stiffness degradation.

However, non-linear material properties, in particular viscoelasticity and viscoplasticity were not

accounted for.

The models detailed above were concerned with the prediction of the effect of matrix micro-cracking

on the elastic properties of the laminate. However, in order to predict the overall behaviour of a laminate

under increasing loading, it is also necessary to predict the multiplication of matrix micro-cracking as a

function of applied laminate load. From this information, the full mechanical response of a laminate

undergoing micro-cracking can be calculated.

Two types of models have traditionally been used to predict crack multiplication. In the case of

strength-based theories [31], cracking is assumed to initiate when the transverse ply stress reaches a certain

threshold level. While this type of model is simple, it fails to explain the decrease in crack initiation strain

with increasing ply thickness that has been observed experimentally. Energy-based models, on the other

hand, determine crack initiation and multiplication by calculating the strain energy released upon crack

Page 39: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

19

multiplication. The stress distribution in the cracked ply with increasing crack density is required, from

which the change in strain energy upon matrix micro-cracking can be calculated based on a critical energy

determined experimentally. In these models, cracks are assumed to nucleate and grow throughout the width

and thickness of the ply instantaneously. The concern is to predict crack multiplication, rather than

propagation [108]; the application of energy-based methods to multiple matrix micro-cracking therefore

assumes that cracks form instantaneously following nucleation. In order to make accurate predictions,

however, energy-based models do require a critical strain energy release rate value for each ply, which

depends on the properties of the matrix, the ply and also on the stacking sequence [106]. It is determined

from experimental data.

Nairn et al [56] found that an energy-based model for crack multiplication was better able to predict

crack density evolution in laminates with different stacking sequences than strength-based models. They

used a variational approach to calculate the stress field in the cracked laminates, from which the strain

energy could be calculated for different crack spacings. They defined the critical energy release rate as a

parameter to fit the experimental data of crack density versus applied laminate load. They also measured

the fracture toughness of an individual ply, and found that it was close to the critical energy release rate

for crack multiplication, but not identical. However, this model was limited to cracking in a single 90° ply,

under uniaxial loading. Moreover, the prediction of micro-crack multiplication was based on an

approximate description of the stress field; a FE-based approach would have been more accurate in

calculating this stress field.

Joffe et al [104] defined the thermo-elastic constants of a laminate containing micro-cracks in terms

of the COD. Using these expressions, they were able to develop a model which predicted crack density

while accounting for stochasticity in matrix micro-cracking. Their stochastic model was aimed at

accounting for the experimental observation that the first micro-cracks in a ply nucleate at locations where

pre-existing defects are present, which leads to non-uniform crack spacing at low crack densities. One of

the key contributions of this work was the stochastic description of micro-crack nucleation; however the

model was restricted to uniaxial loading and cracking in a single layer at a time.

Singh and Talreja [109] developed a COD-based model to obtain crack density evolution in laminates

with off-axis plies. They used FEA to develop expressions for COD evolution as a function of crack

density, from which the energy release rate for crack multiplication could be calculated. They obtained the

critical energy release rate by fitting their model to experimental data. The model was restricted to uniaxial

loading. Moreover, a fitting procedure was required to obtain the critical energy release rate for crack

multiplication.

Montesano and Singh [106] extended this previous model by accounting for multi-axial loading, from

which they were able to plot stiffness degradation curves describing the stiffness loss in terms of the applied

Page 40: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

20

bi-axial loads. Moreover, they developed a computational methodology for calculating the critical energy

release rate for crack multiplication, based on the crack initiation strain for each ply of a laminate. Lastly,

by using a periodic RVE to calculate the COD in each layer of the laminate, they were able to predict the

effect of multi-axial loading on cracking in multiple plies at the same time. This type of model will be used

in this thesis due to its high accuracy and large range of applicability

2.4 Models to predict the combined effects of damage, viscoelasticity and viscoplasticity in composites

Viscoelasticity and viscoplasticity in composites were initially comprehensively modelled by

Schapery and his collaborators [77], [110] [24], [27], [32], [111], [112]. Their model requires creep-

recovery data under different loading levels for parameter calibration. Schapery’s model expresses the

time-dependent strain of a composite using the following equation:

휀(𝑡) = 𝑔0𝐴𝜎(𝑡) + 𝑔1∫ 𝛥𝐴(𝜓 − 𝜓′) 𝑑(𝑔2𝜎)

𝑑𝜏𝑑𝜏

𝑡

0− (2.1)

Where 휀(𝑡) is the time-dependent strain, 𝐴 is the instantaneous elastic compliance, 𝜎(𝑡) is the time-

dependent applied stress, and 𝛥𝐴 is a function which represents the increase in compliance with respect to

time. The parameters 𝑔[0−2]are functions of stress and temperature, and account for the change in

compliance with respect to stress. Lastly, the variable 𝜓 is related to the variable representing time, and is

scaled to take non-linear behaviour into account.

Eq. (2.1) is applicable for uniaxial loading scenarios. In the case of unidirectional plies, the transverse

compliance, and the shear compliance will be accurately modelled by an expression such as Eq. (2.1) as

these components are dominated by the mechanical behaviour of the matrix. The parameters will be

different for these two different components of the compliance matrix. In order to account for multi-axial

loading scenarios to be expected in laminates, the approach used by Tuttle et al [24] can be utilized. In

their approach, a non-linear time-dependent compliance model is defined for both the transverse

compliance, and the shear compliance. When a ply undergoes multi-axial stresses, the non-linear

parameters appearing in Eq. (2.1) are defined in terms of the octahedral shear stress.

Guedes et al [113] extended these works by developing an algorithm capable of predicting the

viscoelastic-viscoplastic response of arbitrary laminates under various loading conditions. Zaoutsos et al

[114] used Schapery’s theory to model the viscoelastic response of a carbon fiber/epoxy composite. Al-

Haik et al [115] showed that artificial neural networks could also be used to predict the viscoelastic

behaviour of composites at high temperatures. Korontzis et al [116] developed a procedure to obtain the

linear viscoelastic properties of a laminate from those of a ply, which were obtained experimentally.

The model of Tuttle et al [24] has been used in this thesis to predict the creep response of laminates

because it is an accurate model with a large range of applicability. However, the model has been improved

Page 41: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

21

upon by combining it with a novel multi-scale modelling methodology to couple viscoelasticity and

viscoplasticity with damage evolution.

As explained in section 2.2.3, the viscoelastic and viscoplastic properties of composites have been

shown to affect the long-term strength of laminates. Under a constant load, the viscoelastic and viscoplastic

behaviour of the matrix will cause changes in the structure of the composite by inducing the nucleation of

micro-cracks, which will eventually lead to critical damage modes, such as delamination [26], [83], [88].

In order to predict the long-term mechanical response of laminates, the progression of damage during creep

loading needs to be modelled. A robust tool for designing high-performance composites that considers

structural durability under demanding thermo-mechanical loading can then be developed.

Towards that goal, Park and Schapery [74] developed one of the earliest theoretical models to predict

progressive damage in a linear viscoelastic particulate composite. They derived an expression for

predicting how micro-damage affects the viscoelastic response of the composite. However, the model

required extensive experiments for calibration. Al-Rub et al [84] and Darabi et al [117] combined several

models to account for non-linear viscoelasticity, viscoplasticity and damage in asphaltic materials and

polymer-based composites. A relatively simple procedure was used to predict the effects of damage on the

mechanical performance. Good agreement with experimental data was observed, however, due to the

complex properties being modelled, extensive experimental calibration was required for the different

components of the constitutive model. Marklund et al [27] showed that damage, viscoelasticity and

viscoplasticity could be combined in a relatively simple constitutive law to predict the creep response of a

bio-composite material with a polymer matrix, using a simpler testing sequence where these properties

were individually tested. Although the constitutive model combining damage with viscoelasticity and

viscoplasticity was accurate, the multiplication of damage was not in fact predicted, but instead was

measured experimentally, limiting the usefulness of the model.

Varna et al [103] and Kumar and Talreja [118] developed a CDM model for linearly viscoelastic cross-

plies under static loading. They developed a FE micro-mechanical model combined with a transversely

isotropic linear viscoelastic model for each ply to predict the evolution of the COD under constant strain

and relate it to the stress relaxation induced by matrix micro-cracking. These models provided new insights

into the interaction between damage and viscoelasticity in composites, however, crack multiplication was

not predicted, the analysis was restricted to cross-plies with cracks in a single layer under uniaxial loading,

and the material model was restricted to linear viscoelasticity, which is usually not accurate at high stresses.

Ahci and Talreja [119] proposed a CDM model accounting for non-linear viscoelasticity and damage to

predict the creep response of a cross-ply laminate undergoing matrix micro-cracking, using experimental

data for model calibration. Crack multiplication was not predicted, and the analysis was restricted to cross-

Page 42: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

22

plies. Giannadakis and Varna [120] proposed a model for predicting the effect of shear damage on the

viscoelastic-viscoplastic creep response of a [±45]s laminate, which used experimental data to determine

the combined effects of damage and time-dependent behaviour on the creep response of the laminate.

Although the constitutive relation describing the combined effects of damage, viscoelasticity and

viscoplasticity on total strain was accurate, the model was highly empirical, and could not be easily

extended to other systems. Nguyen and Gamby [89] developed a shear-lag model to investigate the effects

of loading rate on cracking in non-linearly viscoelastic CFRP laminates. They found that the multiplication

of micro-cracks under varying strain rates could be related to the rate-dependent fracture behavior of the

transverse plies. However, the model was not applicable to different stacking sequences with more

complex damage evolution scenarios. The analysis focused on crack multiplication, instead of creep strain

prediction. The experimental study of Asadi and Raghavan [32] showed that under creep loading

conditions, a laminate can undergo time-dependent micro-cracking due to the viscoelastic properties of the

matrix. They developed a CDM model to explain the effect of matrix micro-cracking on the creep response

of the laminate, which showed excellent accuracy but was highly empirical. A previous study by Birur et

al [87] showed similar experimental results. Similarly, Ogi and his collaborators [121]–[123] studied the

effect of matrix micro-cracking in a cross-ply on the stress relaxation response. They were able to predict

the effect of micro-cracks on stress relaxation by using a time-dependent shear-lag model. However, crack

multiplication was not predicted and the analysis was restricted to cross-plies with cracks in the 90º layer.

Asadi [26] and Asadi and Raghavan [124] extended these previous works; they studied the evolution of

crack densities in a [±45/902]s CFRP laminate in all layers under different creep loads, and temperatures.

Asadi [26] developed a micro-mechanical variational model to predict the time-dependent increase in

micro-crack density, as well as the performance degradation due to the defects. However, the analysis was

restricted to a specific stacking sequence and required complex numerical modelling to solve the

differential equations describing the distribution of stresses in a the cracked laminate.

From this literature review, it is clear that although there are some models capable of predicting the

effects of viscoelasticity and viscoplasticity on damage evolution and the ensuing creep response of

laminates, to date, there is no definitive physics-based model that can predict both matrix micro-crack

density evolution and creep strain evolution in multi-directional laminates with cracks in multiple layers

under multi-axial loading scenarios.

2.5 Structural analysis of composite structures under low-velocity impact

Composite materials are being increasingly used in structural applications due to their high mechanical

performance and lower density as compared to traditional materials. In the case of automotive structures,

impact performance is an important property to control to guarantee the safety of passengers, as well as

Page 43: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

23

pedestrians. Novel composite material systems have been successfully manufactured with improved

impact response [125]–[130], which shows that impact performance can be improved for composites by

optimizing the structure. Under low-velocity impact, the evolution of micro-damage mechanisms might

not be noticeable, however their effect on mechanical properties is very significant, which motivates the

need for an accurate model to predict progressive damage. A damage model for impact damage should be

able to relate the laminate material properties, including ply thickness, stacking sequence and material

system to the evolution of damage throughout the structure. Such a model would allow engineers to

improve the low-velocity impact response of automotive components, guarantee the safety of pedestrians,

and help avoid costly vehicle repairs.

To understand the impact response of advanced laminates, experimental approaches have been used

in the past to investigate the effect of different parameters on micro-damage evolution [126]–[134].

Through a proper control of the material properties and geometrical configurations of the laminate,

significant gains in impact performance have been achieved. For example, Sarasini et al [127] found that

damage evolution in flax/carbon fiber cross-plies depended on the presence of flax plies on the outside or

in the middle of the laminate. Ravandi et al [128] studied the effect of inter-ply stitching on micro-crack

propagation. However, while such studies have been extremely useful in improving the understanding of

impact damage, these experimental works are limited in their ability to generalize their findings to different

material systems. Moreover, these studies are expensive, time-consuming and difficult to conduct. The

results are also sensitive to environmental conditions such as humidity and temperature [135], [136].

On the other hand, numerical studies of micro-damage evolution under impact loading have also been

conducted [137]–[142]. A damage model is typically implemented for each ply, which predicts the loss in

stiffness in terms of applied load. A delamination model can also be used to predict structural failure. Most

of these models are based on Continuum Damage Mechanics (CDM) and make use of empirical parameters

to describe the loss in load-carrying capacity with increasing damage. Some of these models have been

applied to car bumpers [143], [144], allowing for improvements in structural design. However, in all cases,

model parameters need to be obtained prior to simulations, which severely limits the range of applicability

of these models, because parameters need to be experimentally determined for each new laminate structure.

CDM models can predict progressive damage under mechanical loading based on experimentally-

obtained damage parameters. However, the Synergistic Damage Mechanics (SDM) model which has been

developed can help by-pass the need for complicated and costly experiments [59], [105], [107], [145].

SDM is based on micromechanical FEA calculations, which calibrate the parameters of a CDM model,

by-passing costly and complicated experiments. A stiffness degradation matrix is defined for symmetric

laminates, and the evolution of Crack Opening Displacement (COD) with respect to crack density is

calculated using an RVE of the laminate with cracks introduced in the plies. From these FE calculations,

Page 44: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

24

the evolution of stiffness with respect to crack density can be calculated without the use of any

experimental measurements. Based on this modelling approach, crack multiplication can also be calculated

using the crack initiation strain in each ply which can be obtained experimentally. The latest developments

of SDM have incorporated multi-axial tensile loading effects [94], [106] by using periodic RVEs of a

cracked laminate. The advantages of this model include its minimal reliance on experimental data for

predictions, accurate modelling of complex multi-axial loading scenarios, and its relative simplicity, which

allows for its implementation in FEA software. Two previous papers have implemented the SDM model

into commercial FEA software to conduct structural analyses which included the effects of progressive

damage [146], [147]. The model was applied to calculate fatigue damage in composite wind turbine blades,

as well as the delamination response. The model has been highly successful in predicting damage evolution

patterns in complex structures, and the progressive failure of the wind turbine leading to full delamination

was predicted accurately.

From this literature review, it is clear that there is a need for an accurate model of rate-dependent

matrix micro-crack evolution at the structural level. Therefore, the model developed in this thesis should

be implemented in FE software as an advanced damage analysis tool that relies on minimal experimental

data, is highly accurate, can be generalized to many materials, and accounts for time-dependent properties.

2.6 Micro-mechanical models to predict damage evolution in short fiber composites

The failure of short fiber polymer-based composites is governed by the evolution of subcritical micro-

damage. The sequence of damage mechanisms, which include fiber-matrix debonding, matrix micro-

cracking and fiber fracture [61], eventually leads to the failure of the material. Due to the mismatch in

elastic properties between the stiff fiber and more compliant matrix, a shear stress concentration occurs at

the fiber tip, which promotes the nucleation of an interfacial crack. This crack may grow along the fiber-

matrix interface, or propagate into the matrix [71], depending on the properties of the interface and matrix.

Experimental studies have shown [61] that once initial damage initiates from the fiber tips, extensive non-

linear deformation occurs in the matrix, which eventually leads to the nucleation of matrix cracks which

propagate throughout the structure. The evolution of these mechanisms is inherently linked to the

microstructure, and to the mechanical properties of the constituents [69].

The process of fiber-matrix debonding has been studied in depth over the past decades both

experimentally and theoretically [148]–[160], due to its fundamental importance in the onset of critical

damage modes in short fiber composites. Computational modelling of fiber-matrix debonding can provide

more detailed insights than experiments into the effects of material properties and reinforcement

configuration on damage evolution in short fiber composites. Initial studies by Tvergaard and his

collaborators [161]–[163] focused on the behavior of metal matrix composites reinforced with ceramic

Page 45: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

25

whiskers. They were able to predict the effects of material properties, fiber dimensions and loading

scenarios on the evolution of damage in the form of debonding and fiber fracture. A review of the current

literature shows many significant works on the modelling of micro-damage mechanisms in short fiber

composites since those seminal papers [29], [156], [164]–[170]. Some of these investigations have

focussed on modelling micro-damage mechanisms explicitly through the development of complex RVEs

which take into account heterogeneous microstructures [165], [166], [171] and use non-linear material

models for the matrix, fiber, and interface. Other models have been developed which do not explicitly

model a composite microstructure, but instead predict the behavior of macroscopic composite structures

using analytical relationships relating damage density to the local stress field throughout the structure. For

example, Notta-Cuvier et al [156] used analytical relationships relating debonded length to the strain

applied to a short fiber in order to predict the behavior of a composite consisting of many fibers. These

models rely on accurate analytical relationships in order to make accurate predictions of composite

behavior. The parameters used in these models need to be obtained from experimental data, which can be

complicated to conduct. Lastly, many computational studies have analyzed fiber-matrix debonding in

unidirectional composites under transverse tensile loading [148], [158], [172]–[175]. However, although

these studies are very useful for unidirectional composites, the evolution of micro-damage in short fiber

composites is inherently different due to the size of the fibers which induces stress concentration at the

fiber tips.

While much work has been done on modelling short fiber composites using computational methods,

the effect of strain rate on the behavior of thermoplastic short fiber composites is still not fully understood.

Some of the previous FE simulations which explicitly modelled the complex microstructure [165], [166],

[171], [176] assumed in-plane deformation, and did not look at rate-dependent effects in detail. Moreover,

accurate relationships between debond length, microcrack density and local strain levels still need to be

developed in order for higher level analytical models to make accurate predictions of structural behavior.

Strain rate behavior is especially important to account for in the case of automotive applications, for which

crashworthiness needs to be understood. Recent developments have also shown increased interest from the

automotive industry in thermoplastic short fiber composites, which are known to exhibit significant rate-

dependent behavior [38], [177]. Several experimental studies [29], [90], [178] have also shown that the

rate-dependent properties of the matrix can affect the kinetics of damage in glass fiber composites, as well

as the nature of the damage mechanisms themselves. However, to date, no definitive computational

description of the effect of matrix viscoelasticity on damage processes at the fiber scale has been

undertaken. While experimental measurements can provide an averaged quantification of damage

processes, for example through acoustic emission analysis [179], computational modelling is required in

order to obtain detailed insights into the progression of damage.

Page 46: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

26

2.7 Gaps in the literature

The literature review of the previous sections has shown that previous research has been conducted on

the evolution of damage in viscoelastic-viscoplastic composite materials. However, some fundamental

gaps in the literature remain, which can be summarized as follows:

No current model can accurately and efficiently predict the time-dependent multiplication of crack

density under constant multi-axial loading in multi-directional viscoelastic-viscoplastic laminates.

Some models have relied on experimental testing for calibration, and their applicability to different

stacking sequences is limited. Other models have relied on complex micro-mechanical modelling

of ply stress evolution during micro-cracking, but their mathematical formulation restricts them to

particular stacking sequences, thereby preventing their applicability to other composites.

The effect of damage on time-dependent creep strain in laminates is not yet clear. Some previous

models have been developed to quantify this effect, but they are empirical and require extensive

testing for parameter calibration. There is a need for a fundamental understanding of the interaction

between time-dependent properties and damage evolution in laminates.

As shown in the previous sections, studies on the evolution of damage mechanisms in short fiber

composites have not accounted for the effect of the viscoelastic and viscoplastic properties of the

matrix, even though it has been established experimentally that rate-dependent loading affects the

nature of damage mechanisms, as well as the evolution of damage. These effects need to be

accounted for in the development of accurate models to understand the mechanical behavior of

automotive composite materials.

Current structural analyses of composite materials rely on empirical relationships to predict the

evolution of damage and its effect on mechanical performance. These models use simplified

expressions of damage evolution which do not fully account for the physics of crack multiplication

under multi-axial loading, and the non-linear relationship between damage and crack density.

Moreover, rate-dependent behavior of the composites arising from the viscoelastic-viscoplastic

properties of the matrix is usually not modelled. These limitations affect the accuracy of structural

analyses. Therefore, there is a need for the implementation of an accurate damage model into FE

software for the analysis of realistic composite structures under dynamic loading.

2.8 Thesis methodology

The SDM model developed by Singh and Talreja [45] and extended by Montesano and Singh [94] has

proven to be a highly versatile and efficient framework for predicting damage evolution in multi-

directional laminates under multi-axial loading scenarios. It has been combined with an energy-based crack

Page 47: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

27

multiplication model [109], [180] to predict crack density evolution with respect to stress. This multi-scale

methodology can easily be applied to many stacking sequences and requires minimal experimental input:

the elastic properties of the plies, and the ply crack initiation strain. Based on these inputs, micro-damage

FE models are generated to calculate the parameters of a Continuum Damage Mechanics model. These

equations can be used to predict stiffness degradation under complex multi-axial loading conditions. In

this thesis, the strengths of the SDM model are further improved by including time-dependent properties

to predict progressive damage evolution under creep deformation. This improvement would provide

further understanding into the nature of the effect of viscoelastic-viscoplastic matrix behavior on damage

evolution in laminates.

The SDM methodology is however restricted to laminates, in which micro-cracks are assumed equally-

spaced, parallel to the continuous fibers, and to span the thickness of the ply and the width of the laminate.

Damage evolution in short fiber composites has been shown experimentally to be fundamentally different

and less regular. Although models are available for these more complex micro-structures, they do not

account for the viscoelastic and viscoplastic properties of the matrix. Considering the strengths of the SDM

methodology, a highly effective approach for understanding the evolution of micro-damage mechanisms

is the development of a micro-mechanical FE model, where damage mechanisms are explicitly taken into

account. Such a methodology would by-pass the need for complicated experimental calibration, and could

efficiently assess the effects of fiber, matrix and interfacial properties on the progression of damage.

Moreover, the interaction between the viscoelastic-viscoplastic properties of the matrix, and the evolution

of damage could be modelled, and provide insights into experimental observations of damage evolution in

polymer-based short fiber composites. This would constitute a key contribution to the field of damage

mechanics of short fiber composites.

Lastly, the SDM model can be implemented into FE software through user-defined subroutines, which

could include rate-dependent behavior. This would extend the applicability of SDM to the structural scale

and would also provide improved insights into the effects of viscoelasticity and viscoplasticity on damage

evolution at the component level under dynamic loading. The application of interest in this thesis would

be low-velocity impact of an automotive component.

In summary, the overall goal of this thesis is the development of a multi-scale framework to predict

damage evolution in viscoelastic-viscoplastic composites, both with continuous fiber and short fibers. This

multi-scale model should span from the microscopic scale of the fiber-matrix interface, to the structural

level of an automotive component, and include the mesoscopic level of the homogenized plies and the

continuum level of a laminate.

Page 48: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

28

Chapter 3. Methodology

In this chapter, the methodology used to address the objectives of the thesis detailed in the previous

chapters are provided. The main equations and concepts are explained, and the numerical implementation

in FE and numerical software is given.

3.1 Prediction of damage in viscoelastic-viscoplastic laminates under multi-axial creep loading

The Synergistic Damage Mechanics (SDM) model is a CDM-based damage model that can quantify

the stiffness degradation due to matrix micro-cracks in the plies of a symmetric laminate. For each ply

orientation, the micro-cracks in a given layer will be parallel to the fibers. A mode of damage corresponds

to a specific crack orientation. Each mode of damage in the different plies of the laminate is described in

terms of a damage tensor. The reduced stiffness due to the matrix micro-cracks can then be calculated with

evolving crack densities in the different plies. In SDM, it is assumed that cracks are parallel to each other,

equally spaced, span the whole thickness of the ply and the whole width of the laminate (Figure 3-1).

Although a given layer may have micro-cracks, it can still carry load because each micro-crack is

constrained by the adjoining plies of the laminate that are oriented at different angles. Moreover, although

it has been observed that micro-cracks might not open completely throughout the width of the laminate,

for the purposes of this thesis, it is assumed that once micro-cracks nucleate, they span the width of the

laminate and the thickness of the ply.

Page 49: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

29

Figure 3-1. Schematic showing the different modes of ply micro-cracking in a multi-directional laminate. VE refers

to the viscoelastic properties of the plies; VP refers to the viscoplastic behaviour; QS is for quasi-static loading. The

top right graph shows the loading scheme used in the current thesis.

3.1.1 Synergistic Damage Mechanics (SDM) model

Each mode of damage (𝛼) corresponds to a different ply crack orientation dictated by the ply

orientation, and the damage in the ply is represented by a damage tensor of the following form:

𝐷𝑖𝑗(𝛼) =

𝜅𝛼𝑡𝛼2

𝑠𝛼𝑡𝑛𝑖𝑛𝑗 = 𝐷𝛼𝑛𝑖𝑛𝑗 (3.1)

where, 𝑡𝛼 is the thickness of the cracked ply, 𝑠𝛼 is the spacing between cracks in the ply, 𝑡 is the thickness

of the laminate, 𝑛𝑖 represents the components of the vector normal to the crack surface relative to the

loading direction, and 𝜅𝛼 accounts for the effect of adjacent cracks on the COD of the ply crack. Due to

the interactions of the stress fields between cracks in different layers, and within the same layer, 𝜅𝛼 will

change depending on the crack density of each ply. The stiffness of a laminate that has undergone

progressive ply cracking can be defined by:

𝐶𝑖𝑗 =

[

𝐸𝑥0

1 − 𝜈𝑥𝑦0 𝜈𝑦𝑥

0

𝜈𝑥𝑦0 𝐸𝑦

0

1 − 𝜈𝑥𝑦0 𝜈𝑦𝑥

0 0

𝜈𝑥𝑦0 𝐸𝑦

0

1 − 𝜈𝑥𝑦0 𝜈𝑦𝑥

0

𝐸𝑦0

1 − 𝜈𝑥𝑦0 𝜈𝑦𝑥

0 0

0 0 𝐺𝑥𝑦0]

−∑𝑏𝛼𝐷𝛼 [

2𝑎1(𝛼) 𝑎4

(𝛼) 0

𝑎4(𝛼) 2𝑎2

(𝛼) 0

0 0 2𝑎3(𝛼)

]

𝛼

(3.2)

Page 50: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

30

𝐶𝑖𝑗 is the 3x3 stiffness tensor of a symmetric laminate written in Voigt notation, under plane stress. 𝐸𝑥0,

𝐸𝑦0 are the Young’s moduli in axial and transverse directions, 𝜈𝑥𝑦

0 , 𝜈𝑦𝑥0 are the major and minor Poisson’s

ratios, and 𝐺𝑥𝑦0 is the shear modulus of the undamaged laminate. The first term is the stiffness of the

undamaged laminate and can be calculated using Classical Laminate Theory (CLT) from the ply properties.

The second term represents the reduction in stiffness of the laminate due to matrix micro-cracks. It depends

on the damage tensor and a set of stiffness degradation parameters 𝑎[1−4]𝛼 , where there are 4 parameters

for each mode of damage 𝛼. The parameter 𝑏𝛼 depends on the stacking sequence of the laminate, and is

equal to 1 for the ply adjacent to the mid-plane of the laminate and 2 for all other plies. The parameters

𝑎[1−4]𝛼 are obtained by computing the decrease in stiffness of the laminate using FE micro-damage model

for a given crack density; they do not change with increasing crack density.

3.1.2 FE micro-damage model

FE micro-damage modelling of a laminate with matrix micro-cracks is used to evaluate the effect of

matrix micro-cracking on the stiffness of the laminate. The individual plies of the laminate are modelled

in FE using a representative volume element (RVE) with transversely isotropic elastic ply properties. The

required input parameters are elastic constants of the plies, their thickness and the stacking sequence of the

laminate. The cracks are represented in the FE mesh by disconnecting nodes across pre-defined crack

surfaces. Using periodic boundary conditions to the RVE, the crack density can be changed by changing

the size of the model. A certain amount of strain is then applied to the laminate, and the Crack Opening

Displacements (COD) of the cracks in the different layers of the laminate are then calculated. For each

mode of damage, the evolution of the COD of each layer with increase in crack density is recorded. As the

crack density increases in a given layer, the COD for cracks in this same layer decreases due to the

interaction between the stress fields of the cracks. Using the results of these simulations, the evolution of

the COD in terms of the crack density in each layer is approximated by a function consisting of three fitting

parameters, in the following form:

𝜅𝛼 = (∆𝑢2̅̅ ̅̅ ̅)𝛼휀𝑒𝑓𝑓𝑡𝛼

=𝑐1

1 + (𝑐2𝜌𝛼)𝑐3

(3.3)

where (∆𝑢2̅̅ ̅̅ ̅)𝛼 is the average crack opening displacement (COD), 𝜅𝛼 is the normalized COD, and 휀𝑒𝑓𝑓 is

the effective strain causing the cracks to open given by 휀𝑒𝑓𝑓 = 휀22 + 𝜈12휀11 +1

8𝜈21𝛾12. 𝑡𝛼 is the thickness

of the cracked ply, 𝜌𝛼 is the crack density corresponding to the mode of damage 𝛼, and 𝑐1, 𝑐2, and 𝑐3 are

Page 51: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

31

fitting parameters. Once the fitting parameters are obtained, a relationship between crack density and COD

can be established, and the COD can be predicted for a known crack density using the above relation. The

construction of the FE micro-damage model has been discussed in further detail by Montesano and Singh

[94]. Lastly, the average COD appearing in Eq. (3.3) is given by the following equation: (∆𝑢2̅̅ ̅̅ ̅)𝛼 =

1

𝑡𝛼∫ 𝛥𝑢𝑦(𝑧)𝑑𝑧𝑡𝛼2

−𝑡𝛼2

, where 𝛥𝑢𝑦(𝑧) is the distance between nodes on each side of the crack.

3.1.3 Energy-based crack multiplication model

The SDM model described in the previous section can predict the stiffness degradation for a given

state of damage or density of matrix micro-cracks. In order to predict the constitutive response of a laminate

undergoing progressive damage, the evolution of the density of matrix microcracks is required. The

evolution of crack density in each layer of the laminate can be evaluated by using an energy-based model

[106]. This model is inspired by Irwin’s theory of fracture mechanics [109], which states that the strain

energy released upon crack propagation is equal to the amount of work required to close the cracks

following their nucleation. This theory has been modified to predict the multiplication of cracks in the

different layers of the laminate [106], [109].

Based on the evolution of COD with the density of matrix micro-cracks obtained using the FE micro-

damage model, the energy density release rate for crack multiplication can be obtained using the following

equation:

𝑊𝐼 = (𝜎2𝛼)2𝑡𝛼𝐸2

[2�̃�𝑛𝛼 (𝑠𝛼2) − �̃�𝑛

𝛼(𝑠𝛼)] (3.4)

Here, �̃�𝑛𝛼(𝑠𝛼) is the normalized COD for a crack spacing 𝑠𝛼 (equal to 𝜅𝛼), 𝐸2 is the modulus of the ply

in the transverse direction, and 𝜎2𝛼 is the local ply transverse stress (perpendicular to the crack surfaces).

𝜎2𝛼 is calculated based on the local ply strain values, and the linear elastic properties of the plies. Using the

strain energy density release rate calculated with Eq. (3.4), a numerical procedure implemented in

MATLAB is used to predict crack density evolution versus applied global strain on the laminate.

In order to predict crack multiplication, the critical strain energy density release rate 𝐺𝐼𝑐 of the ply is

required. 𝐺𝐼𝑐 is itself calculated based on experimental data on ply cracking, combined with a numerical

procedure described in [106]. After the crack density in each layer has been evaluated, the stiffness of the

laminate can be determined by using Eq. (3.1) - (3.3) with the parameters obtained from the FE micro-

damage model. To incorporate different material models into the framework of SDM model, the global

strain in the material should be determined using the material model. Next, the local stress (𝜎2𝛼) can be

computed from the local strain, which is obtained by transforming the global strain to ply coordinate

Page 52: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

32

system. From there, the energy release rate can be calculated using Eq. (3.4), and the crack density can be

updated.

The crack multiplication model is implemented into MATLAB. At the beginning of the numerical

simulations, each ply is divided into a certain number of segments. This number of segments is given by

two parameters, as follows: 𝑁𝑣 =𝐿𝑜

𝑡𝑝𝑙𝑦. 𝐿𝑜is the length of the laminate, which can be set equal to the length

of the laminates that have been studied experimentally in the literature; 𝑡𝑝𝑙𝑦 is the thickness of the ply; 𝑁𝑣

is the maximum number of cracks in the ply. Once 𝑁𝑣 has been computed, a for loop is initiated; at each

iteration of the for loop, the energy release rate for crack multiplication is calculated using Eq. (3.4); this

energy effectively corresponds to the amount of strain energy that would be released if the segment under

consideration underwent cracking. If 𝑊𝐼 > 𝐺𝐼𝑐, the segment is assumed to have cracked; the number of

cracks in the ply is incremented and, conversely, the number of segments available for cracking is

decremented by 1. The for loop continues to the next segments and the number of cracks is incremented

until the crack multiplication criterion is not satisfied anymore. At that point, the for loop is terminated

and the next loading step is applied. If the crack multiplication criterion is always satisfied, the loop stops

after 𝑁𝑣 iterations, at which point the number of cracks has reached its maximum.

As per the work of Montesano and Singh [106], stochastic behavior is taken into account in this damage

model. At each iteration of the for loop described in the previous paragraph, the value of 𝐺𝐼𝑐is in fact

computed using the following equation: 𝐺𝐼𝑐 = 𝐺𝑜 [𝑙𝑛 (1

1−𝐹)]

1

𝑚, where 𝐺𝑜and 𝑚 are Weibull parameters,

and 𝐹 is a randomly generated number between 0 and 1. Therefore, at each iteration of the loop, there will

be slight variations in the values of 𝐺𝐼𝑐, which has been found to better match experimental data [181].

This can be explained based on the random location of defects in the plies such as inhomogeneous fiber

distributions or voids between the fibers. The parameters 𝐺𝑜and 𝑚 can be adjusted to better represent the

distribution of defects in each ply.

It has been observed experimentally [182] that 𝐺𝐼𝑐 increases with increasing crack density. This effect

has been taken into account in this thesis using the following relationship 𝐺𝐼𝑐 = 𝐺𝐼𝑐𝑜 + 𝐺𝐼𝑐𝑟(1 − 𝑒𝑥𝑝(𝑟𝜌𝛼)

where 𝐺𝐼𝑐𝑟 = 0.8 describes the magnitude of the increase in critical energy release rate for crack

multiplication, and 𝑟 = 15 is a material parameter. 𝐺𝐼𝑐𝑜 is the stochastic value of critical energy release

rate obtained using the Weibull distribution.

Page 53: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

33

3.1.4 Viscoelasticity and viscoplasticity models

The creep strain of a viscoelastic-viscoplastic ply can be calculated as a function of time for different

creep loading levels using Schapery’s thermodynamics-based constitutive model [110]. The creep strain

tensor can be defined using the following equation:

휀𝑖(𝑡) = [𝑔0𝑡𝐴0 + 𝑔1

𝑡𝑔2𝑡∑𝐴𝑟[1 − 𝑒

−𝜆𝑟𝑡

𝑎𝜎𝑇𝑡]

𝑘

𝑟 =1

] 𝜎𝑖 + {𝐶𝑖𝜎𝑖𝑁𝑡}

𝑛 (𝑖 = 2,6)

(3.5)

Where 𝑖 = 2 for the transverse strain component, and 𝑖 = 6 for the engineering shear strain component.

The first term in square brackets represents the viscoelastic response of the material. The 𝑔[0−2]𝑡 and 𝑎𝜎𝑇

𝑡

parameters are stress- and temperature-dependent parameters describing the effect of non-linearity on the

compliance. The summation term corresponds to a Prony series consisting of 𝑘 terms, where 𝐴𝑟 is the

increase in time-dependent compliance corresponding to the 𝑟𝑡ℎterm of the Prony series, and 𝜆𝑟is the

inverse of the time-constant describing the rate of increase in compliance with time. The viscoelastic

model, as well as the non-linearizing parameters, are consistent with Schapery’s formulation of non-linear

viscoelasticity [110]. The second term represents viscoplastic strain expressed in the form of a Zapas-

Crissman [183] functional, which is commonly used to describe the viscoplastic behavior of composite

materials. The constants 𝐶, 𝑁 and 𝑛 define the viscoplastic response of the material. The parameters to

describe the response of the composites investigated here were obtained through fits to experimental data,

as explained by Tuttle et al [24].

For orthotropic viscoelastic-viscoplastic plies, the transverse and shear strains are time-dependent,

while the longitudinal strain can be assumed to be linearly elastic due to the large stiffness of the fibers

[24]. In order to evaluate the viscoelastic-viscoplastic behaviour of the laminate, an approach utilizing the

framework of CLT developed by Dillard [85], and Tuttle and Brinson [184] is adopted in this work. At

each creep time-step in the simulation, the local stress in the ply is used to obtain the non-linear parameters

and the local ply compliance using Eq. (3.5), from which the overall laminate non-linear time-dependent

compliance is obtained using CLT. The unconstrained strain in each ply which would occur if the ply were

not constrained by the overall laminate is then obtained, resulting in an equivalent laminate load, which is

used with the laminate compliance to calculate the overall creep strain of the laminate. Local ply stresses

are then adjusted to account for the difference between desired unconstrained ply strain, and actual

laminate strain. The details of the algorithm implemented in MATLAB are provided in the section 3.1.7.

It should be noted that the critical energy density release rate for viscoelastic materials is dependent

on loading rate [89]. The stresses in the plies as well vary slightly over time due to the different viscoelastic

compliance of each layer. Therefore, the critical energy density release rate can be expressed as a function

of time. In order to obtain the time-dependent form of the critical energy release rate, Asadi and Raghavan

Page 54: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

34

[185] performed experiments on individual plies for the same strain rate and at different temperatures, and

recorded the failure strain and stress. Using a model based on thermal activation theory, they obtained the

time-dependent critical energy release rates from their experiments. In order to predict crack density

evolution during creep loading, when the time-dependent properties of the plies will dominate, in this work

the form of 𝐺𝐼𝑐(𝑡) was chosen to best fit available experimental data of crack density evolution versus

time. The values of 𝐺𝐼𝑐(0) for the different plies were obtained using the numerical procedure of

Montesano and Singh [94]. The precise forms of 𝐺𝐼𝑐(𝑡) for each ply for the laminates considered in this

study are given in chapters 4 and 5.

3.1.5 Combined viscoelastic-viscoplastic-damage model

The present thesis is concerned with the progression of damage during creep loading, and the resulting

creep strain. During creep loading, the crack density in the different plies of the laminate will keep

increasing due to the increase in strain. The model calculates the effect of damage on the time-dependent

behavior through the following strain formulation:

휀𝑖𝐿(𝑡) = 𝑆𝑖𝑗

𝐿 (𝐷)𝜎𝐶𝑅𝐸𝐸𝑃,𝑗𝐿 + 휀𝐶𝑅𝐸𝐸𝑃,𝑖

𝐿 (𝑡) (3.6)

In Eq. (3.6), 휀𝑖𝐿(𝑡) is the time-dependent strain tensor that includes the effects of damage (with 3

components as per the plane-stress approximation, and in Voigt notation). The first term is the elastic part

of the strain under a creep load 𝜎𝐶𝑅𝐸𝐸𝑃,𝑗𝐿 and depends on the damage variable 𝐷 that dictates the increase

in laminate compliance 𝑆𝑖𝑗𝐿 due to damage. The increase in laminate compliance is obtained using the SDM

model using Eq. (3.1)-(3.3) [186]. The second term is the viscoelastic-viscoplastic strain as per Eq. (3.5).

During creep loading, the increase in local ply strain will lead to an increase in the crack density and COD.

The increase in crack density during constant loading will increase the state of damage expressed in terms

of the damage variable 𝐷, thereby enhancing the creep response through its effect on the elastic component

of strain as per Eq. (3.6). Therefore, time-dependent properties will affect damage evolution and the total

creep strain in two ways. First, the time-dependent strain will cause an increase in the local ply strain

during loading, which will increase the time-dependent crack density. Second, the time-dependent strain

will directly increase the total creep strain. In this way, therefore, even though damage is assumed not to

affect the purely time-dependent response of the laminate, the time-dependent response of the plies affect

damage evolution.

Page 55: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

35

3.1.6 Numerical implementation

The SDM model combined with a viscoelastic-viscoplastic material model has been implemented in

MATLAB and the computation steps have been demonstrated in the flowchart in Figure 3-2. The

relationship between crack density and COD has been determined from FE micro-damage model

independently. The viscoelastic-viscoplastic material model parameters were determined from the

literature [24], [26]. The simulation was performed in two steps, (a) quasi-static loading and (b) creep

deformation. At each simulation step, the local energy density release rate for crack multiplication was

calculated using Eq. (3.4). The crack density was then updated when the energy release rate (𝑊𝐼) was

found to be higher than the critical energy release rate (𝐺𝐼𝑐) using the methodology described by

Montesano and Singh [106]. Using the SDM model, the stiffness of the laminate can then be updated, and

the stress-strain response predicted. The axial and transverse loads are increased to their respective creep

load levels (𝜎𝑥0 and 𝜎𝑦

0). Following the quasi-static loading, the loads were maintained constant. A similar

procedure was followed for time stepping during the viscoelastic-viscoplastic deformation except that the

time dependent critical energy release rate (𝐺𝐼𝑐(𝑡)) was used to determine the crack multiplication. Based

on the time-dependent crack density, the total creep strain of the laminated can be computed using Eq.

(3.6).

Page 56: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

36

Figure 3-2. Flowchart explaining the MATLAB program used to performed the progressive damage simulations

under viscoelastic-viscoplastic creep.

3.1.7 Detailed implementation of the model into MATLAB

The constitutive model presented in this chapter has been implemented into MATLAB. Here, the steps

implemented into the code are explained. The algorithm is partly based on the method used to calculate

residual thermal strains in a composite undergoing thermal stresses [6]. In this section, the non-linear

viscoelastic-viscoplastic model parameters obtained from the literature are also provided.

3.1.7.1 Quasi-static loading

The elastic properties of a laminate are dependent on the elastic properties of the individual plies, the

thickness of the plies, and the stacking sequence. In the remainder of this thesis, the mechanical response

Apply stress increment dσ to laminate

Get local crack driving stress 𝜎2𝛼using CLT.

Calculate WI (Eq. 3-4) using 𝑢𝑛𝛼 computed from Eq. 3-3. If WI>GIc

increment crack density using the crack multiplication model

Update laminate stiffness matrix (Eq. 3-2). Obtain laminate strain from new

stiffness and stress values.

Apply time increment dt

Get laminate creep strain 휀𝐶𝑅𝐸𝐸𝑃,𝑖𝐿 (𝑡) using material model (Eq. 3-5)

Calculate WI (Eq. 3-4). If WI> GIc (t), increment crack density using the crack

multiplication model

Update elastic part of compliance (Eq. 3-2). Get elastic strain

𝑆𝑖𝑗𝐿𝜎𝐶 𝑁𝑆𝑇.𝑗𝐿 𝑡 and total strain 휀𝑖

𝐿(𝑡)(Eq. 3-6)

No

Yes

Input ply properties and damage model parameters 𝑎 1−4𝛼 , 𝑐1−3, 𝐺𝐼𝑐

Initialize time t=0 and set 휀𝐶𝑅𝐸𝐸𝑃,𝑖𝐿 (0) =0

Yes

Output variables of

interest (e.g. crack

density, strain, stiffness)

No

𝜎𝑥 = 𝜎𝐶 𝑁𝑆𝑇,𝑥 𝜎𝑦 = 𝜎𝐶 𝑁𝑆𝑇,𝑦

T = tend

Page 57: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

37

of a laminate undergoing damage will need to be calculated. The equations used to obtain the elastic

response of a laminate will be given here, and are explained in detail in the textbook by Herakovich [6].

The thickness of each ply in the laminate is given by 𝑡𝑘. Their orientation is denoted 𝜃𝑘. The direction

along the unidirectional fibers is denoted by “1”, while the local transverse direction is denoted by “2”.

The out-of-plane thickness direction is denoted by “3”. In the case of the coordinate system of the laminate,

the loading axis, which is parallel to direction “1” in the ply with orientation 𝜃𝑘 = 0°, is denoted as “x”.

The transverse laminate direction, which is parallel to the direction “1” in the ply with orientation 𝜃𝑘 =

90° is denoted by “y”. The out-of-plane direction, which in this thesis is always parallel to the ply direction

“3”, is denoted by “z”.

To obtain the relationship between stress and strain in a ply, plane stress conditions are assumed, such

that the only non-zero components of the Cauchy stress tensor are the in-plane components. For the

purposes of this thesis, the analysis is restricted to in-plane loading conditions. Therefore, the ply strain

tensor and the ply stress tensor can be expressed as single columns.

In order to calculate the stiffness matrix of a multi-directional laminate, the stiffness tensor of an

individual ply is needed. In the composites studied in this thesis, it is assumed that the ply is transversely

isotropic. The Young’s modulus of the ply in the “1” direction (parallel to fibers) is denoted 𝐸1, while the

Young’s modulus in the “2” direction (perpendicular to fibers) is denoted 𝐸2. The major Poisson’s ratio,

which relates the transverse strain to the applied axial strain, is denoted 𝜈12. The shear modulus is denoted

𝐺12. The relationship between stress and strain for an individual ply is then given by:

[

𝜎1𝜎2𝜏12] =

[

𝐸11 − 𝜈12 𝜈21

𝜈12𝐸21 − 𝜈12 𝜈21

0

𝜈12𝐸21 − 𝜈12 𝜈21

𝐸21 − 𝜈12 𝜈21

0

0 0 𝐺12]

[

휀1휀2𝛾12] (3.7)

Where 𝛾12is the engineering in-plane shear strain, and 𝜏12is the in-plane shear stress. These columns

are expressed in the local ply coordinate system (1-2-3), and the transverse symmetry of the ply prevents

coupling between normal stresses and shear deformation. We also have 𝜈21 = 𝜈12𝐸2

𝐸1 as the minor Poisson’s

ratio. In the equation above, it is assumed that the coordinate system in which the stress and strain tensors

are expressed is aligned with the fibers. The stress values appearing in the stress tensor, and the strain

values appearing in the strain tensor depend on the coordinate system in which these tensors are expressed.

When the coordinate system is rotated relative to the longitudinal axis, the values appearing in these two

vectors will change.

When a ply is oriented at an angle 𝜃𝑘 different from 0° relative to the loading axis, the relationship

between the global stress tensor, [𝜎]𝐺𝐿 𝐵and the global strain tensor, [휀]𝐺𝐿 𝐵 expressed in the coordinate

Page 58: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

38

system of the laminate depends on the angle 𝜃𝑘. In order to obtain the relationship between the global

stress tensor and the global strain tensor, the stiffness matrix of the ply oriented at an angle relative to the

loading direction “x” needs to be modified. To do so, two coordinate transformation matrices are defined

as follows:

𝑇1 = [𝑚2 𝑛2 2𝑚𝑛𝑛2 𝑚2 −2𝑚𝑛−𝑚𝑛 𝑚𝑛 𝑚2 − 𝑛2

]

𝑇2 = [𝑚2 𝑛2 𝑚𝑛𝑛2 𝑚2 −𝑚𝑛−2𝑚𝑛 2𝑚𝑛 𝑚2 − 𝑛2

]

(3.8)

Where 𝑚 = 𝑐𝑜𝑠 (𝜃𝑘) and 𝑛 = 𝑠𝑖𝑛(𝜃𝑘). These matrices relate the global stress tensor and the global strain

tensor to the local stress tensor and the local strain tensor, according to the following relationships:

[

𝜎1𝜎2𝜏12] = 𝑇1 [

𝜎𝑥𝜎𝑦𝜏𝑥𝑦]

[

휀1휀2𝛾12] = 𝑇2 [

휀𝑥휀𝑦𝛾𝑥𝑦]

(3.9)

In order to obtain the stiffness tensor for a ply oriented at angle 𝜃𝑘with respect to the loading axis, the

following equation is used:

Q̅ = 𝑇1−1Q𝑇2 (3.10)

Where Q is the stiffness tensor of the ply in the ply’s coordinate system (see Eq. 3.7), and Q̅ is the ply’s

stiffness expressed in the global coordinate system of the laminate.

Once the global stiffness tensor of each ply has been obtained, the total stiffness of the laminate

sequence can be obtained using the following expression

𝐶𝑝𝑞 =1

𝑡𝑡𝑜𝑡∑ 𝑡𝑘

𝑛𝑝𝑙𝑦

𝑘=1

Qpq̅̅ ̅̅ ̅ (3.11)

Where 𝐶𝑝𝑞is the overall stiffness tensor of the laminate, and 𝑡𝑡𝑜𝑡is the thickness of the laminate. From the

stiffness of the laminate, the elastic response under any in-plane loading scenario can be calculated,

provided the laminate has not undergone any form of damage which would degrade the elastic properties

of the composite.

The first part of the simulations corresponds to quasi-static loading up to the creep load level. During

this portion of the simulations, the viscoelastic and viscoplastic properties of the plies are ignored. This is

Page 59: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

39

reasonable based on the extensive time scales over which time-dependent deformation start becoming

significant. The damage evolution is calculated using the crack multiplication model, and the stiffness

degradation occurring during the quasi-static loading is obtained using the SDM model.

At each loading step, the value of applied stress on the laminate is incremented. Based on the stiffness

of the laminate at that step, the laminate overall strain can be calculated using the following equation:

휀𝑖 = 𝑆𝑖𝑗(𝐷)𝜎𝑗 (3.12)

Where D describes the level of damage in the laminate, and 𝑆 = 𝐶−1 is the compliance matrix of the

laminate. The transverse stress in each ply can then be updated using the elastic properties of the individual

ply:

𝜎2 =𝐸2

1 − 𝜈12𝜈21𝜖2 −

𝜈12𝐸21 − 𝜈12𝜈21

𝜖1 (3.13)

Based on the updated transverse stress of each ply, the energy release rate for cracking in each ply can

be calculated, using Eq. (3.4). The normalized COD values appearing in Eq. (3.4) are obtained from Eq.

(3.3), where the 𝑐[1−3] parameters were derived from the FE micro-damage model.

The crack multiplication model accounts for the stochasticity in ply micro-cracking that has been

reported experimentally [30], and which Montesano and Singh implemented numerically [106]. Moreover,

the increase in critical energy release rate (𝐺𝐼𝑐) with higher crack density has also been accounted for (see

section 3.1.3). At the start of the quasi-static simulations, each ply is divided into a certain number of

segments. At each loading step, a for loop is run through all the intact segments in each ply; the energy

release rate calculated using Eq. (3.4) is compared to the critical energy release rate at each segment. If

𝑊𝐼 > 𝐺𝐼𝑐, the number of cracks is incremented by 1. As soon as this crack multiplication condition is not

satisfied anymore, the crack density in the ply is calculated, and the loop is terminated. This is performed

for each ply.

Once the crack density in each ply has been obtained by using the crack multiplication model, the

stiffness of the damaged laminate is calculated using Eq. (3.1) - (3.3) The laminate strain is then updated

as per Eq. (3.12).

3.1.7.2 Viscoelastic and viscoplastic material properties

The approach used to calculate the creep response of viscoelastic-viscoplastic laminates was taken

from the literature on time-dependent behavior in composites [26], [32]. The first step in the algorithm is

to provide the model parameters, and the viscoelastic-viscoplastic material model, as per Eq. (3.5). The

stress-dependent and temperature-dependent parameters are obtained for a carbon-fiber/bismaleimide

IM7/5260 composite studied by Tuttle et al [24].

Page 60: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

40

Different material systems are characterized by different non-linear viscoelastic behavior. In the results

reported in the next chapter, the exact form of Eq. (3.5) is changed by replacing the Prony series description

of the time-dependent compliance by a Kohlrausch-Williams-Watts (KWW) stretched exponential, so that

the compliance of a ply is given by the following equation:

𝑆𝑖𝑗(𝑡) = 𝑔0𝑆𝑖𝑗0 + 𝑔1𝑔2𝛥𝑆𝑖𝑗 (1 − 𝑒𝑥𝑝 (−[(

𝑡

𝑎𝜎)𝑐

])) (𝑖 = 2,6) (3.14)

The 𝑖 and 𝑗 subscripts denote the component of the strain tensor. In the current work, the Voigt notation is

used, and it is assumed that plane stress conditions apply in the laminate. Therefore, a value of 𝑖 equal to

2 will denote the transverse component of the strain tensor, while a value of 6 denotes the engineering

shear strain component. The other terms in the compliance matrix of the ply are time-independent.

The 𝑔[0−2]parameters are temperature- and stress-dependent non-linearizing parameters which depend

on the material system. For the carbon-fiber/epoxy material studied by Asadi and Raghavan [26], [32],

[185], these non-linearizing parameters are given in [32].

In the case of the GFRP material system studied by Megnis and Varna [187], the viscoelastic-

viscoplastic model is different. The creep strain of a viscoelastic-viscoplastic ply can be calculated as a

function of time for different creep loading levels using Schapery’s thermodynamics-based constitutive

model [110]. The creep strain is defined for GFRP using the following equation [187]:

where 𝑖 = 2 for the transverse strain component, and 𝑖 = 3 for the shear strain component. The first term in

square brackets represents the viscoelastic response of the material. In the material system studied here,

Megnis and Varna [187] showed that the viscoelastic deformation could be assumed to be linear. The

summation term corresponds to a Prony series consisting of 𝑘 terms, where 𝐴𝑟 is the increase in time-

dependent compliance corresponding to the 𝑟𝑡ℎterm of the Prony series, and 𝜏𝑟 is the time-constant

describing the rate of increase in compliance with time. The second term in Eq. (3.15) represents

viscoplastic strain expressed in the form of the product of a stress-dependent function, and a time-

dependent function. The exact expressions for these functions are given by: 휁66 = 5.974 ∙ 10−9𝜏12

2 +

3.419 ∙ 10−8𝜏12 + 3.621 ∙ 10−7 where 𝜏12 is the shear stress in MPa, and 𝜉(𝑡) = 0.003𝑡0.43. We should

note that according to the experimental study, viscoplastic deformation only occurred in the shear direction.

Following the implementation of Eq. (3.15) for each ply, a numerical procedure developed by Dillard et

휀𝑖(𝑡) = [𝐴0 +∑ 𝐴𝑟[1 − 𝑒−𝑡𝜏𝑟]

𝑘

𝑟 =1

] 𝜎𝑖 + 휁𝑘𝑖(𝜎)𝜉(𝑡)𝜎𝑖 (𝑖 = 2,3) (3.15)

Page 61: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

41

al. [85] was used to obtain the creep behavior of the laminate from the creep behavior of the individual

plies.

In order to obtain the creep deformation of the laminate, the following steps are implemented into

MATLAB following the quasi-static simulations, in a series of functions:

1. Increase the value of the time parameter by the time-step value, dt

2. For each ply, calculate the time-dependent components of the compliance matrix at time t, 𝑆22 and

𝑆66. This is done by using the expressions for the non-linearizing parameters, and Eq. (3.5)/Eq.

(3.14)/Eq.(3.15), depending on the material system of interest. The local ply stress tensor [

𝜎1𝜎2𝜏12] is

used to calculate the non-linearizing parameters. The time-independent components of the

compliance matrix are obtained using the linear elastic properties of each ply. The resulting matrix

is denoted by 𝑆𝑖𝑗𝑡

3. For each ply, repeat the procedure described in point 2, but evaluate the compliance matrix at time

t + dt, where dt is the timestep used in the simulations. It is assumed that the local ply stress tensor

is constant over the time step. The resulting matrix is denoted by 𝑆𝑖𝑗𝑡+𝑑𝑡.

4. Calculate the increase in strain that each ply would undergo if it were not constrained by the other

neighboring plies by using the following equation, where 𝛥휀𝑖 is the increase in the local ply strain

tensor:

𝛥휀𝑖 = (𝑆𝑖𝑗𝑡+𝑑𝑡 − 𝑆𝑖𝑗

𝑡 )𝜎𝑗

5. Using the compliance matrix 𝑆𝑖𝑗𝑡 calculated for each ply at time t in step 2, obtain the compliance

matrix of the laminate at time t using the following equations, based on Classical Laminate Plate

Theory [6]:

Qply = St−1

QLAM̅̅ ̅̅ ̅̅ ̅ =1

𝑡𝑡𝑜𝑡∑ 𝑇1

−1Qply𝑇2𝑡𝑘𝑛𝑝𝑙𝑦

𝑆̅𝐿𝐴𝑀 = QLAM̅̅ ̅̅ ̅̅ ̅−1

In the second of these equations, the matrices 𝑇1and 𝑇2 are used to obtain the stiffness of each ply

in the global coordinate system of the laminate. They are provided above.

6. Using the stiffness matrix of each ply obtained in step 5, as well as the unconstrained strain for

each ply obtained in step 4, calculate the equivalent load per unit width necessary to cause such

deformation, using the following equation, where 𝑡𝑝𝑙𝑦 is the thickness of the ply:

Page 62: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

42

𝛥𝑁𝑖 = 𝑄𝑖𝑗𝑝𝑙𝑦̅̅ ̅̅ ̅̅𝛥휀𝑗𝑡𝑝𝑙𝑦

7. Obtain the total load per unit width that would have to be applied to the laminate to cause the

unconstrained deformation of each ply, using the following equation

𝛥𝑁𝐿𝐴𝑀 = ∑ 𝛥𝑁𝑖𝑛𝑝𝑙𝑦

8. Using the laminate compliance obtained in step 5, and the equivalent load increment obtained in

step 7, obtain the increment in strain of the whole laminate, using the following equation:

𝛥휀𝑖𝐿𝐴𝑀̅̅ ̅̅ ̅̅ ̅̅ =

1

𝑡𝑡𝑜𝑡𝑆�̅�𝑗𝐿𝐴𝑀𝛥𝑁𝑗

𝐿𝐴𝑀

9. In order to account for the difference between the strain of the laminate, and the unconstrained

strain of each ply, the local ply stresses must be adjusted, using the following equation, where the

strains are expressed in the local ply coordinate system:

𝜎𝑖𝑡+𝑑𝑡 = 𝑄𝑖𝑗

𝑝𝑙𝑦(휀𝑗𝐿𝐴𝑀 − 𝛥휀𝑗

𝑝𝑙𝑦)

3.1.7.3 Including the effects of damage during creep loading

In order to predict damage evolution during creep loading, the time-dependent crack density and the

ensuing creep strain response must be calculated. To do so, at each increment of time, after the creep strain

of the laminate has been calculated using the approach described in section 3.1.7.2, the crack density in

each ply must be calculated. This is done using the same approach as in section 3.1.7.1 for quasi-static

loading conditions, except that the critical energy release rate for crack multiplication is a decaying

function of time. The precise form of 𝐺𝐼𝑐(𝑡) is given in the next section for each material system studied

herein. Once the crack density for each ply has been calculated, the stiffness matrix of the laminate is

calculated, and the laminate strain adjusted, using Eq. (3.2) and Eq. (3.6).

We should note, however, that the local ply stress appearing in Eq. (3.4) to calculate the energy release

rate for crack multiplication is not equal to the local ply stress calculated in step 9 of the algorithm

described in section 3.1.7.2. Instead, the local ply stress used in Eq. (3.4) is obtained using the following

equation:

𝜎2 = −𝜈12𝐸2

1 − 𝜈12𝜈21휀1 +

𝐸21 − 𝜈12𝜈21

휀2

Where all the constants correspond to the linear elastic properties of each ply. This form of the local

stress allows to better take into account the time-dependent fracture mechanics of the composite plies.

Page 63: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

43

3.1.7.4 Time-dependent critical energy release rate for crack multiplication.

In order to correctly predict crack density evolution during a creep test, the crack multiplication model

must take into account the rate-dependent fracture behavior of the polymeric matrix. According to the work

of Asadi and Raghavan [185], during a creep test, this behavior can be accounted for by using a time-

decaying function to represent the critical energy density release rate for crack multiplication. They

developed an experimental procedure to identify the exact form of this function. This approach is necessary

in order to account for the microscopic structural changes happening in the polymer during loading.

In the present work, a time-decaying function has been used to represent the reduction in critical energy

release rate for crack multiplication with respect to time. In the case of the material system modelled with

Eq. (3.14) for a CFRP material (Toho G30-500/F263-7), the exact form of this function was identified by

comparing the predictions of the crack multiplication model explained in section 3.1.3 to the experimental

data of time-dependent crack multiplication provided by Asadi [26] and choosing the form that provided

the best agreement. It was an approach of trial-and-error. An exponentially decaying formulation for

critical energy release rate is used herein for the 90º ply, i.e. 𝐺𝐼𝑐(𝑡) = 𝐺𝐼𝑐(0) 𝑒𝑥𝑝(−(𝑡 ∙ 10−10)0.105).

In the case of the material system studied by Tuttle et al [24], namely CFRP IM7/5620, no time-

dependent crack multiplication data was available. Therefore, a different approach was used. Nguyen and

Gamby [89] studied the rate-dependent crack multiplication in a carbon fiber/epoxy cross-ply. They

developed a rate-dependent model for the critical energy release rate for crack multiplication. The time-

dependent compliance function that they developed in their paper for this particular material system was

plotted, and found it was found that this function was very close to the behavior of IM7/5620. Therefore,

it was assumed that the time-dependent critical energy release rate for the two materials would be similar.

The rate-dependent critical energy release rate was converted to a time-dependent function by plotting the

compliance as a function of time and taking the derivative of the compliance to obtain the rate of

deformation; from these two curves, a time-dependent critical energy release rate could be obtained.

In order to take the effects of temperature into account, a different form of the time-dependent decay

in critical energy release rate had to be chosen. In the non-linear viscoelasticity model developed by

Schapery, temperature affects the rate of increase in compliance through a temperature-shift factor, 𝑎𝜎𝑇.

Physically, it was assumed in this thesis that the effect of temperature on the time-dependent compliance

would be similar to its effect on the rate of decrease in critical energy release rate. Therefore, the

temperature-shift factor originally developed for the description of the change in compliance with respect

to time was applied to the time-decaying critical energy release rate equation by multiplying it with the

time constant.

Page 64: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

44

In the case of the GFRP material system studied by Varna and Megnis [187], 𝐺𝐼𝑐(0) was determined

by using experimental data published in the literature [188]. Shokrieh et al [189] obtained a rate-dependent

fracture energy for glass fiber composites. The creep strain predicted by the current model was used to

calculate the time-dependent creep strain rate. This time-dependent equation was then combined with the

strain-rate dependent fracture model of Shokrieh et al to obtain the decay in 𝐺𝐼𝑐(𝑡). The final form is given

by:

𝐺𝐼𝑐(𝑡) =𝐺𝐼𝑐0

2+𝐺𝐼𝑐0

2𝑒𝑥𝑝 (−(

𝑡

104))

0.807

(3.16)

where 𝐺𝐼𝑐0 is the critical energy for crack multiplication at the start of creep and 𝑡 is time since the start

of the creep simulation.

It should be noted that in this part of the model, the time-dependent properties of the matrix are

accounted for through the time-dependent GIc(t), and the time-dependent COD evolution. The ply

properties and the local stresses are assumed to be linearly elastic. This is deemed reasonable given that

the stress relaxation caused by the time-dependent behavior of the matrix are very small.

3.1.8 Development of an RVE to evaluate the effect of damage on laminate stiffness

The SDM model relies on a computational micro-damage FE model to evaluate the constants appearing

in the equations relating crack density in each ply of the laminate to its overall stiffness degradation. In

this section, the construction of the RVE used to evaluate the parameters of the SDM model is reviewed.

The FEA simulations performed to obtain the parameters are then described.

The micro-damage RVE can be built in any commercial FEA software. In the present work, ANSYS

mechanical APDL [190] software has been used. An example is shown in Figure 3-3. Each layer of the

laminate is modelled as a transversely isotropic linearly elastic material, with material properties obtained

from experimental data. Each ply is homogeneous and, therefore, the fibers are not explicitly modelled.

The mesh applied to the RVE is constructed such that the separate plies are perfectly bonded (i.e. no

delamination is modelled, and the only mode of damage considered is matrix micro-cracking). The in-

plane boundaries of the RVE are periodic. This is achieved using the approach described by Montesano

and Singh [94]: the mesh on each side of the RVE is set to match the mesh on the opposite side. The

displacement at each node on each boundary is constrained using equation-type constraints to enforce

Periodic Boundary Conditions. The bottom boundary of the RVE is symmetrical, in order to account for

the symmetry of the laminate with respect to the mid-plane.

The orientation of each ply is taken into account by defining a local coordinate system (CS) orientation

relative to the global coordinate system. This local CS is defined by rotating the X and Y axes around the

Z-axis (laminate thickness direction) so that the local X-axis coincides with the longitudinal direction of

Page 65: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

45

the fibers and the local Y-axis is perpendicular to the fibers. The effect of ply orientation is then accounted

for through the anisotropic properties of the plies. The cracks are created by defining crack surfaces before

meshing the RVE. Each crack face is defined for each ply to be parallel to the fibers. The crack surfaces

are defined at the start of model creation for each ply. In the simulations considered in this thesis, cracking

was modelled in one ply at a time. For each simulation performed, cracking was defined in one ply, and

all other ply cracks were closed by merging nodes across the crack surfaces. In the work of Montesano and

Singh [94], cracking was modelled in several plies at once to obtain the interaction between cracks in

separate layers of the laminate. This was not done in this thesis. Lastly, it should be noted that in SDM

and, more broadly, CDM of composites, the RVE is assumed to represent the mechanical behavior of a

laminate. Due to the periodicity of the boundary conditions, inhomogeneity in crack spacing is not

considered.

An example of a deformed RVE and the resulting stress distribution in shown in Figure 3-3(b). The

90° layer contains a single crack which spans the thickness of the ply, and the width of the RVE. The stress

distribution around the crack tip is non-uniform, as expected. In fact, the highly stressed region around the

crack-tips is susceptible to delamination crack initiation, as shown analytically by Carraro et al [191].

However, this mechanism of damage has not been studied in this thesis.

(a) (b)

Figure 3-3. (a) SDM micro-damage FE model for a [0/90/∓45]s quasi-isotropic laminate, with micro-cracks in the 0, 90 and -45

plies and periodic in-plane boundaries. (b) Stress contour (𝜎𝑦𝑦) in the RVE under uniaxial strain (휀𝑥𝑥 = 0.5%, 휀𝑦𝑦 = 0, 𝛾𝑥𝑦 = 0)

with a single crack in the 90° ply.

The RVE is used to obtain the 𝑎[1−4]𝛼 parameters, as well as the COD model parameters, 𝑐1−3. The

𝑎[1−4]𝛼 parameters describe the loss in stiffness, while the 𝑐1−3 parameters describe the interactions between

adjacent cracks, the constraining effects of adjacent plies, as well as between cracks in different layers.

The 𝑎[1−4]𝛼 parameters are obtained by modelling cracking in each ply, and running three separate

simulations, one for each in-plane strain state (휀𝑥𝑥, 휀𝑦𝑦, 𝛾𝑥𝑦). The value of applied strain can be arbitrary,

Page 66: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

46

however in this thesis a value of 0.5% was chosen for all SDM micro-mechanical simulations. From the

stress distribution in the structure for a given amount of applied strain, the average stress and strain can be

calculated using the following equations [94]:

𝜎𝑖𝑗̅̅̅̅ =1

𝑉∫𝜎𝑖𝑗𝑑𝑉𝑣

=1

𝑉∑𝜎𝑖𝑗

𝑛𝑉𝑛

𝑛

휀𝑖𝑗̅̅ ̅ = 1

𝑉∫휀𝑖𝑗𝑑𝑉𝑣

=1

𝑉∑휀𝑖𝑗

𝑛𝑉𝑛

𝑛

Where 𝑛 is the total number of elements, and 𝑉𝑛 is the volume of the element. These summations were

calculated using Ansys ADPL, by defining a script that iterated through each element in the RVE to obtain

the average stress and strain tensors. From the average stress and strain tensor, the plane-stress stiffness

matrix for the damaged RVE can be obtained as follows, using the results from the three separate

simulations corresponding to the three separate values of the applied in-plane strain:

𝑄11𝐿𝐴𝑀 =

𝜎𝑥𝑥휀𝑥𝑥

𝜈21𝐿𝐴𝑀 =

𝜎𝑦𝑦

휀𝑥𝑥

𝑄22𝐿𝐴𝑀 =

𝜎𝑦𝑦

휀𝑦𝑦

𝑄66𝐿𝐴𝑀 =

𝜏𝑥𝑦

𝛾𝑥𝑦

In order to obtain the parameters appearing in Eq. (3.2), the COD corresponding to the crack density

chosen in these simulations needs to be obtained. Based on this value of the COD parameter, the thickness

of the ply, and the crack spacing, the 𝑎[1−4]𝛼 parameters can be chosen such that the values of the stiffness

tensor predicted by Eq. (3.2) correspond to the values calculated from the FEA model.

In order to account for the effect of crack density multiplication on stiffness, the evolution of the COD

needs to be calculated. In this thesis, cracking is modelled in one ply at a time. In the case of the 𝑐1−3

parameters, a set of simulations is run with increasing crack density, with one set of simulations per ply.

Since the RVE is periodic, the crack density can be increased by decreasing the size of the RVE. For each

simulation, the average COD is calculated through post-processing of the ANSYS output data. The

relationship between COD and crack density can then be approximated by the relationship of Eq. (3.3). In

the present thesis, it is assumed that cracks in separate plies do not interact, so that the COD evolution in

each ply is independent of the evolution in adjacent plies. In reality, cracks in separate plies interact, which

휀𝑥𝑥 = 0.005

휀𝑦𝑦 = 0.005

𝛾𝑥𝑦 = 0.005

Page 67: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

47

can be taken into account by defining a COD that depends on cracks in multiple layers, as shown by

Montesano and Singh [94].

In order to account for multi-axial effects, the role of out-of-plane contraction under transverse load

needs to be assessed. In the case of previous papers on the SDM model [45], [105], simulations were

restricted to uniaxial loading. In order to account for the increase in COD with increasing strain, the COD

was normalized (Eq. 3.3) with respect to an effective strain, and the cracked ply thickness. The effective

strain is given by 휀𝑒𝑓𝑓 = 휀22 + 𝜈12휀11 +1

8𝜈21𝛾12 where 𝜈12 and 𝜈21 are the major and minor Poisson’s

ratios of the ply, respectively, and 𝛾12 is the engineering shear strain. To account for multi-axial effects,

simulations are performed with a transverse strain applied to the laminate, in addition to the axial strain.

The evolution of the COD with respect to crack density is recorded. Following this set of simulations, the

COD is normalized with respect to the effective strain, instead of the transverse strain 휀22. The evolution

of normalized COD with respect to crack density is then approximated using Eq. (3.3) by selecting

appropriate values for the 𝑐[1−3] parameters.

In order to improve the efficiency of the SDM model, Montesano et al [192] developed analytical

expressions for the 𝑐[1−3] parameters appearing in Eq. (3.3). The parameters for each ply are based on the

ply’s thickness, as well as the stiffness and thickness of adjacent plies. In this thesis, these expressions

were implemented into MATLAB.

3.1.9 Assumptions of the model and range of applicability

As explained at the beginning of this chapter, the RVE used in the SDM modelling framework makes

certain assumptions about the distribution of damage throughout the laminate:

The cracks are equally spaced throughout the laminate

They are parallel to the fibers

The growth of the micro-crack is not modelled and once it nucleates, it spans the width of

the laminate and thickness of the ply

The presence of defects within plies leads to a distribution of matrix micro-cracks that is initially not

uniform throughout the laminate. However, with increased loading, the average spacing between matrix

micro-cracks becomes approximately uniform. Stochastic effects are considered in this study by defining

the critical energy release rate for crack multiplication (𝐺𝐼𝑐) in terms of a random variable (see section

3.1.3)

The analysis conducted in this thesis has solely focused on symmetric laminates under in-plane

loading. Laminates that are not symmetric will undergo bending and torsion when loaded in tension or

Page 68: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

48

compression. Further analysis would be required to account for these effects, and they are not the focus of

this thesis.

Lastly, the formalism of SDM is applicable to thin laminates, and there is no out-of-plane stress across

the top and bottom boundaries of the laminate. This assumption leads to the stiffness degradation equations

(see Eq. (3.2)). Further improvements on the model would be required to extend it to thick structures.

3.2 Micro-damage model for a short fibre composite

The fourth goal of this thesis is to understand damage evolution mechanisms in short fiber composites

with a ductile thermoplastic matrix. Specifically, the role of strain rate on the process of fiber-matrix

debonding is to be studied. The modelling approach used in this section is based on FE micro-damage

mechanics, in which a Representative Volume Element (RVE) of a short fiber composite is generated. A

Cohesive Zone Model has been used to predict debonding at the fiber-matrix interface. A constitutive

model for a nonlinear viscoelastic matrix has also been implemented.

3.2.1 Geometry

A schematic of the RVE is shown in Figure 3-4. A rectangular box of height L0 and width W0 was

created. Two half fibers were inserted into the RVE, both with diameter Df and length Lf. The fiber ends

were separated by a small distance (1/20 Lf) to model the microstructure of short fiber composites and the

effect of stress concentration at the fiber tips on micro-damage processes. The RVE was meshed with solid

quadrilateral element C3D8R in Abaqus. The bottom faced of the RVE was pinned by preventing

displacement in all directions. A uniform displacement was applied to the top face in the longitudinal

direction. The sides of the box were maintained plane through multi-point constraints, as recommended by

Liang and Tucker [193], in order to appropriately model the mechanical response of the RVE.

3.2.2 Material models

Cohesive zone models define the stress distribution at an interface in terms of separation [162], [163],

[194], [195]. Needleman [194] used CZM to describe debonding in short fiber composites. Ahead of a

crack-tip, a material will undergo changes in its microstructure due to the applied stress. Due to the

evolution of these damage mechanisms, the separation at the interface increases, until the crack grows

along the interface. This behavior is defined through the CZM with a traction-separation law, which relates

the separation at the interface to the stress at the interface. It is a phenomenological model, which does not

account for the micro-damage mechanisms explicitly, but instead accounts for them in an averaged way

through the traction-separation curve. The CZM for a fiber-matrix interface can be obtained by monitoring

the local separation at the interface, and relating it to the applied normal stress. Once the traction-separation

Page 69: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

49

law has been obtained, the behavior of the interface under different loading conditions, and different

interface shapes can be calculated.

The fiber-matrix interface was modelled in Abaqus using a surface-based Cohesive Zone Model, with

parameters of Li et al [172] fitted experimentally to debonding of a glass fiber/epoxy unidirectional

composite. The CZM is based on a linear representation of the interfacial traction-separation behavior. The

initial elastic portion of the CZM represents the traction-separation behavior when the interface has not

undergone damage. The traction across the interface 𝑻 = {

𝑇𝑛𝑇𝑡1𝑇𝑡2

} is related to the separation 𝜹 = {

𝛿𝑛𝛿𝑡1𝛿𝑡2

}

(where 𝑛 denotes the normal direction, and 𝑡1 and 𝑡2 denote the two shear directions) through the following

equation:

It is therefore assumed that the normal and shear interfacial elastic responses are uncoupled. When the

traction across the interface reaches a critical level, damage initiates, which causes a degradation in the

interfacial stiffness matrix. The critical stress levels in the shear and normal directions are provided in

Table 3-2

In this thesis, the maximum stress criterion for damage initiation is used, which states that damage

initiates when any of the traction forces is larger than its critical value (see Table 3-2)

Following damage initiation, the traction at the interface is related to the separation through the

following expressions: 𝑻 = (1 − 𝐷)𝑻, where 𝑻 is the traction calculated in the absence of damage. The

damage variable 𝐷 describing the loss in load transfer capability at the interface is given by the following

expression:

𝐷 =𝛿𝑚𝑓 (𝛿𝑚 𝑚𝑎𝑥 − 𝛿𝑚

0 )

𝛿𝑚𝑚𝑎𝑥(𝛿𝑚

𝑓− 𝛿𝑚

0 ) (3.18)

where 𝛿𝑚𝑚𝑎𝑥 is the maximum separation that has occurred at the interface; 𝛿𝑚

0 is the interfacial separation

at which damage initiates; and 𝛿𝑚𝑓

is the interfacial separation at which full debonding occurs. In this thesis,

the shear and normal directions have a different value for 𝛿𝑚0 . In the case of mixed-mode behavior, where

both normal and shear separations arise at the fiber-matrix interface, the evolution of the damage variable

is described in terms of the effective separation 𝛿𝑒𝑓𝑓 = √(𝛿𝑛2 + 𝛿𝑡1

2 + 𝛿𝑡22 ) .

𝑻 = [

𝐾𝑛𝑛 0 00 𝐾𝑡1𝑡1 00 0 𝐾𝑡2𝑡2

] 𝜹. (3.17)

Page 70: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

50

Figure 3-4 . Model development for the short fiber RVE. (a) Short fiber RVE used in this thesis to study micro-damage

mechanisms in short fiber composites. (b) Top face of the RVE. (c) Side face of the RVE. Image from “Effect of the matrix

behavior on the damage of ethylene-propylene glass fiber reinforced composite subjected to high strain rate tension”, by Fitoussi

J. et al, 2013 Composites Part B, Vol. 45(1), p. 1183. Copyright (2012) by Elsevier. Reprinted with permission.

Figure 3-5. FE model of the short fiber composite.

The two fibers were modelled using the linear elastic properties measured experimentally for a glass

fiber by Li et al [172]. Other types of fibers were considered as well: carbon and hemp. The matrix was

Top fiber

HDPE matrix

Bottom fiber

(a) (b)

(c)

Page 71: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

51

modelled using different mechanical models in different simulation cases to understand the effect of matrix

properties on damage evolution in the RVE. The first material model assumed perfect linear elasticity of

the matrix. In this case the only damage mechanism accounted for was fiber-matrix debonding. In order to

understand the effects of strain rate on the evolution of damage in the short fiber RVE, the viscoelastic

model developed by Popelar et al [196] for HDPE was then implemented into Abaqus to obtain the stress-

strain response of the RVE under different loading rates. Popelar et al [196] used a non-linear viscoelastic

model for HDPE based on a Prony-series description of time-dependent decay in modulus of elasticity.

The constitutive behavior is given by the following expression:

𝜎(𝑡) = ∫ 𝐸𝑅(𝑡 − 𝜏)𝑑[ℎ(휀)휀]

𝑑𝑡𝑑𝑡

𝑡

0

(3.19)

where 𝐸𝑅(𝑡) represents the time-decay in elastic modulus, 휀 is the strain, and ℎ(휀) is a function of strain

representing the non-linear viscoelastic part of the model, given by ℎ(휀) = (1 + 50휀)−1. The time-

dependent modulus is given by:

𝐸𝑅(𝑡) = 𝐸0 +∑𝐸𝑖(1 − 𝑒−𝑡𝜏𝑖)

10

𝑖=1

(3.20)

where the Prony series parameters are given by Popelar et al [196]. Using Eq. (3.19) and Eq. (3.20), an

analytical relationship was derived for the viscoelastic response under constant strain rate, which was then

implemented into Abaqus.

Table 3-1. Elastic properties of the constituents of the RVE.

Fiber and matrix

Glass Fiber Hemp fiber Carbon fiber HDPE Matrix

Young’s modulus (GPa) 64 [172] 8.5[197] 200 [198] 1.86 [196]

Poisson’s ratio 0.2 0.3 0.3 0.42

Table 3-2. Parameters of the Cohesive Zone Model used to describe the fiber-matrix interface.

Direction Critical stress

(MPa)

Critical

separation (μm)

Initial stiffness

(N/mm3)

Normal 7.1 1 107

Tangential 11.0 1 107

A non-linear static analysis was conducted in Abaqus to predict the deformational response of the

RVE. Dynamic effects were ignored, and the rate of deformation of the RVE was assumed to be uniform.

Output from the simulations was written to the database files at regular intervals to obtain sufficiently

detailed information about the deformation and damage processes.

Page 72: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

52

3.2.3 Validation of the implementation

The stress-strain response of the ductile HDPE matrix was implemented into Abaqus in tabular form

by approximating it as a plastic material with isotropic hardening behavior. The implementation was

validated by comparing the results of the model of Popelar et al [196] to the output data provided by

Abaqus. The results are shown in Figure 3-6 (a). Clearly, the model has been correctly implemented.

Subplot (b) shows the 0.2% offset yield strength for HDPE, as a function of strain rate. While inelastic

behavior in polymers is linked to different mechanisms than in metals, this approximation was deemed

reasonable a low strain levels.

(a) (b)

Figure 3-6. (a) Validation of the implementation of the non-linear viscoelasticity model for HDPE into Abaqus in tabular form.

(b) Yield stress versus applied strain rate.

This particular implementation has several limitations. First of all, it is known that the compressive

behavior of HDPE is different from the tensile behavior [199]. This can be taken into account through a

pressure-dependent plasticity model. However, as suggested by Pan and Pelegri [200], the compressive

behavior of the matrix under tensile loading of a short fiber RVE can be neglected with reasonable

accuracy. Secondly, the unloading response of HDPE cannot be accurately predicted with the current

model, since it assumes all non-linear behavior is due to plasticity, which is not necessarily the case for a

viscoelastic material. Since the current thesis is focused on monotonic loading, this approximation is

deemed reasonable and sufficient for the purposes of this work. Lastly, the behavior at large strains is not

accurately taken into account in the current implementation. Following initial non-linear softening, glassy

polymers such as HDPE are known to undergo significant strain hardening [201] at high strains. However,

the current analysis focuses on the low range of polymer deformation, and therefore the accuracy of the

constitutive model is deemed sufficient.

The CZM is used to represent the fiber-matrix debonding process. In order to accurately model this

process, the interfacial stiffness before the initiation of debonding should be large enough that the load is

carried by the composite constituents instead of being artificially carried by the interface. On the other

Page 73: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

53

hand, the interface stiffness should be low enough that the stresses around the interface can be accurately

obtained without spurious oscillations of the stress field in the RVE, and the debonding progression

correctly modelled. The reader is referred to the work of Turon et al [202] for further details on the issues

surrounding the choice of an appropriate value for interface stiffness. In this thesis, the effect of interface

stiffness was evaluated in order to identify reasonable values to accurately model the process of fiber-

matrix debonding. The Young’s modulus of the RVE has been plotted in Figure 3-7 for different values of

the interfacial stiffness. The modulus of the RVE reaches a constant value of 18 GPa for an interfacial

stiffness of 107 N/mm3 or more. At lower values of interfacial stiffness, the modulus of the RVE is under-

predicted. Inspection of the RVE showed large separation at the fiber tip, even without any interface

damage, suggesting that low values of interfacial stiffness do not accurately represent the mechanical

behavior of the short fiber composite. A supplementary simulation was performed where the fiber and

matrix meshes were merged, and a modulus of 18.5 GPa was obtained, thus showing that the modulus

obtained using CZM with high values of interfacial stiffness is correct. An interfacial stiffness of 107

N/mm3 was used in this thesis to accurately model the interface, while still avoiding spurious stress

oscillations in the RVE. Moreover, the mesh was refined around the fiber tips to better account for the

large stress gradients occurring in these regions.

In order to further verify the accuracy of the model, a mesh size sensitivity analysis was conducted.

The results are shown in Figure 3-8 (a) in the form of stress-strain curves under different strain rates.

Clearly, there is little mesh size sensitivity for an element length less than 4 µm. This was therefore chosen

as the element size in the central region of the RVE. A larger element size was used away from the fiber

tips, where the stress gradients are lower. The effect of element size far from the fiber tips was also

evaluated in the case of a non-linear viscoelastic matrix (see subplot (b)). The element size is clearly not

critical on the stress-strain behavior of the RVE with a viscoelastic matrix, and therefore an element size

of 8 µm (away from the fiber tips) was used in subsequent simulations.

Figure 3-7. Effect of interface stiffness on composite RVE modulus

Page 74: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

54

(a) (b)

Figure 3-8. (a) Effect of mesh size on stress-strain response of composite RVE with linear elastic components. The legend refers

to the element size in the central region of the RVE around the fiber tips ; when the mesh size was constant throughout the RVE,

the term ‘uniform’ is used. (b) Effect of mesh size on stress-strain response of composite RVE with a non-linear viscoelastic

matrix under a strain rate of 10-5 s-1.

3.3 Structural analysis of a composite car bumper

The second objective of this thesis is the implementation of the SDM model described in section 3.1

into FE software to conduct a rate-dependent damage analysis of an automotive component. In this section,

low-velocity impact on a bumper is simulated using explicit dynamics FEA [203].

Figure 3-10 shows the general steps of the numerical implementation of the multi-scale damage model.

First, the SDM model parameters are obtained using the approach provided in section 3.1. Next, an FE

model of a composite car bumper is created in Abaqus, and the low velocity impact response of the

structure is calculated using explicit dynamics. The evolution of damage is predicted, and the effect of

strain-rate dependent behavior is evaluated.

Figure 3-10 shows an automotive bumper subjected to low-velocity external impactor. The impactor

is modelled as a rigid hemispherical shell with diameter 198 mm and mass 10.4 kg and is meshed with a

4-node 3-D bilinear quadrilateral (R3D4) element. As the impactor is rigid, a reference point is considered

to represent the geometry of the impactor. The reference point is constrained in 5 degrees of freedom (X

and Y translations and 3 rotations) and it is only free in the impact direction (Z translation). An initial

velocity is applied to the reference point to perform the impact simulation. Two materials, IM7/5260

Carbon Fiber/Bismaleimide (CFRP) and VICOTEX NVE 913/28%/192/EC9756 (GFRP) are considered

for this study; the material properties are listed in Chapter 5.

In order to take rate-dependency into account, the critical energy release rate for crack multiplication,

𝐺𝐼𝑐, is defined in terms of the loading rate, as per the work of Nguyen and Gamby[89]. The following

equation is used:

Page 75: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

55

Where 𝜎�̇� = 1.3216 𝑀𝑃𝑎/𝑚𝑖𝑛 is a reference loading rate, and 𝑚3 = 0.2 describes the rate-dependency of

𝐺𝐼𝑐. 𝐺𝐼𝑐𝑜 is the critical energy for crack multiplication at the beginning of the cracking process and is

provided for each system in Chapters 4 and 5. 𝐺𝐼𝑐𝑟 = 0.8 𝐺𝐼𝑐𝑜and the shape factor 𝑟 = 1.5 describe the

increase in critical energy release rate with respect to increasing crack density (see [180])

The geometrical dimensions of the bumper are given in Figure 3-10. It is meshed with a 4-node doubly

curved shell element with reduced integration formulation (S4R). The transverse shear stiffness values

were equal to 19.8 MN/m for GFRP and 23.24 MN/m for CFRP. These values were obtained using CLT

and the stiffness properties of the plies. The average element aspect ratio is about 2.0 to maintain an

acceptable accuracy in the results. Moreover, as the subroutine is called at every integration point during

the simulation procedure, the element size is considered larger than the micromechanical RVE in order to

ensure the applicability of the constitutive equations. The contact between the impactor and the bumper is

defined via a general explicit contact algorithm in which the contact force is generated by using the penalty

enforcement contact method. Moreover, a tangential interaction with Coulomb friction coefficient is

considered to account for the shear component of the surface traction τ , which is related to the normal

contact pressure by 𝜏 = 𝜇𝑝. The friction coefficient, 𝜇, depends on the surfaces of the contact materials

and has been calculated for various materials. Here 𝜇 is considered to be 0.2 between the rigid surface and

the composite laminate [137]. To implement the constitutive equations and incorporate the damage model,

a vectored user-defined material (VUMAT) subroutine is developed. At each time increment and at every

integration point in the elements, Abaqus calls the VUMAT subroutine to update the state of the material

and the material mechanical response (i.e. stress and energy) based on the strain increment applied at the

integration point. During this process, if the criterion of damage initiation is satisfied (see Eq. (3.4)), the

crack density is updated and the stiffness degradation is calculated to determine the stress increment at that

integration point. This process continues until the end of the simulation. The flowchart of the VUMAT

subroutine is presented in Figure 3-10. Lastly, it should be noted that delamination is not taken into account

in the current model. The focus is on matrix micro-cracking, which will eventually lead to delamination at

high densities.

𝐺𝐼𝑐(�̇�) = 𝐺𝐼𝑐𝑜 + 𝐺𝐼𝑐𝑟(1 − 𝑒𝑥𝑝(−𝑟 𝜌𝛼)) (�̇�

𝜎�̇�)𝑚3

(3.21)

Page 76: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

56

Figure 3-9. Synergistic Damage Mechanics (SDM) RVE, showing the main features of the model, including uniform ply crack

spacing, cracks parallel to the fibers in each layer, cracking in multiple off-axis orientations and multi-axial loading scenarios.

Figure 3-10. General framework for the numerical implementation of the hierarchical multi-scale methodology.

Page 77: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

57

Chapter 4. Development of an SDM-based

viscoelastic creep damage model

In the present chapter, the results of the predictions of the viscoelastic damage developed in the

previous section are provided, and interpreted. First, the model is validated by comparing its predictions

to experimental data obtained from the literature. Second, the predictions of the model for different

stacking sequences and different loading scenarios are explained and discussed. Note that this chapter only

considers uniaxial loading and viscoelasticity; viscoplasticity and multi-axial loading will be considered

in Chapter 5.

4.1 Multidirectional laminates undergoing cracking in multiple plies.1

Before investigating the effects of creep on damage evolution and performance degradation, the quasi-

static response of the [±45/902]s CFRP laminate is predicted and compared to the experimental test data

reported by Asadi [26]. Eq. (3.2) constitutes the basic constitutive model for the damaged laminate, where

the evolution of ply cracking is predicted using the energy-based approach, see Eq. (3.4). The elastic

properties for the carbon/epoxy plies are shown in Table 4-1. The predicted and experimental crack density

evolution data during quasi-static loading are shown in Figure 4-1(a). The ply cracks initiate in the 90°

plies first, as expected, at about 40MPa applied stress, consistent with experimental observations.

Qualitatively, the overall trends in the evolution of crack densities in different plies agree well with the

test data. Quantitatively, it is clear that the model predictions are accurate for the 90° ply cracks, but less

accurate for the ±45° plies at higher applied stresses. In the study by Asadi [26], experimental tests revealed

that delamination was observed at applied stresses of 75MPa. Since this mechanism of damage is not

considered by the prediction model, which only takes into account sub-critical matrix micro-cracking, this

deviation at higher stresses is expected. Nonetheless, the model is accurate at lower applied stresses, and

since the creep simulations will be performed at approximately 50 MPa prior to the onset of delamination,

this is deemed acceptable. The corresponding quasi-static axial stiffness degradation results are shown in

Figure 4-1 (b). Again, the predictions correlate well with the available experimental data, which are well

within the error bars shown in the plot. The sudden changes in stiffness due to the onset of 90° ply cracks

and 45° ply cracks are also clear. Figure 4-1 (c) shows the predicted stress–strain response, which also

1Reprinted from Int. J. Damage Mech., vol. 25, no. 7, T. Berton, J. Montesano and C.V. Singh, Development of a

synergistic damage mechanics model to predict evolution of ply cracking and stiffness changes in multidirectional

composite laminates under creep, pp. 1060-1078, Copyright (2015), with permission from SAGE Publications. DOI:

10.1177/1056789515605569

Page 78: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

58

shows good agreement with the experimental data. Figure 4-1 thus validates the developed model for

multidirectional laminates during quasi-static loading.

Table 4-1. Elastic ply properties for the [±45/902]s CFRP laminate studied by Asadi [26]

Material property Value

E1 (GPa) 113

E2 (GPa) 9.68 G12 (GPa) 4.83

ν12 0.3

Table 4-2. Parameters for the non-linear creep model described by Eq. (3.14), which were obtained using the experimental data

in Asadi [26].

S22 (transverse) S66 (shear)

S0 (MPa-1) 1.033 x 10-4 S0 (MPa-1) 2.07 x 10-4

ΔS (MPa-1) 0.002675 ΔS (MPa-1) 0.002113

τ (s) 3.49 x 1014 τ (s) 1.52 x 109

c 0.2863 c 0.3638

Next, the deformational response of the [±45/902]s CFRP laminate under creep loading is predicted

and compared to the experimental test data. The model parameters for ply creep compliance (see Eq.

(3.14)) are provided in Table 4-2. It is notable that according to previous experimental studies, only the

transverse and shear components of the ply compliance tensor were found to be time-dependent. The non-

linear stress-dependent ply parameters are also obtained from the experimental data reported in Asadi [26]

and are defined as functions of temperature, T, by the expressions provided in the PhD thesis by Asadi.

Page 79: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

59

Figure 4-1. SDM model results and comparison to experimental data [26] for a quasi-static test performed on a [±45/902]s carbon-

fibre/epoxy laminate: a) crack density evolution in different plies; b) stiffness degradation; c) stress-strain curve. The SDM model

used here does not include viscoelastic effects.

0 0.2 0.4 0.6 0.80

20

40

60

80

100

120

140

Strain (percent)

Str

ess

(MP

a)

Exp.

SDM model

0 20 40 60 80 1005

10

15

20

25

Stress (MPa)

Lo

ng

itu

din

al

Mo

du

lus

(GP

a)

Exp.

SDM Model

0 20 40 60 800

0.5

1

1.5

2

Stress (MPa)

Cra

ck d

ensi

ty (

/mm

)

SDM Model (+45° ply)

SDM Model (-45° ply)

SDM Model (90° ply)

Exp. (90° ply)

Exp. (-45° ply)

Exp. (+45° ply)

+45° ply crack

initiation strain

90° ply crack

initiation strain

(b)

(c)

(a)

Page 80: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

60

Figure 4-2. Results of the extended SDM model for creep strain (a) and damage evolution in the 90° ply (b) for a [±45/902]s

laminate subjected to a constant creep stress of 45 MPa, and comparison to experimental data [32]. Note that the initial

experimental creep strain in a) at time t = 0 s was adjusted so that the predicted results and the experimental data coincided at the

onset of creep.

Two simulations were conducted to predict the laminate creep strain evolution, the first without the

effects of ply cracking (i.e. damage) and the second whereby the evolution of cracking was considered.

The results for the laminate creep strain evolution under a constant creep stress of 45 MPa are shown in

Figure 4-2(a) along with the experimental data. It can be seen that the inclusion of damage in the developed

model significantly improves the predicted creep strain evolution, showing that damage is an important

phenomenon affecting creep for these laminates. The corresponding predicted 90° ply crack density

evolution is shown in Figure 4-2(b), along with the experimental data. The initial crack density at a stress

of 45 MPa is accurately predicted at approximately 0.1 cracks/mm. The subsequent crack evolution

prediction also correlates reasonably well with the experimental data and evolves somewhat similarly to

the creep strain due to an apparent proportionality between creep strain and CODs and thus the crack

density evolution. The crack density at longer times is determined to approach a saturation value of 0.5

cracks/mm, which is slightly lower than the experimental data.

0 5 10 15

x 105

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

Time (s)

Cra

ck D

ensi

ty (

/mm

)

Exp.

Extended SDM model

0 2 4 6

x 105

0.22

0.225

0.23

0.235

0.24

0.245

0.25

Time (s)

Str

ain

(%

)

Exp.

Extended SDM model

Creep model without damage

Increase in

creep strain due

to damage (a)

(b)

Lower creep strain

without damage

Page 81: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

61

As depicted above, using the ply elastic and viscoelastic properties along with the critical energy

release rate, the developed model can predict strain and damage evolutions reasonably well. Minor

discrepancies between predictions and test data can be attributed to inherently different nature of damage

processes under quasi-static and viscoelastic conditions, which the current model may not capture fully.

For instance, the crack density is well estimated up until 6x105 s, whereas the crack density is slightly

under-predicted beyond this time range. Nonetheless, the lack of experimental creep data beyond 106 s

precludes a direct comparison with the model predictions. It should be noted, however, that viscoplasticity

may cause supplementary strains at long times, which are not currently accounted for by the prediction

model. Moreover, the model assumes damage only affects the elastic properties of the laminate and not its

viscoelastic time-dependent compliance. As shown by Varna et al., [103] matrix cracks will cause

degradation in time-dependent stiffness, so that accounting for this effect may aid in improving the results.

Future work should consider all these effects to improve the accuracy of the model.

4.2 Cross-ply laminates undergoing cracking in transverse plies.

In order to showcase the versatility of the model, the quasi-static response of a [03/903]s cross-ply

laminate is also evaluated because it has different thickness ratios of cracking and non-cracking plies, and

the test data are from a different source [89]. Although the experimental data revealed that the laminate

exhibited little time-dependent response, it was stated in the study that the crack density evolution under

quasi-static loading was dependent on the applied strain rate. In order to account for the variation in crack

density evolution at different applied strain rates for our predictions, the critical energy release rate GIc was

defined to be a function of the applied strain rate based on data provided in the cited study. The predicted

and experimental results are shown in Figure 4-3, which demonstrates that the current model accurately

predicts crack initiation at lower stresses for lower applied loading rates, which is expected. It can also be

seen that at high loading rates, when time-dependent effects are minimal, the prediction model is very

accurate. As the loading rate is decreased by a factor of 100, the model predictions are less accurate at

higher applied stresses as shown in Figure 4-3. This is most likely due to the fact that time-dependent

effects were neglected for the quasi-static simulations.

Page 82: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

62

Figure 4-3. Predicted crack density evolution for a [03/903]s cross-ply laminate for different levels of strain rate (low, medium

and high) using the SDM methodology, along with experimental data from Nguyen and Gamby [89].

It is expected that at very low applied strain rates, the laminate will exhibit some creep behaviour due

to the increased loading time, which will drive crack multiplication slightly higher. In order to provide

further validation for the current model, our predictions will be compared to the available experimental

results of Ogi and Takao [121] for a [0/903]s CFRP cross-ply laminate. A quasi-static simulation was first

performed using the elastic ply properties presented in Table 4-3, where the predicted and experimental

crack density evolution results are shown in Figure 4-4. A good agreement between the model predictions

and the test data is shown. The accuracy of the results suggests that time-dependent effects were not

important during loading due to the relatively high applied strain rate.

Figure 4-4. Predicted and experimental crack density evolution during quasi-static loading for a [0/903]s cross-ply laminate; the

prediction assumes no viscoelastic effects.

The creep model presented in the previous section is now applied for the [0/903]s CFRP laminate,

where the ply transverse creep properties were determined by fitting the unidirectional ply creep curves

provided in Ogi and Takao [121] with Eq. (3.14) at an applied stress of 510MPa. It was assumed that all

nonlinear parameters were equal to 1 in this instance because there were not sufficient data in the cited

source to determine them precisely. This implies that the nonlinear parameters are stress independent and

thus, the current analysis is only valid for stresses close to 510MPa. Since the magnitude of creep stress

0 500 1000 15000

0.2

0.4

0.6

0.8

1

Stress (MPa)

Cra

ck d

ensi

ty (

/mm

)

High

Low

Medium

Low

Medium

High

SDM

Model

Exp.

0 200 400 600 8000

0.1

0.2

0.3

0.4

0.5

0.6

0.7

Stress (MPa)

Cra

ck d

ensi

ty (

/mm

)

SDM Model

Exp.

Page 83: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

63

we consider is 530 MPa, this is deemed acceptable. The creep parameters obtained through data fitting are

presented in Table 4-3, where a time constant of 319 seconds was used to evaluate GIc(t). It should be noted

that the magnitude of the ΔS22 value presented in Table 4-3 is notably lower compared with the values

shown in Table 4-2, which is a result of forcing the nonlinear parameters to be stress independent. The

predicted and experimental crack density evolution and laminate creep strain results are shown in Figure

4-5(a) and (b), respectively, for a constant applied creep stress of 530MPa. From Figure 4-5(a), it is clear

that the crack density evolution in the 90° ply correlates well with the experimental data when t>2x105 s,

but is slightly under-predicted during the early stages of creep. From Figure 4-5(b), it can be seen that

including the effect of damage evolution on the elastic properties greatly enhances the accuracy of the

predicted laminate creep strain. The laminate creep strain at longer times is slightly underestimated by the

prediction model and is possibly due to neglected viscoplastic strains.

Table 4-3. Parameters used for the creep compliance model, Eq. (3.14), as obtained by fitting the experimental creep

curve for 90° unidirectional plies, at a stress level of 510 MPa [121]

Non-linear Creep

Parameter Value

dS22 (Pa-1) 9.80 X 10-13

τ22 (s) 319

Page 84: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

64

Figure 4-5. a) Predicted and experimental [121] 90° ply crack density evolution for a [0/903]s cross-ply laminate subjected to a

constant stress of 530 MPa and b) corresponding laminate creep strain.

4.3 Predictions for [±60/902]s and [±75/902]s CFRP laminates.

In order to showcase the predictive capabilities of the developed model for general multidirectional

laminates, the results for two additional CFRP laminates, namely [±60/902]s and [±75/902]s, are analyzed

here. It should be noted that no supplementary data were required to make these predictions. The elastic

and time-dependent ply properties were given in Table 4-1 and Table 4-2 and by Eq. (3.14). Two creep

simulations were performed for two different constant stresses of 16 MPa and 19 MPa, respectively, using

the developed MATLAB algorithm (see Chapter 3).

The creep strain results for the two laminates studied are shown in Figure 4-6. Since the [±75/902]s is

more compliant in the loading direction, the initial strain at time t = 0 s for a constant creep stress of 19

MPa is notably higher compared to the [±60/902]s laminate, which is expected. The effect of damage on

the predicted creep strain is also more noticeable for the [±75/902]s laminate – the [±60/902]s laminate did

not exhibit any significant damage evolution (see Figure 4-6). The strain increases by approximately 0.10%

for the [±75/902]s laminate when damage is not considered, whereas the increase is approximately 0.38%

when the effects of damage are considered. This is due to two main factors. First, the ply crack driving

stresses, 𝜎2𝛼 , from Eq. (3.4) are higher for the more compliant [±75/902]s laminate, which undergoes

greater laminate strains in the loading direction. Thus, as the laminate creep strain continues to increase,

0 2 4 6 8 10 12

x 104

1.3

1.4

1.5

1.6

1.7

Time (s)

Str

ain

(%

)

Creep model without damage

Exp.

Extended SDM model

0 1 2 3 4

x 105

0

0.1

0.2

0.3

0.4

0.5

0.6

Time (s)

Cra

ck d

ensi

ty (

/mm

)

Exp.

Extended SDM model

(a)

(b)

Page 85: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

65

the energy release rate WI increases by a greater magnitude, which favours crack multiplication in both the

90° and 75° plies. For the [±60/ 902]s laminate, the laminate creep strain increases by a lower magnitude,

and thus, its effects on crack multiplication in the 90° and 60° plies is minimal. This same effect can also

be seen in Figure 4-2 for the [±45/902]s laminate since the crack density evolution in the 90° ply has a

similar shape to the creep strain curve; it should be noted that the applied loading for the [±45/902]s

laminate was greater at 45MPa. Second, since the critical energy release rate GIc is an exponentially

decreasing function of time, for the [±75/902]s laminate, it directly enhanced the crack multiplication

process due to the adequate laminate strain magnitudes. For the [±60/902]s laminate, the laminate creep

strain magnitudes were not sufficiently high to promote damage even though GIc decreased with time.

Figure 4-6. Predicted creep strain evolution for [±60/902]s and [±75/902]s CFRP laminates, with and without the effects of

damage, subjected to a constant creep stress of 19 MPa.

Figure 4-7 (a), (b) and (c), respectively, show for the [±75/902]s laminate the evolution of compliance

with laminate strain, the evolution of longitudinal stiffness with crack density, and the evolution of crack

density in the 90° ply. From Figure 4-7(a), we can see that there is no notable change in the compliance

before approximately 0.20% strain, which corresponds to the start of the creep portion of the simulation.

The evolution of crack density, and thus elastic compliance, is minimal during the quasi-static portion,

which is due to the relatively small applied stress of 19 MPa. After the onset of creep, the elastic

compliance increases by 0.075 GPa-1 during the duration of creep loading (up to 0.4% strain), which is a

result of notable crack evolution in the 90° ply. The time-dependent creep compliance increases by 0.02

GPa-1 during the same creep interval, which is approximately a quarter of the increase in the elastic

compliance. This shows that the influence of ply crack evolution during creep loading in the studied

laminates is significant and is a result of the specific layup considered. From Figure 4-7(b), we can see that

laminate stiffness, which is affected by the crack density evolution in all plies, decreases linearly during

the initial stages of creep. It should be noted that during creep loading, the ±75° ply crack density is less

0 2 4 6 8 10

x 105

0.2

0.3

0.4

0.5

0.6

0.7

0.8

Time (s)

Str

ain

(%

)

[±75/902]s without damage

[±75/902]s with damage

[±60/902]s with damage

[±60/902]s without damage

Clearer effect of

damage due to

higher compliance

of the laminate

Page 86: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

66

advanced when compared to a quasi-static loading case, as is shown in Figure 4-8. Therefore, stiffness

degradation is greater for the quasi-static case when compared to creep loading (see Figure 4-7(b)).

Figure 4-7. The predicted evolution of compliance is shown here for the [±75/902]s laminate at a creep stress of 19

MPa: a) different components of compliance (creep and elastic), b) stiffness degradation with crack density in the

90° ply, and c) evolution of 90° ply crack density for creep and quasi-static loading.

From Figure 4-7(c), we can see that crack density evolution in the 90° ply during the load ramp-up is

minor, and that it increases significantly during creep loading. The crack density under creep at a specific

strain magnitude is greater when compared to the quasi-static loading case (see Figure 4-7(c)), which is a

result of time-dependent effects in the thicker 90° ply. This is manifested through the critical energy release

rate and ply CODs, which enhance ply crack multiplication. An increase in crack density tends to decrease

0 0.2 0.4 0.64

5

6

7

8

9

10

Crack density (/mm)

La

min

ate

sti

ffn

ess

(GP

a)

Quasi-static loading

Creep loading

0 0.1 0.2 0.3 0.40

0.05

0.1

0.15

0.2

0.25

Total laminate strain (%)

Co

mp

lia

nce

(/G

Pa

)

Total compliance

Elastic compliance

Creep compliance

Onset of

creep

Significant

cracking

during creep (a)

(b)

Onset of

creep

0 0.1 0.2 0.3 0.40

0.1

0.2

0.3

0.4

0.5

0.6

0.7

Total laminate strain (%)

Cra

ck d

ensi

ty (

/mm

)

Quasi-static loading

Creep loading

Onset of

creep

(c)

Ramp-up for

creep

Page 87: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

67

COD for elastic laminates [94]; however, in the 90° ply the COD increases (see Figure 4-9), which is a

result of increasing creep strain. This demonstrates that the time-dependent effect on the 90° ply COD is

notable for the laminates studied, which is an important result in this study. Furthermore, it can be seen

from Figure 4-8 that the crack density begins to plateau at higher strains; however, the characteristic

damage state (CDS) is not reached before the end of the creep test. This plateau is due to the fact that the

creep strain also reaches a plateau after a notable time increment, which limits the crack driving stresses

during the latter stages of creep loading.

Figure 4-8. Predicted crack density evolution in a [±75/902]s laminate in the angle plies, under creep at 19 MPa and quasi-static

loading. The figure shows that cracks evolve more slowly during creep than during quasi-static loading at low strain levels.

Figure 4-10(a) shows the evolution of laminate axial compliance for two different stress levels during

creep loading of a [±75/902]s laminate. It can be seen that there is a small increase in the creep compliance

component when the stress is increased from 16 MPa to 19 MPa. This is due to the nonlinearity in the

KWW creep model (see Eq. (3.14)), which accounts for a higher compliance at higher stresses through the

nonlinear parameters defined in Eq. (3.14). On the other hand, the elastic compliance at the 16 MPa applied

stress increases from 0.1 GPa-1 to 0.2 GPa-1 during creep, whereas it increases from 0.1GPa-1 to 0.3 GPa-1

at the 19 MPa applied stress. This is due to enhanced crack evolution at the 19 MPa applied stress level,

leading to a greater effect on stiffness degradation. Once again it is demonstrated that ply crack evolution

directly influences the long-term laminate response. It can be inferred from Figure 4-10 that ply cracking

leads to a significant increase in total laminate compliance, even over narrow stress ranges. Figure 4-10(b)

shows the different components of the compliance (creep and elastic) for the [±60/902]s and [±75/902]s

laminates at the same stress level of 19 MPa. The creep compliance of the [±60/902]s laminate is

approximately half that of the [±75/902]s laminate. This is due to the difference in the stacking sequence

and points to the importance of an accurate creep model that considers deformation in both the shear and

transverse directions. The elastic compliance is also much lower for the [±60/902]s and evolves very little

due to the decreased amount of ply cracking, which is a result of lower energy release rates caused by

lower laminate creep strains.

0 0.1 0.2 0.3 0.40

0.5

1

1.5

2

Strain (%)

Cra

ck d

ensi

ty (

/mm

)

Quasi-static loading, +75° ply

Quasi-static loading, -75° ply

Creep loading, -75° ply

Creep loading, 75° ply

Page 88: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

68

Figure 4-9. Predicted COD evolution for a 90° ply in a [±75/902]s laminate undergoing creep at 19 MPa, showing a clear increase

linked to creep strain evolution. The increased in COD affects crack density evolution through its effect on the energy release rate

WI.

Figure 4-10. a) Plot showing the predicted evolution of the elastic and time-dependent (creep) components of the

[±75/902]s laminate compliance, at two different stress levels, demonstrating the impact of damage on compliance

increase. b) Plot showing the predicted evolution of the different components (creep and elastic) of the compliance

for the two different laminates, at a stress of 19 MPa.

0 2 4 6 8 10

x 105

0.8

0.9

1

1.1

1.2x 10

-6

Time (s)

CO

D (

mu

m)

0 2 4 6 8 10

x 105

0

0.1

0.2

0.3

0.4

0.5

0.6

Time (s)

Com

pli

an

ce (

/GP

a)

Creep

Elastic

Total ([±60/902]s)

Total ([±75/902]s)

Elastic

Creep

0 2 4 6 8 10

x 105

0

0.1

0.2

0.3

0.4

Time (s)

Com

pli

an

ce (

/GP

a)

Total (19 MPa)

Creep

Elastic

Total (16 MPa)

Elastic

Creep

Change in creep

compliance

Significant

increase in elastic

compliance

(a)

(b)

Page 89: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

69

Summary for Chapter 4

This chapter presented a physics-based multi-scale model to predict the simultaneous ply crack

evolution and time-dependent deformational response of general multidirectional laminates subjected to

quasi-static and creep loads. The model follows the framework of SDM and utilizes computational

micromechanics to define the corresponding damaged laminate stiffness degradation parameters, requiring

only ply-level elastic and viscoelastic empirical data as input. The critical energy release rates associated

with ply crack multiplication are also evaluated computationally through a crack closure concept. A

nonlinear viscoelastic Schapery-type model is used to represent the ply behaviour, which is cast into the

framework of CLT for predicting the total laminate time-dependent response. The effects of time-

dependent behaviour and damage evolution are superimposed to evaluate the laminate compliance and the

corresponding creep strains for various cross-ply and multidirectional laminates. Model predictions for the

studied laminates subjected to quasi-static and creep loading correlated well with available experimental

data, with respect to the evolution of ply crack density and laminate creep strain. The predictive capabilities

of the model were further shown through a parametric evaluation of [±θ/902]s laminates subjected to creep

loading at various applied stresses. The developed model is the first reported model that considers the

effects of viscoelasticity, ply crack evolution and stiffness degradation for general multidirectional

laminates simultaneously and will provide an effective means to assess the long-term durability of

structures made from these laminates.

The model presented in this chapter is applicable to uniaxial loading scenarios and the ensuing

evolution in micro-crack density in the different plies of multi-directional laminates. The model is has been

developed for symmetric laminates specifically, with a focus on in-plane mechanical properties. Constant

loading was the focus of this chapter; future works should consider the effects of different loading

scenarios. In the next chapter, the effects of viscoplasticity, as well as bi-axial loading scenarios, are

considered.

Page 90: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

70

Chapter 5. Time-dependent bi-axial stiffness

degradation envelopes for structural

composites2

In this chapter, the SDM model is extended to include viscoplastic behavior of the plies. The effect of

bi-axial loading scenarios on damage evolution in symmetric laminates is studied by using COD-based

descriptions of progressive damage. Carbon fiber and glass fiber composites are studied. The chapter

concludes with the development of stiffness degradation envelopes which describe loss in performance

with respect to time for different biaxial loading scenarios.

5.1 Model validation

In order to demonstrate the accuracy of the current model, and its ability to predict the time-dependent

behavior of viscoelastic-viscoplastic composites, the model has been validated with respect to

experimental data, as well as previous models. Due to the current unavailability of experimental data for

time-dependent matrix micro-crack multiplication in viscoelastic-viscoplastic laminates, the different

components of the model have been validated separately. The three parts of the model were presented in

Chapter 3, and consist of a viscoelastic-viscoplastic material model for laminates (not including the effects

of damage), a time-dependent micro-crack multiplication model, and a stiffness degradation model based

on Synergistic Damage Mechanics. While the combined effects of viscoelasticity, viscoplasticity and

damage have not been studied experimentally in laminates in a single procedure, the validation of the

separate parts of the model is sufficient for the purposes of this thesis given the significant quality of

research that has been conducted in the past on the separate issues of matrix micro-cracking and

viscoelastic-viscoplastic behavior in laminates.

2Reprinted from Compos. Struct., vol. 203, March, T. Berton, S. Haldar, J. Montesano and C.V. Singh, Time-

dependent damage analysis for viscoelastic-viscoplastic structural laminates under biaxial loading, pp. 60-70,

Copyright (2018), with permission from Elsevier. DOI: 10.1016/j.compstruct.2018.06.117

Page 91: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

71

Figure 5-1. Schematic showing the different modes of ply micro-cracking in a multi-directional laminate. VE refers to the

viscoelastic properties of the plies; VP refers to the viscoplastic behaviour; QS is for quasi-static loading. The top right graph

shows the loading scheme used in the current chapter.

5.1.1 Validation of the viscoelastic-viscoplastic implementation

We first validate the nonlinear viscoelastic-viscoplastic model, without any damage, by comparing the

predictions from our implementation to the theoretical creep simulation results for a GFRP laminate under

uniaxial loading previously published by Megnis and Varna [187]. The parameters to describe the

viscoelastic-viscoplastic behavior of the individual plies of the GFRP laminate as defined in Eq. (3.15) are

given in Table 5-1. Figure 5-3(a) shows the prediction of creep strain following a quasi-static loading to a

stress of 50 MPa in the [±45]2s GFRP laminate and comparison with the material model previously

developed by Megnis and Varna for viscoelastic-viscoplastic plies [187]. It can be observed from the figure

that the predictions of the current model match very well with the previous work, thus validating the creep

model.

Page 92: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

72

Figure 5-2. Flowchart explaining the MATLAB program used to perform the progressive damage simulations under

viscoelastic-viscoplastic creep.

Table 5-1. Parameters for the viscoelastic part of the model [187]

A1 (Pa-1) A2 (Pa-1) A3 (Pa-1) A4 (Pa-1)

Transverse

Shear

𝜏𝑟(s)

8.12 x 10-12

3.24 x 10-11

2400

2.81 x 10-12

-3.09 x 10-11

14000

6.14 x 10-12

4.64 x 10-11

25000

8.25 x 10-12

2.29 x 10-11

550000

Table 5-2. Elastic properties of the GFRP plies, obtained from [187].

Elastic property Value

E1(GPa)

E2(GPa)

𝜈12 G12 (GPa)

45

14.6

0.32

4.95

Apply stress increment dσ to laminate

Get local crack driving stress 𝜎2𝛼using CLT.

Calculate WI (Eq. 3-4) using 𝑢𝑛𝛼 computed from Eq. 3-3. If WI>GIc

increment crack density using the crack multiplication model

Update laminate stiffness matrix (Eq. 3-2). Obtain laminate strain from new

stiffness and stress values.

Apply time increment dt

Get laminate creep strain 휀𝐶𝑅𝐸𝐸𝑃,𝑖𝐿 (𝑡) using material model (Eq. 3-5)

Calculate WI (Eq. 3-4). If WI> GIc (t), increment crack density using the crack

multiplication model

Update elastic part of compliance (Eq. 3-2). Get elastic strain

𝑆𝑖𝑗𝐿𝜎𝐶 𝑁𝑆𝑇.𝑗𝐿 𝑡 and total strain 휀𝑖

𝐿(𝑡)(Eq. 3-6)

No

Yes

Input ply properties and damage model parameters 𝑎 1−4𝛼 , 𝑐1−3, 𝐺𝐼𝑐

Initialize time t=0 and set 휀𝐶𝑅𝐸𝐸𝑃,𝑖𝐿 (0) =0

Yes

Output variables of

interest (e.g. crack

density, strain, stiffness)

No

𝜎𝑥 = 𝜎𝐶 𝑁𝑆𝑇,𝑥 𝜎𝑦 = 𝜎𝐶 𝑁𝑆𝑇,𝑦

T = tend

Page 93: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

73

Table 5-3. Damage model parameters for the glass fiber/epoxy composite

Laminate [0/90/∓45]s [0/90]s

Ply orientation 0° 90° -45° 45° 0° 90°

Stiffness

degradation

parameter

(GPa)

a1 0.77 8.08 5 5 0.77 8.14

a2 7.45 0.84 4.92 5.01 7.55 0.82

a3 1.74 1.54 0.90 0.84 1.30 1.61

a4 4.78 5.37 3.48 2.56 6.64 3.07

Critical

energy

release rate 𝐺𝐼𝑐0 300 300 300 300 300 300

COD

evolution

parameters

c1 3.15 1.28 1.28 1.26 3.15 1.26

c2 0.95 0.35 0.35 0.7 0.95 0.7

c3 1.69 1.63 1.63 1.61 1.69 1.61

5.1.2 Validation of the crack multiplication model

The crack multiplication was validated by comparing its predictions to experimental results provided

in the literature for the glass fiber/epoxy material system studied in this chapter [188]. The predictions of

the crack multiplication model are compared to the experimental data in Figure 5-3(b). The critical energy

release rates for crack multiplication in each ply were obtained through trial-and-error, until agreement

with experiments was satisfactory. The values 𝐺𝐼𝑐0 for each laminate and each ply are given in Table 5-3.

In the case of the simulations performed in this chapter, it was necessary to predict the evolution of

cracking during constant creep loading. As explained in Chapter 3, this was accomplished by implementing

a time-dependent critical crack multiplication energy, 𝐺𝐼𝑐(𝑡). Using this time-dependent critical energy,

as well as the creep model for viscoelastic-viscoplastic composites, crack multiplication was predicted for

a [±45/902]s CFRP laminate for which time-dependent micro-cracking has been studied experimentally by

Asadi [26]. The results of the model have been compared to the experimental measurements in Figure

5-3(c), showing almost perfect agreement.

5.1.3 Validation of the SDM model

The stiffness degradation parameters used for the SDM model, namely 𝑎1−4𝛼 , were obtained for the

GFRP laminates using FEA micro-mechanical simulations and are provided in Table 5-3.The parameters

for the evolution of COD with respect to crack density are provided in Table 5-3 as well. In order to validate

the SDM model for the GFRP, independent FEA simulations were performed for different crack densities

from which stiffness was obtained. The predictions of the independent FEA simulations were compared to

the predictions of the SDM model, as shown in Figure 5-3(d). Clearly, the predictions of the SDM model

Page 94: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

74

are extremely close to the independent simulation results, thus validating the analytical COD formulation,

and the 𝑎[1−4 ]𝛼 parameters.

5.2 Model predictions

5.2.1 GFRP

Two laminates with different stacking sequences and the same ply material properties were considered in

this work to investigate the effect of ply lay-up. A detailed parametric study has been performed to predict

the crack density and stiffness degradation in the [0/90/∓45]s and [0/90]s glass fiber/epoxy laminates with

viscoelastic-viscoplastic properties described by Eq. (3.15) and material properties obtained from Megnis

and Varna [187]. The parameters for the energy-based crack multiplication model and the SDM model are

given in Table 5-3. The ply thickness was set to 0.5 mm.

Figure 5-4(a) shows the evolution of crack density in the 90° ply of a [0/90/∓45]s GFRP quasi-isotropic

viscoelastic-viscoplastic laminate during quasi-static and under creep deformation at a constant axial load

of 210 MPa. The different curves correspond to different transverse loads. The initial portion of the curves

to the left of the red vertical line corresponds to the quasi-static loading, while the portion to the right of

the red line corresponds to the creep loading. It can be observed from the figure that cracking initiates

during the quasi-static loading, and the crack density continues to increase during the creep test. The crack

density evolution shown in Figure 5-4(a) can also be qualitatively compared to experimental measurements

of crack density evolution during creep loading (see [32]). From Figure 5-4(a), it can be seen that an

increasing applied transverse stress 𝜎𝑦 does not lead to a large difference in crack density evolution in the

90° layer. The crack density in the 90° ply reaches approximately 1.5 cr/mm at long times under all

transverse loads. The evolution of crack density in the +45° ply is shown in Figure 5-4(b). The overall

trends of the evolution of crack density are similar to that in the 90° ply. However, transverse loading has

a large effect on the crack density in the off-axis ply. This effect is due to the orientation of the plies for

which a transverse laminate stress causes a larger local transverse stress driving crack multiplication. The

crack density is equal to 0.6 cr/mm under uniaxial loading in the 45° ply, while it reaches 1.4 cr/mm under

a transverse load of 180 MPa.

To quantify the increase in crack density during creep under biaxial loading, crack densities before

(t=0) and after (t = tend) the viscoelastic-viscoplastic deformation are shown in Figure 5-5 for the 90° and -

45° plies of the [0/90/∓45]s quasi-isotropic laminate. In the case of the 90° ply, the crack density increases

with the applied axial load 𝜎𝑥, with crack initiation occurring at around 140 MPa, as can be seen in subplot

(a). The effect of creep strain is to lower the stress necessary to reach a given crack density, shifting the

crack initiation stress to around 80 MPa, as can be seen in subplot (a). The transverse load has very little

Page 95: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

75

effect on the crack density evolution, both before and after creep loading. This can be explained from the

stacking sequence of this laminate. In the case of the 90° ply of the quasi-isotropic laminate, the transverse

load causes a slight contraction of the laminate in the axial direction, preventing cracking. However, the

out-of-plane contraction enhances crack multiplication. In this particular stacking sequence, these effects

cancel out, such that the transverse load has almost no effect on crack density evolution in the 90° ply. In

the case of the -45° ply, shown in subplot (b), the crack density also increases with increasing axial stress.

The transverse stress has a very large effect on the crack initiation stress, however, reducing it from more

than 240 MPa to about 80 MPa at time t=0, and from 150 MPa to less than 0 MPa at time t = tend as seen

in subplot (b). This can be explained from the fact that the transverse stress causes an increase in the local

ply COD and local transverse stress by increasing the local ply transverse strain. After creep loading has

occurred, the crack density in the -45° ply has significantly increased.

(a) (b)

(c) (d)

Figure 5-3. a) Viscoelastic-viscoplastic creep strain prediction for a [±45]s GFRP laminate under a creep stress of 50

MPa, and comparison to a previous model from the literature [187]. b) Evolution of crack density under quasi-static

loading for a cross-ply GFRP composite, and comparison to experimental data [188] c) Time-dependent crack density

evolution in each layer of a [±45/902]s CFRP laminate predicted by the model, and comparison to experimental data

under a creep stress of 45 MPa [26]. (d) Predictions of stiffness degradation with respect to crack density in a

[0/90/∓45]s GFPR laminate using the SDM model (Eqs. 3.1-3.3) and independent FEA simulations in APDL

Page 96: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

76

(a) (b)

Figure 5-4. Crack density evolutions versus simulation time for a [0/90/∓45]s GFRP laminate loaded to an axial

load of 210 MPa. a) 90° ply crack density. b) 45° ply crack density.

(a) (b)

Figure 5-5. Crack density versus axial stress with different levels of transverse stress in the 90° and 45° plies of the

quasi-isotropic GFRP [0/90/∓45]s CFRP laminate before (t=0) and after (t= tend) creep deformation.

To show the effect of biaxial load ratio on crack density evolution and stiffness, the evolution of crack

density in the different plies of the [0/90/∓45]s laminate and its axial and transverse stiffness with respect

to the biaxial load ratio are shown in Figure 5-6(a). The results correspond to an axial load 𝜎𝑥 level of 150

MPa and a varying transverse load 𝜎𝑦 of 0 – 300 MPa. It can be seen for the quasi-isotropic laminate that

the crack density in 45o and -45o plies are most affected by the transverse load. While the crack density in

90o ply increases from 0.9 to 1.15 cr/mm, the crack density in -45o ply increases from around 0 cr/mm to

more than 1.7 cr/mm. As expected, the increase in transverse load significantly affects the transverse

stiffness as can be seen in the figure. It also affects the axial stiffness significantly, leading to a decrease

in modulus from 0.91 to 0.7 of the initial value. In the case of the cross-ply laminate shown in subplot (b),

crack density in both layers increases due to transverse load, leading to more severe stiffness degradation

in both laminate directions. The predictions of crack density evolution up to around 0.8 cr/mm and

Page 97: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

77

associated stiffness degradation by 20 % in these laminates are in good agreement with experimental

observations [54]. We should note that the results of Figure 5-6 are taken at the end of the creep

simulations.

The axial stiffness due to damage progression under quasi-static loading (i.e. before creep) and that

under creep is depicted in Figure 5-7 for the two GFRP laminates. The solid lines represent the stiffness

of the laminates after quasi-static load, i.e. before creep, and the dashed lines represent the stiffness of the

laminates after creep. As can be seen in Figure 5-7(a), the initial axial stiffness of the undamaged

[0/90/∓45]s CFRP starts degrading beyond 140 MPa of uniaxial quasi-static loading and keeps decreasing

with increasing applied load. The normalized stiffness reduces to a value of 85 % of the undamaged value

under a uniaxial load of 300 MPa. Low transverse loads do not have a noticeable effect on damage

initiation; however, when the transverse load reaches 180 MPa, damage initiates at an axial load of 80

MPa. Under transverse loading, cracking is enhanced in all layers of the laminate, causing a larger loss in

stiffness with axial loading. Under a transverse load of 180 MPa, the stiffness has reduced to 77 % of the

undamaged value. Creep deformation enhances cracking, and causes further stiffness degradation. Figure

5-7 (b) shows the axial stiffness of the [0/90]s CFRP laminate under quasi-static loading (before creep) and

after creep. Under uniaxial loading, the stiffness starts degrading under a load of 150 MPa and keeps

decreasing with higher loading levels, due to increased cracking. At the end of 300 MPa of quasi-static

loading the axial stiffness reduces to a value of 85 % of the initial value. The presence of transverse stress

enhances the damage progression leading to a larger reduction in stiffness as can be seen by comparing the

solid lines of different colors. Similarly, creep deformation causes further stiffness degradation, with the

stiffness reaching a minimum value of 79%, at which point crack density is saturated.

(a) (b)

Figure 5-6. Evolution of crack density and normalized stiffness for the GFRP composite under different levels of

transverse loading: (a) Quasi-isotropic [0/90/∓45]s laminate under an axial load of 150 MPa, and (b) cross-ply

laminate under an axial load of 150 MPa

Page 98: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

78

(a) (b)

Figure 5-7. Axial modulus degradation versus applied axial stress for (a) [0/90/∓45]s and (b) [0/90]s GFRP laminates,

at the start of the creep tests, and at the end, for different biaxial loads.

In order to better understand the effects of biaxial loading on stiffness degradation, time-dependent

stiffness degradation envelopes have been developed to determine the amount of stiffness loss for different

values of biaxial loading in both laminates studied in this chapter. In Figure 5-8 (a), we show the stiffness

degradation contour lines in terms of the axial and transverse loads for the [0/90/∓45]s GFRP laminate,

before the creep test, and at the end. In each case, the stiffness degradation contour lines correspond to a

loss of 5 %, 10% and 15% in the initial axial modulus, 𝐸𝑥0, or in the initial transverse modulus, 𝐸𝑦

0. Looking

first at the 5% stiffness degradation contour, at time t = 0 (red), we can see that for low transverse loads

the line is nearly vertical, showing that the transverse load does not affect crack density evolution. This

can be confirmed by inspecting Figure 5-5, where it is seen that the crack density in the 90° ply is barely

affected by the transverse load. Although the 45° and -45° layer are affected by the transverse load, their

effect on stiffness loss at low loads is much lower. Similarly, the 5% contour line is horizontal when the

transverse loads are much higher, at which point the stiffness loss is dominated by cracking in the 0° ply.

Looking at the trends for the two other stiffness contour lines, corresponding to a loss in stiffness of 10%

and 15%, the lines are more slanted. This is due to the effect that transverse loads have on crack

multiplication at high loading levels, in the +45° and -45° layer. From inspection of the stiffness

degradation contour lines at the end of the creep simulations, it is clear that the lines have been shifted to

much lower values of stress. This suggests that the viscoelastic-viscoplastic properties of the matrix affect

the patterns of stiffness degradation under uniaxial loading. For example, the 5% stiffness degradation

contour line starts at 125 MPa of axial load, instead of 180 MPa. Under transverse loading, the 5% stiffness

degradation contour line crosses the vertical axis at 80 MPa when viscoelastic-viscoplastic deformation

has been taken into account, while it only crosses the axis at 120 MPa when creep viscoelastic-viscoplastic

deformation is ignored. These marked effects are due to two factors which affect the multiplication of

micro-cracks in the different plies of the laminate. First, viscoelastic-viscoplastic deformation increases

Page 99: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

79

the driving force for crack multiplication by increasing the amount of strain for a given applied load.

Second, under time-dependent deformation, the critical energy release rate (see Eq. (3.16)) degrades due

to the behavior of the matrix, which enhances crack multiplication under constant loading.

In Figure 5-8(b), we have shown the stiffness degradation contour lines for the [0/90]s GFRP cross-

ply, with the same ply properties as the [0/90/∓45]s laminate. The contour lines correspond to 5%, 10%

and 15% stiffness loss in either the axial, 𝐸𝑥0, or transverse, 𝐸𝑦

0, modulus. Looking at the 5% stiffness loss

line at time t = 0, we can see that the 5% stiffness degradation occurs at 175 MPa in the axial, and 100

MPa in the transverse direction. When viscoelastic deformation is taken into account (there is no

viscoplasticity in the cross-ply because of the absence of shear stress), 5% stiffness degradation occurs at

125 MPa of axial load, and 75 MPa of transverse load. All the contour lines are slanted, due to the effect

of transverse loading on crack multiplication. This can be explained by inspection of Figure 5-5(b), where

it can be seen that a transverse stress causes crack density in the 90° ply to increase, for a given axial load.

In the case of the 0° ply, large transverse stresses cause an enhancement in crack density, leading to further

stiffness degradation. As for the [0/90/∓45]s laminate, the creep strains cause stiffness degradation contour

lines to be shifted to much lower stresses; the shapes are also changed because of the creep strain

deformation that occurs in the laminate.

(a) (b)

Figure 5-8. The lines in the stiffness degradation contour map correspond to constant levels of stiffness loss after

time-dependent deformation under biaxial loading of (a) [0/90/∓45]s and (b) [0/90]s GFRP laminate.

5.2.2 CFRP

A detailed parametric study has been performed to predict the crack density and stiffness degradation

in the [0/90/∓45]s and [0/90]s carbon fiber-bismaleimide (IM7/5260) laminates with viscoelastic-

viscoplastic properties described by Eq. (3.5) and material properties obtained from Tuttle et al [16]. The

ply thickness was set to 0.14 mm. The parameters for the energy-based crack multiplication model and the

SDM model are given in Table 5-4. The effect of both axial (𝜎𝑥) and transverse (𝜎𝑦) tensile loads on the

evolution of crack density in the plies of the two CFRP laminates has been predicted, as well as their effect

Page 100: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

80

on the laminate stiffness in the axial (X) and transverse (Y) directions. When a transverse load (𝜎𝑦) is

applied to the laminate along with an axial load, the crack density in each layer for a given axial stress (𝜎𝑥)

may or may not be significantly affected depending on the orientation of the ply and the stacking sequence

of the laminate.

Table 5-4. Parameters for the viscoelastic part of the model [24]

A1 (Pa-1) A2 (Pa-1) A3 (Pa-1) A4 (Pa-1)

Transverse

Shear

𝜆𝑟(min-1)

1.200 x 10-12

5.723 x 10-12

0.001

1.795 x 10-12

3.942 x 10-12

0.01

8.823 x 10-13

3.874 x 10-12

0.1

1.954 x 10-13

1.001 x 10-12

1

Table 5-5. Elastic properties of the plies, measured at 93 °C, obtained from Tuttle et al [24]

Elastic property Value

E1(GPa)

E2(GPa)

𝜈12 G12 (GPa)

157

8.3

0.3

5.81

Table 5-6. Parameters for the viscoplastic part of the model [24]

n N

Transverse component

Shear component 0.16

0.33

6.25

10.11

Table 5-7. Damage model parameters for IM7/5260

Laminate [0/90/∓45]s [0/90]s

Ply angle 0° 90° -45° 45° 0° 90°

Stiffness

degradation

parameter

(GPa)

a1 0.41 4.17 3.92 4.01 0.41 4.28

a2 4 0.44 3.53 4.03 4.03 0.44

a3 1.66 1.84 0.48 0.5 2.08 2.27

a4 2.56 2.68 -0.036 -0.5 2.57 2.73

Critical

energy

release rate

GIc0

80 80 80 80 100 100

COD

evolution

parameters

c1 3.1159 1.1957 1.1957 1.1773 3.1159 1.1773

c2 0.266 0.098 0.098 0.196 0.266 0.196

c3 1.5484 1.5422 1.5422 1.5248 1.5484 1.5248

Time-dependent stiffness degradation envelopes have been developed to determine the amount of

stiffness loss for different values of biaxial loading in both laminates studied in this thesis. In Figure 5-9(a),

we show the stiffness degradation contour lines in terms of the axial and transverse loads for the

[0/90/∓45]s CFRP laminate, before the creep test, and at the end. In each case, the stiffness degradation

contour lines correspond to a loss of 1 %, 2% and 3% in the initial axial modulus, 𝐸𝑥0, or in the initial

Page 101: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

81

transverse modulus, 𝐸𝑦0. Looking first at the 1% stiffness degradation contour, at time t = 0 (red), we can

see that for low transverse loads the line is vertical, showing that the transverse load does not affect crack

density evolution. Similarly, the 1% contour line is horizontal when the transverse loads are much higher,

at which point the stiffness loss is dominated by cracking in the 0° ply. Looking at the trends for the two

other stiffness contour lines, corresponding to a loss in stiffness of 2% and 3%, the lines are much more

slanted. Looking next to the stiffness degradation contour lines at the end of the creep simulations, it is

clear that the shapes are very similar. However, the contour lines are shifted to lower loads. The 1%

stiffness degradation contour line starts at 500 MPa of axial load, instead of 750 MPa.

In Figure 5-9(b), we have shown the stiffness degradation contour lines for the [0/90]s CFRP cross-

ply, with the same ply properties as the [0/90/∓45]s laminate. The contour lines correspond to 1%, 2%and

3% stiffness loss in either the axial, 𝐸𝑥0, or transverse, 𝐸𝑦

0, modulus. Looking at the 1% stiffness loss line

at time t = 0, we can see that the 1% stiffness degradation occurs at 900 MPa in the axial, and 500 MPa in

the transverse direction. All the contour lines are slightly curved, due to the effect of transverse loading on

crack multiplication. As for the [0/90/∓45]s laminate, the creep strains cause stiffness degradation contour

lines to be shifted to much lower stresses, however the overall shapes are maintained.

5.2.3 Effect of temperature on the stiffness degradation envelopes

The effect of temperature on the damage behavior and stiffness degradation has been studied using the

SDM model and temperature dependent material properties for [0/90/∓45]s and [0/90]s laminates. The

material model (Eq. 3.5) is applicable up to a temperature of about 150 °C [24]. Therefore 150 oC was

chosen as the maximum simulation temperature. In order to perform the simulations at high temperatures,

the parameters of the SDM model were calculated by accounting for the different elastic properties of the

plies at 150 °C. The critical energy release rate for crack multiplication was obtained using the trends

reported by Asadi [26]. The time-dependent part of the critical energy release rate was obtained using the

method detailed in Chapter 3.

Page 102: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

82

(a) (b) Figure 5-9. Stiffness degradation contour lines. The lines in the stiffness degradation contour map correspond to constant levels

of stiffness loss after viscoplastic deformation under biaxial loading of (a) [0/90/∓45]s and (b) [0/90]s CFRP laminate.

Figure 5-10(a) shows the stiffness degradation contour lines for the [0/90/∓45]s laminate at two

different temperatures, 93 °C and 150 °C. The different lines correspond to a loss in stiffness of 1%, 2%

and 3%. It is clear that temperature has almost no effect on the shapes of the envelopes in this laminate,

and only shifts the envelopes towards lower values of stress. This suggests that the dominant effect of

viscoelasticity-viscoplasticity on damage evolution is through the change in critical energy release rate

with respect to time, and not the increase in strain. As for the lower temperature, at 150 °C, the 1% stiffness

degradation contour line is vertical for low transverse stresses and horizontal for low axial stresses, due to

the low effect of transverse load on cracking in the 90° ply of the laminate, and the low effect of axial load

on cracking in the 0° ply. At larger applied loads, the contour lines are more slanted. In Figure 5-10(b), we

have shown the stiffness degradation contour lines for the cross-ply [0/90]s laminate at two temperatures

of 93 °C and 150 °C. For the cross-ply laminate, the effect of temperature is lower than for the quasi-

isotropic laminate, and the shapes of the contour lines are not affected significantly. In fact, the 1% stiffness

degradation contour line at 150°C is located slightly beyond the 1 % stiffness degradation contour line at

93°C at some points due to the larger initial critical energy release rate (𝐺𝐼𝑐0 ) at higher temperature.

However, the rate of decay in critical energy release rate with respect to time causes most of the contour

lines at 150°C to be shifted towards lower values of biaxial loads.

Page 103: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

83

(a) (b)

Figure 5-10. Stiffness degradation contour lines for different temperatures. a) Stiffness degradation contour lines at the end of the

creep simulations (T = tend) at two different temperatures for (a) [0/90/∓45]s and (b) [0/90]s IM7/5260 laminates. This material

has a glass transition temperature of 260 °C.

5.3 Discussion

The model results presented in this chapter are an improvement on the SDM model that includes the

effects of time-dependent damage due to viscoelasticity and viscoplasticity on the creep response of multi-

directional laminates under multi-axial loading. The viscoplastic model used in this chapter was based on

the Zapas-Crissman functional [183] which relates the viscoplastic strain to the loading history of the plies.

Different model parameters were used for the GFRP and CFRP composites.

In order to succinctly summarize the findings of this chapter, the stress-strain curves under uniaxial

loading were plotted for the two material systems in Figure 5-11. For each material system, four stress-

strain curves are plotted: linear elastic model without damage; linear elastic model with damage (before

creep); viscoelastic-viscoplastic model without time-dependent damage; viscoelastic-viscoplastic model

with time-dependent damage (after creep deformation). In the case of the GFRP composite (subplot (a)),

the effect of damage is large, and therefore, before creep deformation, significant non-linearity in the

stress-strain curve can already be observed. Time-dependent deformation, and then time-dependent

damage, cause more severe non-linearity in the stress-strain response of the composite. The curve which

includes the time-dependent damage evolution constitutes a worst-case scenario for the uniaxial

mechanical response of the composite, and could be used as a design tool for applications where constant

loading at higher temperatures might be important. In the case of CFRP, the degradation in stiffness is

much lower, therefore the stress-strain response is not affected much by micro-crack multiplication.

Similarly, the viscoelastic-viscoplastic deformation is much less significant due to the stiffer carbon fibers,

and the stress-strain response is not significantly affected by time-dependent damage. The plies are also

thinner, which constrains crack opening and leads to less stiffness degradation. However, micro-crack

density reaches high values at high loading levels, which will eventually lead to delamination and critical

Page 104: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

84

failure. Therefore, even though the stress-strain response is not remarkably affected by micro-cracking,

the model developed in this thesis can provide information into the sequence of mechanisms leading to

failure of the laminate.

(a) (b) Figure 5-11. (a) Stress-strain curves for a [0/90/∓45]s GFRP laminate under uniaxial loading using different material models. (b)

Stress-strain curves for a [0/90/∓45]s CFRP laminate.

The viscoelastic-viscoplastic properties of the ply will cause the stress to vary with time during

constant loading. When cracking occurs, the local stress redistribution will cause the stiffness (elastic) of

the laminate to decrease, which is accounted for in the SDM model. The effect of damage on the evolution

of time-dependent stresses is not explicitly accounted for in our model, as explained in Chapter 3. As of

now, the micro-damage model is limited to the degradation in elastic properties, and viscoelastic-

viscoplastic stress redistribution under micro-cracking is beyond the capabilities of the micro-damage FE

model that has been developed. This approximation is deemed sufficient based on the results of Figure

5-3(c), which show that the model is accurate enough to predict creep strain in a viscoelastic laminate

undergoing time-dependent micro-cracking. Although cracking is assumed not to affect the creep response

under constant crack density, the crack multiplication model predicts crack multiplication under constant

loading, which affects the creep response. Therefore, damage and creep are coupled accurately, even

though the local time-dependent stress redistribution at the microscopic level that occurs upon cracking is

not accounted for explicitly.

Under constant loading, the creep strain of a multi-directional viscoelastic-viscoplastic laminate will

increase. However, even though the globally applied stress is constant, each ply undergoes time-dependent

deformation at a different rate due to their different orientations. This leads to time-dependent changes in

ply stresses. Overall, however, the laminate stress is constant and equal to the applied load, as per Classical

Laminate Plate Theory. The evolution of stresses during creep loading have been plotted in Figure 5-12(a)

for the GFRP quasi-isotropic laminate under a uniaxial constant load of 300 MPa. For both plies, the

Page 105: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

85

transverse stress decreases with respect to time following ramp-up to the constant loading level. This is

due to the presence of a very stiff 0° ply, whose behavior in the axial direction is elastic and time-

independent. This layer inhibits time-deformation of the laminate, and effectively reduces the stress in the

more compliant plies.

As explained in Chapter 3 of this thesis, the energy release rate for crack multiplication assumes a

linear elastic relationship between stress and strain in the plies. In reality, as shown in Figure 5-12 (a), the

stresses in the plies are time-dependent due to the viscoelastic-viscoplastic behavior of the matrix. The

effect of including stress relaxation on the crack multiplication predictions has been investigated. The

effect of including stress relaxation is shown in Figure 5-12(b) for a GFRP laminate and compared to the

predictions which do not include stress relaxation. As expected, there is a slight effect on crack

multiplication, however, stress relaxation does not significantly affect the evolution of crack density. The

biggest effect of viscoelastic-viscoplastic properties on crack multiplication is through the decay in 𝐺𝐼𝑐

with respect to time, and the effect of stress relaxation can therefore be ignored.

(a) (b) Figure 5-12. (a) Time-dependent evolution of transverse ply stress for the [0/90/∓45]s under uniaxial constant load of 300 MPa,

during ramp-up and subsequent creep deformation. (b) Effect of including viscoelastic-viscoplastic stress relaxation on crack

multiplication predictions for a [0/90/∓45]s GFRP laminate.

Summary for Chapter 5

A continuum damage mechanics-based SDM model was developed to incorporate the effects of

viscoelastic-viscoplastic properties in GFRP laminates. Schapery’s thermodynamics-based viscoelastic-

viscoplastic constitutive model was implemented in the framework of the SDM model to investigate

progressive damage by matrix micro-cracking in laminates. After initial validation with available

experimental data, two GFRP laminates, namely [0/90/∓45]s and [0/90]s, were studied in this chapter.

The damage behavior was predicted under an initial quasi-static step followed by viscoelastic-viscoplastic

Page 106: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

86

creep deformation under biaxial loading. The laminate was quasi-statically loaded up to the creep load

level and then it was allowed to undergo creep at that constant load.

It was observed that the presence of a transverse load can increase the crack density compared to that

under uniaxial loading. In the quasi-isotropic laminate, an increase of transverse load from 0 to 180 MPa

with a constant longitudinal load of 210 MPa led to increase in crack density from 0.6 cr/mm to 1.5 cr/mm

in the 45° ply. In the 90° ply, on the other hand, the crack density was barely affected when a transverse

load is applied. For a [0/90]s cross-ply, a transverse load for a given axial load enhanced cracking in both

plies.

The viscoelastic-viscoplastic deformation was found to reduce the micro-cracking initiation stress for

both laminate stacking sequences. In the quasi-isotropic laminate, damage initiation was found at a uniaxial

load of 140 MPa in the longitudinal direction under quasi-static deformation. However, when viscoelastic-

viscoplastic deformation was accounted for, the damage initiation in the laminate was at 80 MPa of

longitudinal uniaxial load. In the case of the cross-ply, the crack initiation stress decreased from 150 MPa

to 80 MPa when viscoelastic-viscoplastic deformation was taken into account. For both stacking

sequences, an increase in transverse stress enhanced stiffness degradation.

To comprehensively quantify the damage under biaxial loading, stiffness degradation maps were

developed under different combinations of axial and transverse loads. It was found that a given stiffness

degradation occurs at lower axial loading levels when a transverse load is applied to the laminate. When

the effect of creep was accounted for, the level of stiffness degradation for a given biaxial loading scenario

was always more severe. Depending on the effect of transverse load on the evolution of damage in the

different plies of the laminates, the slopes of the contour lines were different. Due to the extensive effects

of viscoelastic-viscoplastic deformation in these composites, the contour lines describing the damage state

after creep were shifted to lower loads, indicating more significant damage, and underwent changes in their

shape, due to the time-dependent evolution in creep strain.

Taken together, these results demonstrate the ability of the SDM model to predict time-dependent

damage progression in viscoelastic-viscoplastic laminates under multi-axial constant loading. Through its

ability to accurately take into account multi-axial loads, and to predict the time-dependent damage

progression in individual plies, the model paves the way for the design of damage-tolerant composites

structures which can withstand more demanding thermo-mechanical environments.

In this thesis, the focus has been on constant loading and creep deformation. The model can be further

extended to include different loading scenarios. In order to extend the methodology to time-dependent

loading, the micro-damage FE model should be augmented to be able to model the viscoelastic and

viscoplastic behavior of each ply.

Page 107: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

87

Chapter 6. Strain rate-dependent damage model

for a short fiber composite

In this chapter, the effects of strain rate on the evolution of interfacial debonding in a short fiber

thermoplastic matrix composite have been studied using computational micro-damage modelling. A three-

dimensional RVE of a short glass fiber/thermoplastic matrix composite has been developed and a non-

linear viscoelastic model has been implemented for the matrix. A bilinear Cohesive Zone Model has been

used to model the behavior of the fiber/matrix interface. Following validation with respect to a previous

numerical study, an extensive study of the effects of strain rate, fiber diameter and fiber type on micro-

damage evolution has been performed.

6.1 Model validation

In order to validate the model, the methodology was applied first to a short carbon fiber/epoxy

composite, studied by Pan and Pelegri [200]. An axisymmetric model was created, and the interface

modelled using a CZM. The results of our implementation are shown in Figure 6-2, and compared to the

predictions of Pan and Pelegri. Clearly, the present model perfectly predicts debonding evolution versus

applied strain. Damage initiates at approximately 0.22% applied strain, and reaches its maximum value

(full debonding) at 0.6% strain. Interestingly, the shape of the damage evolution curve is similar to that

observed for crack multiplication in laminates reported in Chapters 4 and 5, even though the microscopic

phenomena that these curves describe are fundamentally different. Due to the complexity involved in

measuring debonded area fraction versus strain experimentally, the experimental validation of this curve

is not possible at this point in time, and therefore the results shown in Figure 6-2 are deemed satisfactory

for the purposes of this thesis.

6.2 Effect of strain rate on damage evolution

Following validation of the model with respect to experimental data, the CZM was applied to a short

glass fiber composite embedded in a HDPE matrix. Initially, the matrix and fiber were modelled as linearly

elastic, and the fiber-matrix interface was modelled using the CZM. The fiber diameter was set to Df = 40

μm and the length Lf = 200 μm (giving a total length of 400 μm when including both fibers). In Figure

6-1, the Von Mises stress profile has been plotted under an applied strain of 0.23 %. The contour plot in

the inset shows a zoomed-in part of the RVE around the two fiber tips. Note that interfacial damage has

not been accounted for in these results. Stresses provided in the legend are expressed in MPa, and distances

Page 108: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

88

are in micrometers. The matrix stress is on average much lower than the fiber stress, because of its lower

modulus of elasticity. The stress in each fiber is lower at the tip, but increases along the length. At the fiber

tip, there is a mismatch between the elastic properties of the matrix and fiber, which leads to the occurrence

of a stress concentration. The maximum stress in the fiber reaches approximately 125 MPa, while the

matrix is under a stress of 10 MPa approximately. In Figure 6-1 (b), the displacement in the vertical

direction has also been plotted. Displacement is almost linear with respect to position along the RVE, going

from 0 μm at the bottom face, to 0.94 μm at the top face. Around the region of the fiber tips, the

displacement is not uniform along the cross-section of the RVE due to the stress concentration occurring

as a result of the mismatch in elastic properties of the fiber and matrix.

In the current chapter, the evolution of damage at the interface is taken into account in the calculations

by defining the degradation in stress transfer with respect to normal and shear displacements across the

interface. The CZM provides a local damage variable which describes the loss in stress transfer across the

interface. On the other hand, a global damage variable can be defined through a damage variable 𝐷 equal

to 0 when the total interface area is intact, and 1 when the interface has failed at all points. It is given by

𝐷 =𝐴

𝐴0, where 𝐴0 is the initial area of the fiber-matrix interface, and 𝐴 is the damaged area following the

onset of debonding from the fiber tip. In Figure 6-1(c), a contour plot of the bottom half-fiber shows the

value of the local damage variable, which corresponds to the loss in load-transferring capability of the

interface and is obtained using the cohesive zone model. A value of 1 shows the interface has failed, while

a value of 0 shows the interface is intact. The transition from an intact interface to a completely failed

interface occurs over a very small distance along the half-fiber, as shown by the small transition zone from

completely damaged interface to intact interface. Damage is also uniform around the circumference of the

fiber, which is as expected given the symmetry assumed in the model. In Figure 6-1(d), we have plotted

the damage variable based on debonded area fraction in red. We have also plotted the damage variable

based on the reduction in secant modulus (obtained from the stress-strain curve), defined as follows: 𝐷 =

1 −𝐸

𝐸0, where 𝐸0 is the initial secant modulus (equal to Young’s modulus of the RVE), and 𝐸 is the current

secant modulus. Clearly, the two quantities are related, as expected, and the secant modulus loses 90 % of

its original value due to debonding. This observation further validates the model.

Page 109: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

89

(a) (b)

(c) (d)

Figure 6-1. (a)Von Mises stress contour for the short fiber composite RVE undergoing longitudinal deformation with inset

showing the stress in the neighborhood of the fiber tips. Stresses are expressed in MPa. (b) Longitudinal displacement contour

plot of the RVE under an applied strain of 0.1 %. Displacements are expressed in μm. (c) Local damage variable for a fiber Df =

40 μm and Lf = 200 μm. (d) Global damage variable versus applied strain and secant modulus loss curve

In Figure 6-3, the Von Mises stress distribution under increasing deformation has been plotted for one

of the RVEs along a vertical cross-section. The matrix here is modelled as linearly elastic; viscoelasticity

is accounted for in the next simulations. At low strain, no debonding has occurred, and stress is transferred

from the matrix to the fibers, leading to an increase in stress from the fiber tips to the midpoint. The matrix

region located directly between the fiber tips shows a very low Von Mises stress. However, the region

around the fiber tips shows a large stress concentration, with the Von Mises stress reaching more than 15

MPa. Under further loading, at 0.76% strain shown in subplot (b), some fiber-matrix debonding has

occurred. The stresses in the fibers are much lower in the regions that have undergone debonding. The

stresses in the matrix around these regions is also much higher. This observation suggests that stress around

the fiber tips is carried by the matrix, as the reinforcing effect of the fiber is inhibited by the debonding

process. Finally, at 4 % strain shown in subplot (c), the interface has almost completely failed, and the

stress in the fibers is very low; there is a slight increase far from the fiber tips, where the interface is still

intact. The stress in the matrix is much higher than in the fiber, as it is carrying almost the entire load

applied to the RVE.

𝐷 =𝐴

𝐴0

Global damage

variable

Local damage

variable

Page 110: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

90

(a) (b) (c)

Figure 6-2. (a) Axisymmetric short glass fiber/epoxy composite RVE used to validate the model. (b) Displacement response under

1% strain. (c) Debonded area fraction versus applied strain for a short glass fiber/epoxy composite, and comparison to the work

of Pan and Pelegri [200].

In Figure 6-4, the Von Mises stress distribution under increasing deformation has been plotted for the

same RVE. The strain rate applied is 10-5 s-1. From the first image, corresponding to 0.088% strain, it can

be seen that the stress along the fibers increases from the tip to the midpoint due to stress transfer from the

matrix to the fiber. As the strain is increased further, to 0.76%, full normal separation occurs at the fiber

faces. Some partial debonding also occurs at the fiber tips. This leads to a lower stress at the ends of the

fibers. At 4 % strain, significant inelastic deformation of the matrix around the to fiber has occurred due

to the high stresses in that region. An interfacial crack has also opened at the fiber tips. The matrix region

away from the fiber tips is still in the linear range of deformation. Debonding has not extended along the

fiber-matrix interface, and there is still effective stress transfer between the matrix and fiber further away

from the fiber tips. Interestingly, the matrix region around the lower fiber tip has not undergone extensive

deformation.

Page 111: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

91

Figure 6-3. Von Mises stress distribution in the short fiber RVE with Df = 40 µm and Lf = 200 µm under different strain levels

with a linear elastic HDPE matrix.

In order to understand the effect of matrix non-linearity on damage evolution at the fiber-matrix

interface, the strain rate applied to the RVE was changed by implementing the non-linear behavior into

Abaqus based on the stress-strain response calculated using the non-linear viscoelastic model for HDPE

(see Chapter 3).

In Figure 6-5 (a), we have plotted the global damage variable which describes the fraction of debonded

interface area with respect to applied strain for a fiber diameter Df = 40 μm and a fiber length Lf = 200 μm.

The different curves correspond to the different strain rates applied to the RVE. The simulations were

stopped at 4% strain because damage did not evolve significantly past this strain level. In the case of the

linear elastic matrix, damage initiates early on in the simulations, at approximately 0.1% . It evolves at a

constant rate with respect to applied axial strain until the global strain reaches a strain of 3.6 %, after which

damage reaches its maximum value. As the strain rate is decreased and the non-linear behavior of the

HDPE matrix becomes more pronounced, the evolution of damage is affected. In all cases, damage initiates

under an applied strain of 0.1 %. However, with decreasing strain rate, damage evolves more slowly. Under

a strain rate of 102 s-1, the damage variable reaches a maximum value of 0.95; when the strain rate is lower,

it only reaches a maximum value of 0.3. This behavior suggests that in the case of a relatively thick fiber

(Df = 40 μm), debonding evolution is significantly affected by the viscoelastic behavior of the matrix.

휀 = 0.08 % 휀 = 0.76 % 휀 = 4 %

Page 112: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

92

Figure 6-4. Von Mises stress distribution in the short fiber RVE with Df = 40 µm and Lf = 200 µm under different strain levels

for a viscoelastic matrix under a strain rate of 10-5 s-1.

The stress-strain curves obtained for these RVEs using Abaqus are shown in Figure 6-5 (b), with

different applied strain rates. The initial part of the curves is always linear, because the matrix has not been

strained high enough to enter the non-linear region, and no debonding has occurred. In the case of a linearly

elastic matrix, the response becomes non-linear above an applied strain of 0.5 % due to the onset of fiber-

matrix debonding, which reduces the load-carrying capacity of the short fiber composite. As the strain rate

is decreased using the viscoelastic constitutive model for the matrix, the stress is lower. This non-linear

behavior of the RVE is affected by matrix non-linearity in two different ways: the viscoelastic behavior of

the matrix induces a lower matrix stress as the strain rate is decreased, and the viscoelastic behavior also

affects damage evolution. As the strain rate is decreased, the evolution of damage is delayed, because the

interface shear stresses are lower. The fact that stress is reduced when the non-linear behavior of the matrix

becomes more pronounced shows that debonding does not dominate the response of this RVE, and that

instead matrix inelasticity largely governs the stress-strain response. This can be illustrated by calculating

the amount of work applied to the RVE which is dissipated as matrix inelastic deformation, versus the

amount which is dissipated in the form of damage (i.e. debonding) evolution. These quantities have been

plotted in Figure 6-6 (a). Clearly, most of the applied work is dissipated into inelastic deformation, while

the remainder is stored in the form of elastic energy. The amount of energy dissipated in the form of

damage is minimal at lower strain rates.

The inelastic strain was calculated in the RVE using the following equation: 휀𝑝𝑙 =1

𝑉∑ 휀𝑖𝑛

𝑖 𝑑𝑉𝑖𝑖 where

𝑉 is the volume of the RVE, 𝑉𝑖 is the volume of element 𝑖, 휀𝑖𝑛𝑖 is the inelastic strain in the element obtained

4 %0.76 %0.088 %

Interfacial crack

openingPartial

debonding

along sides

Full debonding

at fiber face

Page 113: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

93

from Abaqus, and the summation runs over all matrix elements in the RVE. In Figure 6-6 (b) the matrix

inelastic strain has been plotted, which shows that most of the deformational response of the matrix is

dominated by its inelastic behavior. By inspection of the RVE, it was found that inelastic strain was

concentrated at the fiber tips (see Figure 6-4(c)). Away from the fiber tips, strain was elastic. This

observation suggests that the stress concentration at the fiber tips induces the nucleation of a zone of

extensive inelastic deformation. Inspection of the RVE also showed the growth of interfacial opening at

the fiber tip, instead of debonding, when viscoelastic properties were accounted for.

The COD at the fiber tip obtained in these simulations (where strain rate was varied by changing the

constitutive behaviour of the matrix) describes the growth of the interfacial opening, and is shown in Figure

6-7. When the matrix is modelled with a linear elastic property, the COD is close to 0, and there is no

microcrack opening. However, as the strain rate is lowered and inelastic effects become more pronounced,

the COD increases much more significantly. Under a strain rate of 10-2 s-1, the COD reaches 6.5 μm; under

a strain rate of 10-5 s-1, the COD reaches 9 μm. This suggests that under low strain rate, microcrack opening

will dominate the response of a short fiber composite, most likely leading to crack growth into the matrix

as interfacial cracks from adjacent fibers coalesce. These observations are in agreement with experimental

observations by Jerabek et al [204], who found that under high strain rates, debonding in a glass particle

reinforced plastic dominated the response, while matrix shear deformation was more important at low

strain rates. It should be noted that the present model does not account explicitly for matrix microcracking.

Instead, the focus is on interfacial microcrack opening from the fiber tips.

The results reported here can be used in analytical studies of damage in short fiber composites. For

example, Notta-Cuvier et al [156] implemented an analytical model for predicting damage and failure in

randomly oriented short fiber composites. They defined an expression relating the applied strain to

debonded area fraction. They also postulated that at a certain critical void volume fraction, crack

propagation and critical failure of the material could be assumed. The crack opening shown in Figure 6-7

could be used in such an analytical model to predict strain-rate dependent failure induced by the

coalescence of interfacial cracks nucleated at fiber tips. These results improve on previous studies by

accurately accounting for the viscoelastic properties of the matrix.

Page 114: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

94

(a) (b)

Figure 6-5. (a) Evolution of the damage variable (fraction of debonded area) with respect to applied strain for different strain rates

in a short glass fiber composite with fiber diameter Df = 40 μm. (b) Stress-strain curves.

(a) (b)

Figure 6-6. (a) Energy dissipation in the RVE of a short glass fiber composite (Df =40 μm) for a viscoelastic matrix deformed at

10-1 s-1. (b) Inelastic strain evolution under different strain rates.

Figure 6-7. Evolution of COD under different strain rates for an interfacial crack at the fiber tips in a short glass fiber composite

with Df = 40 μm.

In Figure 6-8 (a), the interfacial shear stress along the fiber-matrix interface has been plotted at 0.08%

strain (before the onset of debonding) and at 4% strain (after full debonding), for a linearly elastic matrix,

Page 115: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

95

and a viscoelastic matrix under a strain rate of 10-5 s-1. Subplot (b) shows the axial stress along the fiber

under the same loading conditions. Looking first at subplot (a), in the case of the linearly elastic matrix,

the shear stress is highest at the fiber tip due to the stress concentration effect. The maximum stress reaches

11 MPa, which is the stress at which mode II debonding initiates, according to the CZM implemented in

this thesis. As one moves along the length of the fiber, the shear stress eventually reaches 0. After

significant debonding has occurred (blue dotted curve), the shear stress along the interface is 0 until one

reaches a distance of 170 μm along the fiber. At this location along the fiber, there is a transition from a

fully debonded region, to a fully bonded region. The shear stress increases at that point due to the effective

load transfer enabled by the pristine section of the interface. Looking next to the non-linear matrix, prior

to debonding (yellow dashed curve), the shear stress is lower than in the case of the linearly elastic matrix,

which can be attributed to the lower load-carrying capability of the matrix. At 4 % strain (purple dashed

curve), some debonding has occurred, but it only extends up 60 μm along the fiber. The shear stress

concentration in that case only reaches 5 MPa, which is not high enough to induce mode II debonding.

This therefore suggests that the viscoelastic behavior of the matrix leads to a lower shear stress

concentration at the fiber tips, which inhibits mode II debonding.

Looking next at subplot (b), which shows the axial stress in the fiber, in the case of a linearly elastic

matrix prior to debonding, the stress increases along the length of the fiber, due to the effective transfer of

load from the matrix to the fiber. After significant debonding has occurred, the stress in the fiber is much

lower as the load is mainly carried by the matrix. The portion that is still bonded to the matrix still carries

some load. In the case of the viscoelastic matrix, prior to debonding, the stress in the fiber is lower than in

the linearly elastic matrix due to the lower shear stress across the interface. After debonding, however, the

stress in the fiber is much larger than in the linearly matrix at the same level of applied strain due to less

extensive debonding and more effective stress transfer across the interface. These results therefore suggest

that in a viscoelastic matrix, more load will be carried by the fiber, which might lead to the onset of fiber

breakage.

Table 6-1. Fiber dimensions studied in this thesis

Lf (μm) Df (μm)

200 10

200 20

200 40

Page 116: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

96

(a) (b)

Figure 6-8. (a) Shear stress distribution along the fiber matrix interface at two different applied strains, for a linearly elastic matrix,

and a viscoelastic matrix under a strain rate of 10-5 s-1. (b) Axial stress distribution.

(a) (b)

Figure 6-9. (a) Damage variable versus applied strain for 3 fiber dimensions with a linearly elastic matrix. (b) Maximum damage

reached for 3 fiber dimensions as a function of strain rate.

The effect of strain rate on debonding evolution was studied for 2 other fiber diameters, provided in

Table 6-1. The corresponding evolution of the global damage variable when the matrix is modelled as

linearly elastic is plotted in Figure 6-9 (a) as a function of applied strain. In the case of the thicker fiber,

damage initiates at a slightly lower value of applied strain and reaches its maximum more quickly. In the

case of the thinner fibers, damage initiates at larger applied strains, and does not evolve as quickly. These

results are due to the different distribution of interfacial stresses induced by fibers with different diameters.

The effects of viscoelasticity and strain rate were studied next, by plotting the maximum amount of

damage reached for the different fiber dimensions as a function of strain rate. This maximum value was

reached at a strain of 4%. The results are plotted in Figure 6-9(b). In the case of the thinner fibers, the

viscoelastic properties cause a significant decrease in the maximum debonded area fraction reached in the

simulations, even at large strain rates. In these cases, most of the applied work is dissipated into inelastic

Page 117: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

97

deformation which induces the growth of the interfacial opening. In the case of the thickest fiber,

debonding evolution is more gradual with respect to strain rate, and even at relatively low strain rates

debonding is significant. Taken together, these results suggest that fiber diameter has an effect on the rate-

dependent micro-damage processes in short fiber composites. These observations can be rationalized based

on the different stress concentration enforced at the fiber tips under different fiber diameters. Large fibers

undergo debonding more readily at lower applied strain, at which point inelastic deformation of the matrix

is minimized. Thinner fibers, on the other hand, do not undergo debonding as readily, such that inelastic

deformation becomes dominant before debonding can initiate.

The mechanical behavior of short fiber composites is largely governed by the interfacial shear strength

(IFSS). This property can be partly controlled through proper processing. In order to understand the effect

of this parameter on the progression of micro-damage mechanisms, its value was varied for the fiber

diameter of 20 µm. The values chosen were: 11 MPa (used for the results shown in the previous

paragraphs), 20 MPa and 30 MPa. This range of interfacial shear strengths is representative of experimental

investigations for this material system. In order to understand the effect of IFSS on the deformation of the

matrix, the contour plot of strain energy density obtained using Abaqus has been plotted for these three

different values of interfacial shear strengths under a strain of 4 % in Figure 6-10 for a viscoelastic matrix

under 10-1 s-1 strain rate. The units are N/mm2. In the first case (lowest interfacial shear strength),

debonding has propagated a certain distance along the top fiber. At the same time, significant inelastic

deformation has occurred at the top fiber tip, accompanied with interfacial crack opening. The maximum

strain energy density reaches 40 MPa. In the second case, the debonded area has not propagated as far

along the top fiber, but the crack opening is larger, and the strain energy density for the same strain level

is also much larger. Lastly, in the third case, the interfacial crack opens at the bottom fiber tip, and the

strain energy density reaches an even higher value of 116 MPa.

In order to quantify the evolution of maximum strain energy density, its maximum value throughout

the RVE has been plotted for the three values of IFSS in Figure 6-11. As the applied strain increases, the

maximum strain energy density increases due to the significant inelastic deformation of the matrix at the

fiber tips. When the IFSS is increased, the maximum stored strain energy density increases significantly.

This suggests that in addition to the opening of the interfacial crack, the strain energy density in the matrix

adjacent to the crack is very large, which would most likely lead to the nucleation of a micro-crack into

the matrix region close to the fiber tips.

In summary, increasing the IFSS leads to less extensive debonding, larger crack opening displacement,

and more inelastic deformation in the matrix. Although the material model used in this chapter for the

matrix does not predict high strain behavior and matrix failure, it can be argued that the increase in strain

energy density with applied strain would lead to the nucleation of a micro-crack in the highly strained

Page 118: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

98

region around the fiber tips. With a larger IFSS, matrix micro-cracking becomes more likely. Moreover,

coalescence between interfacial cracks on adjacent fibers also becomes more likely as the COD at the fiber

tip increases.

(a) (b) (c)

Figure 6-10. Effect of IFSS on strain energy density at 4 % applied strain: (a) 𝜏𝐼𝐹𝑆𝑆 = 11 𝑀𝑃𝑎, (b) 𝜏𝐼𝐹𝑆𝑆 = 20 𝑀𝑃𝑎, (c) 𝜏𝐼𝐹𝑆𝑆 =30 𝑀𝑃𝑎 for a matrix loaded under a strain rate of 10-1 s-1. The glass fiber diameter is 20 μm.

Figure 6-11. Evolution of maximum strain energy density versus applied strain for three different values of the interfacial shear

strength. The glass fiber diameter is 20 μm.

The effect of fiber type was investigated next. Composites considered in the automotive industry are

based on hemp, glass and carbon fibers [205], [206]. Due to the different mechanical properties of these

fibers, specifically their Young’s moduli and Poisson’s ratios, it can be assumed that the evolution of

debonding will be affected. It is also likely that the effect of strain rate on damage evolution will depend

on the mechanical properties of the fibers. In the next set of results, the fiber diameter was maintained at

20 µm, and the length to 400 µm.

First, the evolution of debonded area fraction in terms of applied strain has been plotted for the different

fiber types (see Figure 6-12 (a)). The behavior of the carbon fiber is similar to the that of the glass fiber,

Page 119: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

99

with damage initiating at a low strain level (0.1 %). The behavior of hemp is different, with damage

initiation at approximately 0.75 %. Note that in all three cases, the interfacial shear strength is constant at

11 MPa. The different behavior of hemp can most likely attributed to its very low stiffness as compared to

the other two fiber types. This lower stiffness leads to a lower stress concentration at the fiber tips, thereby

inhibiting the growth in debonded area fraction.

In subplot (b), the maximum value of the damage variable (i.e. the debonded area fraction) has been

plotted for the three types of fibers with the RVE deformed under different strain rates. From this figure,

it is clear that the rate-dependency of debonding is similar for the types of fibers considered in this chapter.

Surprisingly, the rate-dependency of debonding for hemp is similar to that of carbon and glass at lower

strain rates, even though its Young’s modulus is much smaller. This observation suggests that the rate-

dependency of the debonding process occurring around the fibers tips is not significantly influenced by the

mechanical properties of the fibers.

(a) (b)

Figure 6-12. (a) Damage evolution for different types of fibers with different mechanical properties with a linear elastic matrix.

(b) Damage variable at 4% strain as a function of strain rate for different fiber types. In all cases, the fiber diameter is Df = 20 µm.

Summary for Chapter 6

The evolution of fiber/matrix debonding in short fiber composites with a thermoplastic matrix has been

studied with computational micro-damage mechanics using a bilinear Cohesive Zone Model for the

interface, and a nonlinear viscoelastic material model for the matrix. The CZM was validated with respect

to a previous model. The evolution of micro-damage was precisely monitored through quantitative post-

processing of the simulation data. It has been found that under high strain rate behavior, debonding

dominated the nonlinear response of the RVE, while it was less significant with decreasing strain rate. At

lower strain rates, the stress concentration at the tip of the short fibers induced the opening of an interfacial

crack. The rate of growth of this interfacial micro-crack was significantly affected by the strain rate level.

Page 120: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

100

The effects of fiber dimensions on the evolution of micro-damage mechanisms were also studied.

Composites with thin fibers were found to be more prone to void opening. Larger fibers, on the other hand,

were found to be more susceptible to debonding, with interfacial crack opening dominating the response

at low strain rates. The effects of interfacial shear strength and fiber types were also studied. A stronger

interface led to less significant debonding and a higher likelihood of micro-cracking. The fiber type, on

the other hand, did not significantly affect the strain-rate dependency of damage evolution.

One limitation of the current approach is that microcrack growth into the matrix following interfacial

separation is not explicitly predicted. It is not clear at this point at which level of COD the interfacial

micro-crack would grow into the matrix and cause composite failure. While different values of the COD

have been calculated for the different fiber sizes and different strain rates, it is not clear whether the results

demonstrate that micro-crack growth is more likely in certain RVEs. Future studies should focus on micro-

crack growth into the matrix in order to develop accurate criteria for rate-dependent failure of short fiber

composites. The results of the analysis are applicable to matrices with high ductility (more than 100%) due

to the large deformation occurring at the fiber tips. Moreover, the analysis is restricted to the lower range

of interfacial shear strengths (IFSS). The fibers are modelled using a linear elastic isotropic material model.

The loading scenario used in this chapter is purely quasi-static, even when high loading rates are

applied. The deformation calculated using FEA assumes no dynamic behavior. Under impact loading, it is

likely that the time-dependent evolution of stresses will lead to different damage mechanisms, and different

damage evolution patterns. While these effects will be important to consider in future studies, the goal of

this chapter is to isolate the effect of matrix viscoelasticity on the competition between damage

mechanisms.

Temperature and humidity are known to affect both the viscoelastic-viscoplastic properties of the

matrix, the natural fiber and the interface [75], [207], [208]; the material models should be adjusted

accordingly for these effects to be properly accounted for. Temperature and humidity will cause different

time-dependent behavior in the polymer. Residual stresses might also affect the debonding process, which

needs to be understood in the case of natural fibers, which are highly susceptible to moisture. The behavior

of the fibers was assumed to be linearly elastic, which might not be the case in natural fibers, which could

require more sophisticated mechanical models for a proper description of damage evolution at the fiber

level [209].

In this chapter, damage evolution has been studied from a microscopic perspective, focusing on a

single fiber composite. In order to incorporate these results into a multi-scale modelling approach, the

effects of fiber orientation, and fiber distribution will need to be accounted for. Debonding and microcrack

evolution are highly sensitive to the local environment, including the orientation of neighboring fibers, and

their length [69]. Once these effects have been accounted for, expressions can be developed for predicting

Page 121: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

101

damage evolution in terms of the applied stress. These expressions can then be implemented in structural

analysis software for improving the strain-rate dependent damage predictions of composite structures with

significant viscoelastic behavior. The results of this study therefore provide valuable insights into the

processes of rate-dependent damage in thermoplastic composites and pave the way for the design of

recyclable composites with a thermoplastic matrix to be used in rate-sensitive structural applications.

Page 122: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

102

Chapter 7. Structural analysis using a rate-

dependent SDM model

In this chapter, the multi-axial SDM model developed in Chapter 3 is implemented into Abaqus by

using a VUMAT. Low-velocity impact on an automotive bumper is simulated using explicit dynamics

FEA. The effect of different design parameters on the results are evaluated and discussed.

7.1 Implementation validation

In Chapter 3, the SDM model was first developed in MATLAB (section 3.1); it was then programmed

into Abaqus using a custom VUMAT subroutine (section 3.3) which implements the SDM model

numerically in the FE software. In this section, the implementation of the SDM model into Abaqus is

validated by comparing the output from FE simulations to the results of the MATLAB implementation.

This was performed by applying a uniaxial in-plane load to a single shell element in the FE software, and

obtaining the stiffness degradations and crack densities in terms of the applied load. In Figure 7-1(a), the

stiffness predicted in MATLAB for a quasi-isotropic GFRP laminate has been plotted. Also plotted are the

results obtained from the implementation of the SDM model into Abaqus. Almost perfect agreement can

be observed. Similarly, the crack densities plotted in subfigure (b) are in good agreement with the

MATLAB numerical implementation of the SDM model.

Next, the VUMAT was augmented to include viscoelastic-viscoplastic properties through Eq. (3.21),

which defines the rate-dependency of the critical energy release rate for crack multiplication (𝐺𝐼𝑐). In

Figure 7-2, the evolution of stiffness in terms of strain has been plotted for different loading rates for a

single shell element in Abaqus loaded under different rates. At high loading rates (more than

3.61%/second), crack multiplication is inhibited due to the increase in critical energy release rate for crack

multiplication, which leads to lower crack densities in all plies of the laminate, and consequently lower

stiffness degradation. It can therefore be concluded that the rate-dependent SDM model has been correctly

implemented into Abaqus using the VUMAT.

Page 123: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

103

(a) (b)

Figure 7-1. Validation of the implementation of the SDM model into Abaqus FEA software using a VUMAT subroutine, as

compared to the MATLAB predictions: (a) Modulus degradation. (b) Crack multiplication in the 90º ply of the [0/90/∓45]s GFRP

composite.

Figure 7-2. Effect of strain rate on stiffness degradation as predicted by the rate-dependent SDM model implemented into

MATLAB for the [0/90/∓45]s GFRP composite.

7.2 Model validation

In order to verify the accuracy of the present damage model, a comparative study is undertaken here.

The results of the present study are compared with those obtained by Schoeppner and Abrate [210]. They

studied the impact response of laminate plates before and after delamination using experimental testing.

In the test, a 12.7 mm × 12.7 mm laminate plate with stacking sequence [0/90]6s was impacted by a

spherical nose impactor made of steel with diameter 25.4 mm, mass 3.1 kg and initial velocity 1.72 m/s

(or initial kinematic energy equal to 4.6 J). The orthotropic properties of the laminate are given in [210].

Figure 7-3 shows time histories of the contact force obtained in the present study alongside the results of

Schoeppner and Abrate. The comparison of total contact time shows a relative discrepancy of about 25%

between the results of the current study and those obtained by Schoeppner and Abrate. Three explanations

can be provided for the differences between the results. First, the impact event takes place in small time

intervals (milliseconds) and the results of Schoeppner and Abrate were obtained using experimental tests.

Therefore, due to the complexity involved in measuring the displacement response experimentally, there

Page 124: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

104

might be small discrepancies between experimental results and computational results. Second, FE

problems with contact definitions such as the impact simulations considered here can involve considerable

numerical inaccuracies due to the complexity of the contact models. Moreover, when defining the

interaction properties in a contact problem, different contact algorithms (i.e. general contact, surface to

surface contact penalty and kinematic methods) can predict different contact forces. Third, the impactor of

present simulation is considered to be a rigid body, however, in the experimental test of [210], the impactor

is made of steel which is a deformable material. Based on these remarks, the differences between

simulation results and the experimental results shown in Figure 7-3 are reasonable.

Figure 7-3. Comparisons of time histories of contact force for a [0/90]6s composite laminate between simulation and experiments

[210].

7.3 Parametric studies

In this section, a parametric study is performed on the effects of the impactor's initial velocity, laminate

stacking sequence, bumper's cross section profile and rate-dependency. Figure 7-4 shows the schematic of

the assembled model of rigid impactor and the composite bumper considered for this study; the model

dimensions are also provided. The material properties are given in Table 7-1. To analyze the effect of the

impactor's initial velocity, three different sets of initial velocities, 4.6 m/s, 8.4 m/s, and 13.1 m/s are

considered. Figure 7-5 shows the time history of contact force for different initial velocities for the

[0/90/∓45]s GFRP composite. It can be found that the velocity has a significant effect on the maximum

contact force, however, the patterns in all results are almost identical. The contact force initially increases

upon impact due to resistance from the bumper. It then decreases as the strain energy of the bumper is

transferred back to the impactor. The oscillations in all three curves are due to the dynamic behavior of the

bumper.

Figure 7-6 shows the evolution of the crack density in each ply of the GFRP [0/90/∓45]s laminate as

a function of time. The crack densities plotted correspond to the maximum values reached throughout the

[210]

Page 125: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

105

bumper. The three different plots correspond to different initial impactor velocities. Looking first to subplot

(a), the crack multiplication process initiates first in the 0° ply, at approximately 0.2 ms after impact. This

suggests that the local laminate stress is mainly transverse at this location. The crack density increases,

until it saturates at 750 cr/m. At 1.6 ms after impact, damage initiates in the ∓45° plies. The crack density

reaches a value 1200 cr/m in the -45° while it only reaches 590 cr/m in the 45° because of its lower

thickness. Crack multiplication initiates in the 90° ply at 2.5 ms, reaching a maximum value of 750 cr/m.

These trends for crack multiplication in the different plies suggest a complex local loading scenario. We

should note that the results of the simulations are written to the output files every 0.24 ms. Therefore, the

results shown in Figure 7-6 contain slight numerical inaccuracies due to the very low time scales at which

crack multiplication evolves, which cannot be exactly captured with this relatively large time step.

When the impactor velocity is increased to 8.4 m/s, cracking initiates sooner in the ∓45º plies. This is

due to the larger increase in local ply transverse stress with increasing velocity, which promotes crack

multiplication. The maximum crack density in all plies increases relative to the simulation with lower

impactor velocity. The same trends can also be observed when the impactor velocity is increased to 13.1

m/s. However, the crack density in the 0º ply does not increase significantly with higher impactor velocity.

This saturation effect can be explained based on the crack multiplication model, in which the interaction

between adjacent cracks prevents further increase in crack density.

The SDM methodology implemented in this chapter is based on shell elements, and therefore

delamination cannot be modelled explicitly using, for example, cohesive zone modelling. However, the

current model provides accurate insights into the evolution of sub-critical damage mechanisms which will

eventually lead to failure of the structure. Understanding the progression of damage in the separate plies

of the different laminates paves the way for failure models of laminates under dynamic impact loading. As

was shown in Figure 7-6, the crack density in the 0° ply saturates, which can accurately be taken into

account through the evolution of COD with respect to increasing crack density. Moreover, the local ply

stresses throughout the bumper induce a complex loading scenario on the different stacking sequences,

which will cause complex patterns of damage evolution. To our knowledge, the SDM model, which uses

fully periodic RVEs for calculating progressive damage evolution, is one of the only models which can

account for these complex multi-axial loading scenarios with minimal experimental input. Taken together,

we believe the observations reported in these figures demonstrate the effectiveness of the damage model.

Figure 7-7 illustrates the contours of 0° ply crack density for the quasi-isotropic GFRP laminate for

different initial velocities of the impactor. As seen in the three plots, the dominant damage occurs in the

region of impact. As the impactor's velocity increases, additional damage is induced in the regions close

to the boundaries. Under a velocity of 4.6 m/s, the maximum crack density in the 0° ply is 710 crack per

meter (cr/m). It reaches 910 cr/m when the impactor velocity is increased to 13.1 m/s. This increase in

Page 126: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

106

crack density is due to the larger in-plane deformation that the impactor causes on the bumper as its velocity

is increased, which drives crack multiplication. We should note that the crack density contours shown in

this figure are envelope plots, which display results at the integration points having the largest crack density

within the shell elements used in this analysis.

Figure 7-4. Schematic of an automotive bumper subject to low-velocity impact. (a) an example of the mesh pattern, (b)

geometrical dimensions of the bumper and impactor.

Table 7-1. Material properties of CFRP and GFRP unidirectional lamina

Material Density Orthotropic properties

CFRP 1550 kg/m3 E1 =157 GPa, E2 = E3 = 8.3 GPa

𝜈12 = 𝜈13 = 0.32 𝜈23 = 0.32 G12=G13 = 5.81 GPa, G23 = 3.14GPa

GFRP 1600 kg/m3 E1 = 45 GPa, E2 = E3= 14.6 GPa

𝜈12 = 𝜈13 = 0.32 𝜈23 = 0.42 G12 = G13 = 4.95 GPa, G23 = 5.14GPa

Page 127: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

107

Figure 7-5. Effect of the impactor’s initial velocity on time histories of the contact force for a [0/90/∓45]s GFRP laminate.

(a) (b)

(c)

Figure 7-6. Time history of the evolution of crack density for the different plies of a [0/90/∓45]s GFRP laminate. Initial velocity

of the impactor is equal to : (a) 4.6 m/s, (b) 8.4 m/s, (c) 13.1 m/s.

Page 128: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

108

Figure 7-7. Contour of 0° ply crack density for [0/90/∓45]s GFRP laminate for different initial velocities of the impactor: (a) v =

4.6 m/s, (b) v = 8.4 m/s, (c) v =13.1 m/s.

(a) (b)

Figure 7-8. Effect of laminate stacking sequence on (a) displacement versus contact force and, (b) time histories of contact force

for GFRP laminates under an impactor velocity of 4.6 m/s.

Figure 7-8 shows the results of displacement versus contact force and time histories of the contact

force for a GFRP with three different laminate stacking sequences under an impactor velocity of 4.6 m/s.

The [0/90/∓45]s laminate has the highest maximum contact force compared to the others due to its higher

stiffness. On the other hand, the [0/90]s laminate has the highest maximum displacement in the contact

region, due its much lower torsional stiffness relative to the other two laminates. Moreover, the shear

stiffness of [0/90]s is significantly degraded under impact due to damage, which explains the lower

contact force. From Figure 7-8(b), it can be seen that the [0/90/∓45]s reaches a larger contact force (30

kN versus 23 kN for [0/90]sand 28 kN for [∓45]s). This is due to the larger stiffness properties of this

stacking sequence relative to the other two laminates.

Page 129: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

109

(a) (b)

Figure 7-9. Effect of laminate stacking sequence on (a) displacement versus contact force, and (b) time histories of contact force

for CFRP laminates under an impactor velocity of 4.6 m/s

Figure 7-9 shows the contact force - displacement curves and the contact force - time curves for three

CFRP stacking sequences under an impactor velocity of 4.6 m/s. The trends observed for the CFRP

material are the same here: the quasi-isotropic laminate induces a larger contact force on the impactor,

while the angle-ply laminate is the least resistant to impact. However, the contact force caused by the

GFRP composites is always lower than for CFRP (see Figure 7-8). For example, the GFRP quasi-isotropic

laminate induces a maximum contact force of 30 kN at 3.5 ms, while the CFRP quasi-isotropic laminate

leads to a maximum force of 60 kN at 3.5 ms. This is due to the much larger in-plane stiffness of CFRP

relative to GFRP.

Table 7-2 shows the maximum amount of damage in each laminate studied in this chapter. The CFRP

composites undergo less damage than the GFRP composites. This is due to the larger in-plane stiffness of

CFRP, which constrains the COD and inhibits stiffness degradation. The [∓45]s composite undergoes the

most stiffness degradation due to its off-axis orientation. The cross-ply and the quasi-isotropic laminate

undergo similar amounts of stiffness degradation, however the shear modulus of the cross-ply is degraded

more significantly due to the absence of ∓45° plies in its stacking sequence.

Figure 7-10 shows the effect of rate-dependence for the [0/90/∓45]s CFRP laminate, based on Eq.

(3.21). Rate-dependence has a slight effect on impact displacement and contact force evolution when the

impactor has an initial velocity of 13.1 m/s. Including rate-dependence causes micro-cracking to be

inhibited at higher loading rates. This leads to larger in-plane stiffness. This explains the slightly different

contact force observed when rate-dependent cracking is accounted for. However, due to the low loading

rates involved in low velocity impact, this effect is negligible, thus warranting the use of the simpler rate-

independent crack multiplication model developed in this chapter.

Page 130: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

110

Table 7-2. Maximum degradation of material properties of laminates with different stacking sequences due to low-velocity impact.

Laminate Ex Ey Gxy 𝜈xy

CFRP, [0/90/∓45]s 2.8% 4.6% 3.4% +2.6 %

CFRP, [0/90]s 2.7% 3.7% 42.3% -54.8 %

CFRP, [0/90/∓60]s 4.5% 2.0% 6.4% -2.2%

CFRP, [∓45]s 74% 74% 2.7% +22 %

GFRP, [0/90/∓45]s 13.0 % 17.6 % 14.2 % +1.5 %

GFRP, [0/90]s 17.2 % 18.2% 40.4 % -64.1 %

GFRP, [0/90/∓60]s 19.6% 10.6% 16.7% -18.2%

GFRP, [∓45]s 72.9 % 72.9 % 16 % +40%

(a) (b)

Figure 7-10. Effect of rate-dependency on (a) displacement versus contact force, (b) time histories of contact force for a CFRP

[0/90/∓45]s laminate.

Figure 7-11 shows the crack density contours for the [0/90/∓60]s CFRP laminate with three different

bumper cross-sections when the impactor velocity is 8.4 m/s. The first cross-section is square, and the

maximum crack density in the 0° ply reaches 1000 cr/m. The second cross-section is curved, and shows a

maximum crack density of 820 cr/m, while the last cross-section contains grooves and the crack density

reaches 980 cr/m. These differences between the different bumpers can be attributed to the shapes of the

cross-section, which induce a different stress distribution upon impact, and therefore a different driving

force for crack multiplication. The first bumper is more susceptible to damage evolution, because of the

irregular stress state induced on its surface. These results therefore suggest that damage tolerance in

composite car bumpers is related to the shape of the cross-section, which can be accounted for using the

methodology presented in this chapter.

Page 131: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

111

Figure 7-11. Effect of the cross section profile of the bumper on 0º crack density contours for a [0/90/∓60]s CFRP laminate under

an impactor velocity of 8.4 m/s.

Figure 7-12 shows the effect of cross-section on the contact force and displacement of the bumper for

the [0/90/∓60]s CFRP laminate at an impactor velocity of 8.4 m/s. The overall responses for all three cross-

sections is very similar. There are some significant differences in the mechanical responses of the different

bumpers: the bumper with a humped cross-sectional profile is much more rigid, leading to a maximum

force of 60 kN on the impactor. The other two types of bumpers are more compliant. This can be attributed

to the different distribution of damage in the bumper upon impact, as was illustrated in Figure 7-11. It can

also be attributed to the different stress distribution occurring in the different bumpers, which induces a

different load-displacement response, even when neglecting the effects of damage. This suggests that

cross-sectional shape partly governs the homogenized mechanical behavior of the bumper and that it has

an effect on the damage evolution patterns throughout the structure.

(a) (b)

Figure 7-12. Effect of bumper cross-section shapes on (a) displacement versus contact force and, (b) time histories of contact

force for CFRP laminates for the [0/90/∓60]s laminate.

As shown in Figure 7-10, the effect of including a rate-dependent crack multiplication criterion to

represent the viscoelastic-viscoplastic behavior of laminates did not lead to significant differences in the

Page 132: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

112

load-displacement curve at the location of impact. In order to more fully understand rate-dependent effects

in this structure, the crack density in the 0° ply has been plotted in Figure 7-13, by ignoring rate-

dependency on the one hand, and by including it on the other hand. Clearly, even though the rate-dependent

behavior did not lead to significant changes in the load-displacement curves, the crack density distribution

is significantly affected: the maximum reaches 910 cr/m when rate-dependency is ignored, while it only

reaches 765 cr/m when rate-dependency is accounted for. Moreover, damage is less extensive throughout

the structure and covers a more limited area. This can be explained based on the increase in energy required

to increase crack density when rate-dependency is included, which reduces crack density. This effect is

more clearly shown in Figure 7-14, where the crack density in the 0° ply has been plotted. Clearly, rate-

dependency has a large effect on crack density evolution. This effect therefore needs to be accounted for

a proper prediction of damage evolution. However, the resulting effect on the contact force-displacement

curve is low due to the restricted area over which micro-cracking extends.

(a) (b)

Figure 7-13. Crack density in the 0° ply of a GFRP [0/90/∓45]s laminate under low-velocity impact at 13.1 m/s : (a) neglecting

rate-dependent behavior and (b) including rate-dependent behavior.

Figure 7-14. Effect of including rate-dependency on crack density evolution in the 0° ply of the [0/90/∓45]s GFRP bumper under

low velocity impact.

Page 133: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

113

Summary for Chapter 7.

In this chapter, a robust and accurate damage model considering matrix micro-cracking in laminates

has been used to predict the low-velocity impact response of a car bumper. FE micro-damage modelling

is used to obtain the parameters of the damage model, in lieu of extensive and complicated experiments

that have traditionally been used. The model is then implemented into Abaqus using a VUMAT. The

impact response of different stacking sequences and different material systems is simulated using explicit

dynamics in Abaqus. The results show the following:

GFRP laminates undergo more extensive deformation than the CFRP laminates, due to their lower

in-plane stiffness, and more extensive damage evolution. The larger damage susceptibility of

GFRP can be explained from the lower stiffness of the plies relative to CFRP, which limits the

constraining effects of adjacent plies on the Crack Opening Displacements.

It was found that for both material systems, the quasi-isotropic laminate underwent less

deformation upon impact, and induced a larger contact force on the impactor; these observations

were attributed to the larger bending stiffness of this stacking sequence.

The effect of incorporating a rate-dependent crack multiplication model was also considered and

showed that rate-dependency only had some effect of damage evolution, but an insignificant effect

on the load-displacement response of the bumper.

Lastly, the effect of bumper cross-section shape was studied, and it was found that the damage

evolution of composite car bumpers is sensitive to geometry.

In summary, the results of the parametric studies demonstrate the ability of the composite damage

model to predict micro-cracking in the different plies of GFRP and CFRP laminates with different stacking

sequences and cross-sectional shapes, as well as the mechanical response of damaged composites under

dynamic loading.

The focus of this chapter has been on low-velocity impact of laminates, which is important to

understand in order to predict damage tolerance following collision of a thin composite with an impactor.

However, in the case of vehicles colliding at very high speeds, such as might be expected practically, the

modelling approach developed in this chapter would not be applicable, due to the onset of critical damage

modes such as delamination and fiber fracture, which are not accounted for in this model. Future works

should attempt to extend the model to these critical damage modes, and predict the interactions between

matrix micro-cracking, delamination and fiber fracture. Such a complete model would help elucidate the

progression of damage that unfolds during a crash event and help in the design of failure-tolerant

automotive structures [211]. Furthermore, the model implemented in this chapter is concerned with thin

structures, which can be modelled using shell elements; a rate-independent version of this model has been

Page 134: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

114

used before [212], [213] for wind turbine blades. Thicker structures would require a different formulation

of the SDM model which accounts for out-of-plane behavior. Such improvements are left to future works.

In the case of short fiber composites, for example bio-composites, the low-velocity impact response

will be different due to the different evolution of damage processes. The in-plane mechanical properties of

the material will be isotropic due to the random orientation of the fibers. Damage will nucleate at several

locations in the structure due to the stress concentrations effected by the fiber tips. The diffuse damage

will grow with increasing loading, and eventually lead to the coalescence of matrix micro-cracks and the

nucleation of a macroscopic crack, resulting in part failure. The exact shape of the contact force response

might be different from that of the laminate due to the elastic isotropy, but the main difference between

the two types of composites will be the evolution of damage.

Page 135: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

115

Chapter 8. Conclusions and future work

8.1. Conclusions

In this thesis, computational modelling tools have been developed to study the effects of the

viscoelasticity and viscoplasticity of polymer-based composites on the nature and evolution of damage

prior to failure. The investigation has spanned from microscopic damage modelling of debonding in short

fiber composites, to ply-level micro-crack modelling in continuous fiber composites, and up to the

structural level, where damage was predicted in an automotive composite structure. A wide range of

computational techniques have been used to complete the objectives of this thesis.

In Chapters 4 and 5 of this thesis, micro-damage modelling of matrix micro-cracking in symmetric

multi-directional laminates has been used to calibrate the equations of a Synergistic Damage Mechanics

model to predict stiffness degradation, as well as micro-crack multiplication in viscoelastic-viscoplastic

composites. An algorithm based on Classical Laminate Plate Theory has been implemented to predict the

bi-axial creep of laminates, and combined with the SDM model to predict the evolution of damage during

creep deformation. Excellent agreement of the predictions of the Continuum Damage Model with

experimental data has been obtained. The effect of viscoplasticity has also been considered in Chapter 5.

Parametric studies have been conducted to understand the effects of biaxial loading ratios, fiber type (glass

and carbon), stacking sequence and temperature. The strengths of the model include its straightforward

implementation for a large range of stacking sequences, its ability to accurately account for different multi-

axial in-plane loading scenarios, and its minimal reliance on experimental data. Therefore, objective 1 of

this thesis has been addressed. The key findings of the two chapters based on this multi-scale model are as

follows:

The evolution of creep strain under constant loading could be captured using the damage

model, as shown in Chapter 4 in Figure 4-2. Therefore, matrix micro-cracking can be assumed

to only affect the elastic properties of the laminate during creep deformation, which results in

a relatively simple constitutive model for the creep of multi-directional laminates.

The time-dependent behavior of the matrix mainly affects damage evolution through the

decrease in critical energy release rate for crack multiplication (𝐺𝐼𝑐) with respect to time. The

evolution of stress in the plies during the creep test does not significantly affect crack

multiplication, as shown in Figure 5-12.

Creep deformation shifts the stiffness degradation envelopes to lower values of stress, as

shown in Figure 5-8-Figure 5-10. These curves can be used to guide the design of composites

used in demanding thermo-mechanical applications

Page 136: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

116

Temperature has a slight effect on damage evolution, which has been taken into account by

using temperature-dependent ply stiffness values and time-dependent critical energy release

rates for crack multiplication.

The multi-scale damage model has been implemented into FE software with a user-defined subroutine

in Chapter 7. The effect of low-velocity impact on the dynamic evolution of matrix micro-cracking has

been studied for different stacking sequences, material systems and impactor velocities. It has been found

that some stacking sequences are more prone to damage than others, as shown by the more significant

stiffness degradation levels: the angle-ply underwent the most stiffness degradation, reaching more than

73% stiffness loss when the GFRP was considered, and 74% for the CFRP. The cross-ply underwent

significant loss in shear modulus (more than 40%), while the quasi-isotropic laminates were the most

resistant to damage and underwent the least deformation. The evolution of matrix micro-cracking has been

quantified during the impact event while taking into account complex multi-axial loading scenarios. A

loading-rate dependent crack multiplication model has been implemented, and it has been shown that strain

rate has an effect on damage evolution, although the subsequent effect on the load-displacement response

of the bumper is not very significant. Objective 3 of this thesis has therefore also been completed.

Lastly, in Chapter 6, a new micro-damage model for short fiber composites has been developed to

predict debonding evolution under different strain rates. It models the behavior of the interface using a

bilinear Cohesive Zone Model, and the matrix non-linearity is accounted for using a viscoelastic model. A

range of strain rates has been studied, from 102 s-1 to 10-5 s-1. The effects of fiber diameter, fiber type and

interfacial strengths have been studied as well, providing new insights into the evolution of micro-damage

in short fiber composites with a ductile matrix, as follows:

The viscoelastic behavior of thermoplastic matrices leads to a transition from debonding to

interfacial crack opening with decreasing strain rate. For a 40 μm fiber, at 4% strain, the

debonded area fraction goes from 95% debonding at 102 s—1 strain rate, to less than 30 %

under a strain rate of 10-5 s-1.

Higher interfacial strength lead to a higher likelihood of matrix cracking, and a lower

likelihood of debonding: at 30 MPa interfacial shear strength, the strain energy density reached

in the matrix 110 MPa at 4% strain, while it only reached 40 MPa at 11 MPa interfacial shear

strength. Debonding was also less extensive with increasing interfacial shear strength.

Thinner fibers are less prone to debonding, and are more likely to lead to interfacial crack

opening: for a 40 μm diameter fiber, the debonded area fraction reaches 30 % at a strain rate

of 10-5 s-1, while for a 10 μm diameter fiber, it only reaches 15 %.

Page 137: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

117

Hemp fibers are less susceptible to debonding, even at high strain rates, as compared to the

glass and carbon fibers.

The central focus of this thesis has been on the effect of viscoelastic and viscoplastic properties on

damage evolution in composites. These properties relate to the time-dependency of loading, and have been

studied in several ways: constant loading, in which these properties will lead to creep deformation; quasi-

static loading under different strain rates, in which viscoelasticity and viscoplasticity will lead to stress

relaxation and different shapes of the stress-strain curve; and dynamic loading, where the viscoelastic-

viscoplastic properties lead to different material responses, as well as different patterns of damage

evolution due to the rate-dependent failure of the matrix.

8.2 Future work

Based on the findings of this thesis, the following recommendations for future work can be made:

A fully non-linear FE model of matrix micro-cracking should be developed to more accurately

represent the effect of matrix viscoelasticity and viscoplasticity on damage evolution. In this thesis,

the SDM model was based on RVEs with elastic plies, and the effect of matrix micro-cracking on

viscoelastic and viscoplastic compliance were not modelled explicitly. Future work should

improve the accuracy of the model used in Chapters 4 and 5 of this thesis to apply to more complex

loading scenarios, such as dynamic loading, and fatigue. Moreover, the relationship between the

fracture toughness of the matrix, and the critical energy release rate for crack multiplication (𝐺𝐼𝑐)

should be understood, in order to by-pass the complicated set of experiments required to obtain

this parameter.

A more in-depth analysis of micro-damage mechanisms in short fiber composites is required. The

effect of surrounding fibers on the evolution of debonding and matrix cracking needs to be

evaluated in more depth so as to provide more detailed insight into the failure of short fiber

composites. This could be done by augmenting the RVE used in Chapter 6 to include multiple

fibers with different lengths and orientations. A more sophisticated model accounted for the

failure of the polymeric matrix should also be developed to account for the competition between

damage modes, especially for strong interfaces and brittle matrices, which are more likely to

undergo cracking.

The user-defined material subroutine should be extended to account for rate-dependent behavior

at high-temperatures. In this thesis, viscoelastic and viscoplastic behavior were accounted for by

Page 138: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

118

using a rate-dependent critical energy release rate for crack multiplication, which is only valid at

room temperature. At high temperatures, the non-linear properties of the plies will also lead to

complex mechanical behavior. This should be accounted for in future investigations, so as to fully

extend the capabilities of the SDM model.

In conclusion, the results of this thesis demonstrate the accuracy of FE modelling of damage

mechanisms, and also provide key results and fundamental insights into damage evolution, at multiple

scales. The methodologies and results can be used for the design of composite components in a wide range

of applications.

Page 139: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

119

REFERENCES

[1] D. L. Greene and A. Schafer, “Reducing Greenhouse Gas Emissions From U.S. Transportation.”

Pew Center on Global Climate Change, pp. 1–58, 2003.

[2] Ruth Heuss, N. Müller, W. van Sintern, A. Starke, and A. Tschiesner, “Lightweight, heavy impact

- How carbon fiber and other lightweight materials will develop across industries and specifically

in automotive.” McKinsey & Company, pp. 1–24, 2012.

[3] “Learn the facts: Weight affects fuel consumption,” Natural Resources Canada, 2017. [Online].

Available: http://www.nrcan.gc.ca/energy/efficiency/transportation/cars-light-

trucks/buying/16755. [Accessed: 07-Oct-2017].

[4] H. C. Kim and T. J. Wallington, “Life-cycle energy and greenhouse gas emission benefits of

lightweighting in automobiles: Review and harmonization,” Environ. Sci. Technol., vol. 47, no. 12,

pp. 6089–6097, 2013.

[5] J. C. Kelly, J. L. Sullivan, A. Burnham, and A. Elgowainy, “Impacts of Vehicle Weight Reduction

via Material Substitution on Life-Cycle Greenhouse Gas Emissions,” Environ. Sci. Technol., vol.

49, no. 20, pp. 12535–12542, 2015.

[6] C. T. Herakovich, “Mechanics of Fibrous Composites.” John Wiley & Sons, Inc., 1998.

[7] Magna, “Magna and Ford Collaborate on Multi-Material Lightweight Vehicle Concept Car,” 2014.

[Online]. Available:

https://www.magna.com/company/newsroom/releases/release/2014/06/11/news-release---magna-

and-ford-collaborate-on-multi-material-lightweight-vehicle-concept-car. [Accessed: 13-Dec-

2018].

[8] J. L. Hardcastle, “GM to Cut Vehicle Weight ‘Up to 15%’ by 2017,” Environmental Leader, 2013.

[9] E. V. Olsen and H. G. Lemu, “Mechanical Testing of Composite Materials for Monocoque Design

in Formula Student Car,” Int. Sch. Sci. Res. Innov., vol. 10, no. 1, pp. 1–8, 2016.

[10] EPA, “Sources of Greenhouse Gas Emissions,” 2017. [Online]. Available:

https://www.epa.gov/ghgemissions/sources-greenhouse-gas-emissions. [Accessed: 02-Oct-2017].

Page 140: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

120

[11] S. Mazumdar, M. Benevento, D. Pichler, E. Witten, and J. Hinrichsen, “The 2018 State of the

Composites Industry Report,” Composites Manufacturing, 2018.

[12] S. Mazumdar, “Opportunity and Challenges in Automotive Composites Industry.” Lucintel, 2013.

[13] F. Vilaplana, E. Strömberg, and S. Karlsson, “Environmental and resource aspects of sustainable

biocomposites,” Polym. Degrad. Stab., vol. 95, no. 11, pp. 2147–2161, Nov. 2010.

[14] A. K. Mohanty, M. Misra, and L. T. Drzal, “Sustainable Bio-Composites from Renewable

Resources : Opportunities and Challenges in the Green Materials World,” J. Polym. Environ., vol.

10, no. April, pp. 19–26, 2002.

[15] J. Varna, E. Sparninš, R. Joffe, K. Nättinen, and J. Lampinen, “Time-dependent behavior of

flax/starch composites,” Mech. Time-Dependent Mater., vol. 16, no. 1, pp. 47–70, 2012.

[16] O. Faruk, A. K. Bledzki, H.-P. Fink, and M. Sain, “Biocomposites reinforced with natural fibers:

2000–2010,” Prog. Polym. Sci., vol. 37, no. 11, pp. 1552–1596, Nov. 2012.

[17] A. Soroudi and I. Jakubowicz, “Recycling of bioplastics, their blends and biocomposites: A

review,” Eur. Polym. J., vol. 49, no. 10, pp. 2839–2858, 2013.

[18] A. Arbor, “The Bio-Based Materials Automotive Value Chain,” no. February, 2012.

[19] “Recyclable materials,” Green Vehicle Disposal, 2017. [Online]. Available:

http://www.greenvehicledisposal.com/environment/recyclable-materials/#. [Accessed: 10-Feb-

2017].

[20] H. Hajiha and M. Sain, “High toughness hybrid biocomposite process optimization,” Compos. Sci.

Technol., vol. 111, pp. 44–49, 2015.

[21] “Composites Market Size, Share & Trends, Global Industry Report, 2024,” Grand View Research,

2017.

[22] Going to Extremes: Meeting the Emerging Demand for Durable Polymer Matrix Composites.

National Research Council, 2005.

[23] D. A. Dillard and H. F. Brinson, “A Numerical Procedure for Predicting Creep and Delayed

Page 141: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

121

Failures in Laminated Composites,” ASTM, pp. 23–37, 1983.

[24] M. E. Tuttle, A. Pasricha, and A. F. Emery, “The Nonlinear Viscoelastic-Viscoplastic Behaviour

of IM7/5260 Composites Subjected to Cyclic Loading,” J. Compos. Mater., vol. 29, no. 15, pp.

2025–2046, 1995.

[25] A. Pramanick and M. Sain, “Nonlinear Viscoelastic Creep Prediction of HDPE-Agro-fiber

Composites,” J. Compos. Mater., vol. 40, no. 5, pp. 417–431, 2005.

[26] A. Asadi, “A model for time-independent and time-dependent damage evolution and their influence

on creep of multidirectional polymer composite laminates,” University of Manitoba, 2013.

[27] E. Marklund, J. Eitzenberger, and J. Varna, “Nonlinear viscoelastic viscoplastic material model

including stiffness degradation for hemp/lignin composites,” Compos. Sci. Technol., vol. 68, no. 9,

pp. 2156–2162, 2008.

[28] Z. Jendli, J. Fitoussi, F. Meraghni, and D. Baptiste, “Anisotropic strain rate effects on the fibre-

matrix interface decohesion in sheet moulding compound composites,” Compos. Sci. Technol., vol.

65, no. 3–4, pp. 387–393, 2005.

[29] Z. Jendli, F. Meraghni, J. Fitoussi, and D. Baptiste, “Multi-scales modelling of dynamic behaviour

for discontinuous fibre SMC composites,” Compos. Sci. Technol., vol. 69, no. 1, pp. 97–103, 2009.

[30] Z. Vinogradov, V., Hashin, “Probabilistic energy based model for prediction of transverse cracking

in cross-ply laminates,” Int. J. Solids Struct., vol. 42, no. 2, pp. 365–392, 2005.

[31] J. A. Nairn, “Matrix Microcracking in Composites,” Polym. Matrix Compos., vol. 2, p. 34, 2001.

[32] A. Asadi and J. Raghavan, “Influence of time-dependent damage on creep of multidirectional

polymer composite laminates,” Compos. Part B Eng., vol. 42, no. 3, pp. 489–498, Apr. 2011.

[33] J. Tong, F. J. Guild, S. L. Ogin, and P. A. Smith, “On Matrix Crack Growth in Quasi-Isotropic

Laminates-I. Experimental Investigation,” Compos. Sci. Technol., vol. 51, no. 97, pp. 1537–1545,

1997.

[34] L. N. McCartney, “Predicting transverse crack formation in cross-ply laminates,” Fail. Criteria

Page 142: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

122

Fibre-Reinforced-Polymer Compos., vol. 58, pp. 157–178, 2004.

[35] J. Varna, R. Joffe, N. V. Akshantala, and R. Talreja, “Damage in composite laminates with off-axis

plies,” Compos. Sci. Technol., vol. 59, no. 14, pp. 2139–2147, 1999.

[36] P. Lundmark and J. Varna, “Constitutive Relationships for Laminates with Ply Cracks in In-plane

Loading,” Int. J. Damage Mech., vol. 14, no. 3, pp. 235–259, 2005.

[37] P. Wang, J. Liu, W. Yu, and C. Zhou, “Dynamic rheological properties of wood polymer

composites: from linear to nonlinear behaviors,” Polym. Bull., vol. 66, no. 5, pp. 683–701, Oct.

2010.

[38] L. Rozite, J. Varna, R. Joffe, and a. Pupurs, “Nonlinear behavior of PLA and lignin-based flax

composites subjected to tensile loading,” J. Thermoplast. Compos. Mater., vol. 26, no. 4, pp. 476–

496, Nov. 2011.

[39] M. Sain, S. Panthapulakkal, and S. Law, “Manufacturing process for hybrid organic and inorganic

fibre-filled composite materials,” 2009.

[40] S. Panthapulakkal and M. Sain, “Injection-Molded Short Hemp Fiber / Glass Fiber- Reinforced

Polypropylene Hybrid Composites — Mechanical , Water Absorption and Thermal Properties,”

vol. 103, pp. 2432–2441, 2006.

[41] R. Talreja and C. V. Singh, Damage and Failure of Composite Materials. Cambridge University

Press, 2012.

[42] H. Ku, H. Wang, N. Pattarachaiyakoop, and M. Trada, “A review on the tensile properties of natural

fiber reinforced polymer composites,” Compos. Part B Eng., vol. 42, no. 4, pp. 856–873, 2011.

[43] D. S. Dandy, “What is Chemical Engineering?,” Colorado State University, 2007. [Online].

Available: http://navier.engr.colostate.edu/whatische/ChEL05Body.html. [Accessed: 04-Oct-

2017].

[44] R. M. Koide, G. V. Z. De Franca, and M. A. Luersen, “An Ant Colony Algorithm Applied to Lay-

up Optimization of Laminated Composite Plates,” Lat. Am. J. Solids Struct., vol. 10, pp. 491–504,

Page 143: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

123

2013.

[45] C. V. Singh and R. Talreja, “A synergistic damage mechanics approach for composite laminates

with matrix cracks in multiple orientations,” Mech. Mater., vol. 41, no. 8, pp. 954–968, 2009.

[46] C. V. Singh, “Multiscale modeling of damage in multidirectional composite laminates,” Texas

A&M University, 2008.

[47] “Accelerated Aging of Materials and Structures: The Effects of Long-Term Elevated-Temperature

Exposure.” The National Academies Press, pp. 1–65, 1996.

[48] J. Tong, F. J. Guild, S. L. Ogin, and P. A. Smith, “On matrix crack growth in quasi-isotropic

laminates—I. Experimental investigation,” Compos. Sci. Technol., vol. 57, no. 11, pp. 1527–1535,

1997.

[49] A. Parvizi and J. E. Bailey, “On multiple transverse cracking in glass fibre epoxy cross-ply

laminates,” J. Mater. Sci., vol. 13, no. 10, pp. 2131–2136, 1978.

[50] J. E. Bailey, P. T. Curtis, and A. Parvizi, “On the transverse cracking and longitudinal splitting

behaviour of glass and carbon fibre reinforced epoxy cross ply laminates and the effect of Poisson

and thermally generated strain,” Proc. R. Soc. London, vol. 366, pp. 599–623, 1979.

[51] D. L. Flaggs and M. H. Kural, “Experimental Determination of the In Situ Transverse Lamina

Strength in Graphite/Epoxy Laminates,” J. Compos. Mater., vol. 16, no. 2, pp. 103–116, 1982.

[52] P. A. Smith, L. Boniface, and N. F. C. Glass, “A Comparison of Transverse Cracking Phenomena

in (0/90) s and (90/0) s CFRP Laminates,” Appl. Compos. Mater., vol. 5, no. 1, pp. 11–23, 1998.

[53] S. Liu and J. A. Nairn, “The Formation and Propagation of Matrix Microcracks in Cross-Ply

Laminates during Static Loading,” J. Reinf. Plast. Compos., vol. 11, no. 2, pp. 158–178, 1992.

[54] O. Rubenis, E. Sparniņš, J. Andersons, and R. Joffe, “The effect of crack spacing distribution on

stiffness reduction of cross-ply laminates,” Appl. Compos. Mater., vol. 14, no. 1, pp. 59–66, 2007.

[55] P. W. Manders, T. W. Chou, F. R. Jones, and J. W. Rock, “Statistical analysis of multiple fracture

in 0°/90°/0° glass fibre/epoxy resin laminates,” J. Mater. Sci., vol. 18, no. 10, pp. 2876–2889, 1983.

Page 144: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

124

[56] J. A. Nairn, S. Hu, and J. S. Bark, “A critical evaluation of theories for predicting microcracking in

composite laminates,” J. Mater. Sci., vol. 28, no. 18, pp. 5099–5111, 1993.

[57] K. W. Garrett and J. E. Bailey, “Multiple transverse fracture in 90 cross-ply laminates of a glass

fibre-reinforced polyester,” J. Mater. Sci., vol. 12, no. 1, pp. 157–168, 1977.

[58] J. A. Nairn and S. Hu, “The initiation and growth of delaminations induced by matrix microcracks

in laminated composites,” Int. J. Fract., vol. 57, no. 1, pp. 1–24, 1992.

[59] C. V. Singh and R. Talreja, “A synergistic damage mechanics approach for composite laminates

with matrix cracks in multiple orientations,” Mech. Mater., vol. 41, no. 8, pp. 954–968, 2009.

[60] D. Hull and Y. B. Shi, “Damage mechanism characterization in composite damage tolerance

investigations,” Compos. Struct., vol. 23, no. 2, pp. 99–120, 1993.

[61] N. Sato, T. Kurauchi, S. Sato, and O. Kamigaito, “Microfailure behaviour of randomly dispersed

short fibre reinforced thermoplastic composites obtained by direct SEM observation,” J. Mater.

Sci., vol. 26, no. 14, pp. 3891–3898, 1991.

[62] S. Fu, B. Lauke, and M. A. Edith, “Hybrid effects on tensile properties of hybrid short-glass-fiber-

and short-carbon-fiber-reinforced polypropylene composites,” J. Mater. Sci., vol. 36, pp. 1243–

1251, 2001.

[63] M. F. Arif, F. Meraghni, Y. Chemisky, N. Despringre, and G. Robert, “In situ damage mechanisms

investigation of PA66/GF30 composite: Effect of relative humidity,” Compos. Part B Eng., vol.

58, pp. 487–495, 2014.

[64] J. Lindhagen and L. Berglund, “Microscopical damage mechanisms in glass fiber reinforced

polypropylene,” J. Appl. Polym. Sci., vol. 69, pp. 1319–1327, 1998.

[65] E. Belmonte, M. De Monte, C. J. Hoffmann, and M. Quaresimin, “Damage initiation and evolution

in short fiber reinforced polyamide under fatigue loading: Influence of fiber volume fraction,”

Compos. Part B Eng., vol. 113, pp. 331–341, 2017.

[66] S. Marfia and E. Sacco, “TFA-based Homogenization for Composites Subjected to Coupled

Page 145: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

125

Damage-friction Effects,” Procedia Eng., vol. 109, pp. 113–120, 2015.

[67] MT-Aerospace AG, “Fracture surface of a CVI-SiC/SiC-material (fibre reinforced ceramics with

SiC-fibres and SiC-matrix, manufactured via a CVI-process (Chemical Vapour Infiltration)),”

2006.

[68] “Composite Release Films and Mould Release Agents,” LPD Lab Services, 2017. [Online].

Available:

https://www.lpdlabservices.co.uk/problem_solving/typical_technical_expertise/composite_release

_films_and_mould_release_agents.php. [Accessed: 08-Oct-2017].

[69] X. Hu, J. Fang, F. Xu, B. Dong, Y. Xiao, and L. Wang, “Real internal microstructure based key

mechanism analysis on the micro-damage process of short fibre-reinforced composites,” Sci. Rep.,

vol. 6, no. June, p. 34761, 2016.

[70] S. S. Wang and E. S. M. Chim, “Fatigue Damage and Degradation in Random Short-Fiber SMC

Composite,” J. Compos. Mater., vol. 17, no. 2, pp. 114–134, 1983.

[71] A. C. Johnson, S. A. Hayes, and F. R. Jones, “The role of matrix cracks and fibre/matrix debonding

on the stress transfer between fibre and matrix in a single fibre fragmentation test,” Compos. Part

A Appl. Sci. Manuf., vol. 43, no. 1, pp. 65–72, 2012.

[72] E. Marklund, J. Varna, and L. Wallström, “Nonlinear Viscoelasticity and Viscoplasticity of

Flax/Polypropylene Composites,” J. Eng. Mater. Technol., vol. 128, no. 4, p. 527, 2006.

[73] M. Lévesque, K. Derrien, D. Baptiste, and M. D. Gilchrist, “On the development and parameter

identification of Schapery-type constitutive theories,” Mech. Time-Dependent Mater., vol. 12, no.

2, pp. 95–127, Mar. 2008.

[74] S. W. Park and R. a. Schapery, “A viscoelastic constitutive model for particulate composites with

growing damage,” Int. J. Solids Struct., vol. 34, no. 8, pp. 931–947, 1997.

[75] A. Pramanick and M. Sain, “Temperature-Stress Equivalency in Nonlinear Viscoelastic Creep

Characterization of Thermoplastic/Agro-fiber Composites,” J. Thermoplast. Compos. Mater., vol.

Page 146: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

126

19, no. 1, pp. 35–60, 2006.

[76] M. M. Sain, J. Balatinecz, and S. Law, “Creep fatigue in engineered wood fiber and plastic

compositions,” J. Appl. Polym. Sci., vol. 77, no. 2, pp. 260–268, 2000.

[77] Y. C. Lou and R. A. Schapery, “Viscoelastic characterization of a nonlinear fiber-reinforced

plastic,” J. Compos. Mater., vol. 5, pp. 208–234, 1971.

[78] A. Gupta, “Creep of plain weave polymer matrix composites,” University of Manitoba, 2009.

[79] Z. Hashin, “Complex moduli of viscoelastic composites-I. General theory and application to

particulate composites,” Int. J. Solids Struct., vol. 6, no. 5, pp. 539–552, 1970.

[80] S. E. Hamel, N. J. Bechle, J. C. Hermanson, and S. M. Cramer, “Characterizing Macroscopic Creep

Behavior of Wood-Plastic Composites in Tension and Compression,” pp. 139–144, 2009.

[81] S. E. Hamel, J. C. Hermanson, and S. M. Cramer, “Mechanical and time-dependent behavior of

wood-plastic composites subjected to tension and compression,” J. Thermoplast. Compos. Mater.,

vol. 26, no. 7, pp. 968–987, 2013.

[82] P. Wang, J. Liu, W. Yu, and C. Zhou, “Dynamic rheological properties of wood polymer

composites: From linear to nonlinear behaviors,” Polym. Bull., vol. 66, no. 5, pp. 683–701, 2011.

[83] R. M. Guedes, “An energy criterion to predict delayed failure of multi-directional polymer matrix

composites based on a non-linear viscoelastic model,” Compos. Part A Appl. Sci. Manuf., vol. 35,

no. 5, pp. 559–571, 2004.

[84] R. K. Abu Al-Rub, A. H. Tehrani, and M. K. Darabi, “Application of a large deformation nonlinear-

viscoelastic viscoplastic viscodamage constitutive model to polymers and their composites,” Int. J.

Damage Mech., vol. 24, no. 2, pp. 198–244, 2014.

[85] D. A. Dillard, D. H. Morris, and H. F. Brinson, “Creep and Creep Rupture of laminated

graphite/epoxy composites,” 1980.

[86] R. H. Moore and D. A. Dillard, “Time-dependent matrix cracking in cross-ply laminates,” Compos.

Sci. Technol., vol. 39, no. 1, pp. 1–12, 1990.

Page 147: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

127

[87] A. Birur, A. Gupta, and J. Raghavan, “Creep Rupture of Multidirectional Polymer Composite

Laminates — Influence of Time-Dependent Damage,” J. Eng. Mater. Technol., vol. 128, no. 4, pp.

611–617, 2006.

[88] A. Birur, “Time-dependent damage evolution in multidirectional polymer composite laminates,”

University of Manitoba, 2008.

[89] T. H. Nguyen and D. Gamby, “Effects of nonlinear viscoelastic behaviour and loading rate on

transverse cracking in CFRP laminates,” Compos. Sci. Technol., vol. 67, no. 3–4, pp. 438–452,

2007.

[90] J. Fitoussi, M. Bocquet, and F. Meraghni, “Effect of the matrix behavior on the damage of ethylene-

propylene glass fiber reinforced composite subjected to high strain rate tension,” Compos. Part B

Eng., vol. 45, no. 1, pp. 1181–1191, 2013.

[91] S. G. Lim and C. S. Hong, “Prediction of Transverse Cracking and Stiffness Reduction in Cross-

Ply Laminated Composites,” J. Compos. Mater., vol. 23, no. 7, pp. 695–713, 1989.

[92] Z. Hashin, “Analysis of cracked laminates: a variational approach,” Mech. Mater., vol. 4, no. 2, pp.

121–136, 1985.

[93] E. Adolfsson and P. Gudmundson, “Thermoelastic properties in combined bending and extension

of thin composite laminates with transverse matrix cracks,” Int. J. Solids Struct., vol. 34, no. 16,

pp. 2035–2060, 1997.

[94] J. Montesano and C. V. Singh, “A synergistic damage mechanics based multiscale model for

composite laminates subjected to multiaxial strains,” Mech. Mater., vol. 83, pp. 72–89, 2015.

[95] J. Montesano and C. V. Singh, “Critical stiffness damage envelopes for multidirectional laminated

structures under multiaxial loading conditions,” Mater. Des., vol. 91, pp. 218–229, 2016.

[96] J. Varna, R. Joffe, and R. Talreja, “A synergistic damage-mechanics analysis of transverse cracking

in [±θ/904]s laminates,” Compos. Sci. Technol., vol. 61, no. 5, pp. 657–665, 2001.

[97] R. Talreja, “Transverse Cracking and Stiffness Reduction in Composite Laminates,” J. Compos.

Page 148: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

128

Mater., vol. 19, no. 4, pp. 355–375, 1985.

[98] R. Talreja, “A Continuum Mechanics Characterization of Damage in Composite Materials,” Proc.

R. Soc. A Math. Phys. Eng. Sci., vol. 399, no. 1817, pp. 195–216, 1985.

[99] P. Gudmundson and S. Ostlund, “First Order Analysis of Stiffness Reduction Due to Matrix

Cracking,” J. Compos. Mater., vol. 26, no. 7, pp. 1009–1030, 1992.

[100] P. Gudmundson and W. Zang, “An Analytic Model For Thermoelastic Properties of Composite

Laminates,” Int. J. Solids Struct., vol. 30, no. 23, pp. 3211–3231, 1993.

[101] P. Lundmark and J. Varna, “Crack face sliding effect on stiffness of laminates with ply cracks,”

Compos. Sci. Technol., vol. 66, no. 10, pp. 1444–1454, 2006.

[102] J. Varna, “Modelling mechanical performance of damaged laminates,” J. Compos. Mater., vol. 47,

no. 20–21, pp. 2443–2474, Dec. 2012.

[103] J. Varna, A. Krasnikovs, R. S. Kumar, and R. Talreja, “A Synergistic Damage Mechanics Approach

to Viscoelastic Response of Cracked Cross-ply Laminates,” Int. J. Damage Mech., vol. 13, no. 4,

pp. 301–334, Oct. 2004.

[104] R. Joffe, A. Krasnikovs, and J. Varna, “COD-based simulation of transverse cracking and stiffness

reduction in [S/90n]s laminates,” Compos. Sci. Technol., vol. 61, no. 5, pp. 637–656, 2001.

[105] C. V. Singh and R. Talreja, “Analysis of multiple off-axis ply cracks in composite laminates,” Int.

J. Solids Struct., vol. 45, no. 16, pp. 4574–4589, 2008.

[106] J. Montesano and C. V. Singh, “Predicting evolution of ply cracks in composite laminates subjected

to biaxial loading,” Compos. Part B Eng., vol. 75, pp. 264–273, 2015.

[107] C. V. Singh, “A higher order synergistic damage model for prediction of stiffness changes due to

ply cracking in composite laminates,” Comput. Mater. Contin., vol. 34, no. 3, pp. 227–249, 2013.

[108] D. G. Callister, W. D., & Rethwisch, Fundamentals of materials science and engineering: An

integrated approach., 6th ed. Hoboken, NJ: John Wiley & Sons, Inc., 2008.

[109] C. V. Singh and R. Talreja, “Evolution of ply cracks in multidirectional composite laminates,” Int.

Page 149: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

129

J. Solids Struct., vol. 47, no. 10, pp. 1338–1349, May 2010.

[110] R. A. Schapery, “On the characterization of nonlinear viscoelastic materials,” Polym. Eng. Sci., vol.

9, no. 4, pp. 295–310, 1969.

[111] A. Pramanick and M. Sain, “Application of the ‘theory of mixtures’ to temperature-stress

equivalency in nonlinear creep of thermoplastic/agro-fiber composites,” Polym. Polym. Compos.,

vol. 14, no. 5, pp. 455–472, 2006.

[112] W. S. Lin, A. Pramanick, and M. Sain, “Development of TTMSP to Predict the Performance of

Discontinuous Natural Fibre-Polymer Composites,” Sci. Eng. Compos. Mater., vol. 11, no. 2–3,

pp. 201–208, 2004.

[113] R. M. Guedes, A. T. Marques, and A. Cardon, “Analytical and experimental evaluation of nonlinear

viscoelastic-viscoplastic composite laminates under creep, creep-recovery, relaxation and ramp

loading,” Mech. Time-Depend. Mater., vol. 2, no. 1986, pp. 113–28, 1998.

[114] S. P. Zaoutsos, G. C. Papanicolaou, and A. H. Cardon, “On the non-linear viscoelastic behaviour

of polymer-matrix composites,” Compos. Sci. Technol., vol. 58, no. 6, pp. 883–889, 1998.

[115] M. S. Al-Haik, M. Y. Hussaini, and H. Garmestani, “Prediction of nonlinear viscoelastic behavior

of polymeric composites using an artificial neural network,” Int. J. Plast., vol. 22, no. 7, pp. 1367–

1392, 2006.

[116] D. T. H. Korontzis, L. Vellios, and V. Kostopoulos, “On the Viscoelastic Response of Composite

Laminates,” Mech. Time-Dependent Mater., vol. 4, pp. 381–405, 2000.

[117] M. K. Darabi, R. K. Abu Al-Rub, E. A. Masad, C.-W. Huang, and D. N. Little, “A thermo-

viscoelastic–viscoplastic–viscodamage constitutive model for asphaltic materials,” Int. J. Solids

Struct., vol. 48, no. 1, pp. 191–207, Jan. 2011.

[118] R. S. Kumar and R. Talreja, “A continuum damage model for linear viscoelastic composite

materials,” Mech. Mater., vol. 35, no. 3–6, pp. 463–480, Mar. 2003.

[119] E. Ahci and R. Talreja, “Characterization of viscoelasticity and damage in high temperature

Page 150: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

130

polymer matrix composites,” Compos. Sci. Technol., vol. 66, no. 14, pp. 2506–2519, Nov. 2006.

[120] K. Giannadakis and J. Varna, “Analysis of nonlinear shear stress–strain response of unidirectional

GF/EP composite,” Compos. Part A Appl. Sci. Manuf., vol. 62, pp. 67–76, 2014.

[121] K. Ogi and Y. Takao, “Evolution of transverse cracking in CF/Epoxy cross-ply laminates under

creep loading,” J. Reinf. Plast. Compos., vol. 18, no. 13, pp. 1220–1230, 1999.

[122] K. Ogi and P. A. Smith, “Modeling creep and recovery behavior of a quasi-isotropic laminate with

transverse cracking,” Adv. Compos. Mater., vol. 11, no. 1, pp. 81–93, 2002.

[123] K. Ogi and T. Shiraishi, “Viscoelastic shear lag analysis of a cross-ply laminate with transverse

cracking,” in The Japan Society of Mechanical Engineers, 2001, pp. 277–280.

[124] A. Asadi and J. Raghavan, “Model for prediction of simultaneous time-dependent damage

evolution in multiple plies of multidirectional polymer composite laminates and its influence on

creep,” Compos. Part B Eng., vol. 79, pp. 359–373, 2015.

[125] I. Taraghi, A. Fereidoon, and F. Taheri-behrooz, “Low-velocity impact response of woven Kevlar

/ epoxy laminated composites reinforced with multi-walled carbon nanotubes at ambient and low

temperatures,” J. Mater., vol. 53, pp. 152–158, 2014.

[126] A. K. Bandaru, V. V. Chavan, S. Ahmad, R. Alagirusamy, and N. Bhatnagar, “Low velocity impact

response of 2D and 3D Kevlar/polypropylene composites,” Int. J. Impact Eng., vol. 93, pp. 136–

143, 2016.

[127] F. Sarasini et al., “Damage tolerance of carbon/flax hybrid composites subjected to low velocity

impact,” Compos. Part B Eng., vol. 91, pp. 144–153, 2016.

[128] M. Ravandi, W. S. Teo, L. Q. N. Tran, M. S. Yong, and T. E. Tay, “Low velocity impact

performance of stitched flax/epoxy composite laminates,” Compos. Part B Eng., vol. 117, pp. 89–

100, 2017.

[129] B. Yang, Z. Wang, L. Zhou, J. Zhang, and W. Liang, “Experimental and numerical investigation

of interply hybrid composites based on woven fabrics and PCBT resin subjected to low-velocity

Page 151: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

131

impact,” Compos. Struct., vol. 132, pp. 464–476, 2015.

[130] G. Simeoli, D. Acierno, C. Meola, L. Sorrentino, S. Iannace, and P. Russo, “The role of interface

strength on the low velocity impact behaviour of PP/glass fibre laminates,” Compos. Part B Eng.,

vol. 62, pp. 88–96, Jun. 2014.

[131] S. Abrate, R. Sierakowski, G. Newaz, S. Reid, and G. Zhou, Impact on Composite Structures.

Cambridge, 1998.

[132] G. Perillo, N. P. Vedivik, and A. T. Echtermeyer, “Damage development in stitch bonded GFRP

composite plates under low velocity impact: Experimental and numerical results,” J. Compos.

Mater., vol. 49, no. 5, pp. 601–615, 2015.

[133] R. Petrucci et al., “Impact and post-impact damage characterisation of hybrid composite laminates

based on basalt fibres in combination with flax, hemp and glass fibres manufactured by vacuum

infusion,” Compos. Part B Eng., vol. 69, pp. 507–515, 2015.

[134] O. T. Topac, B. Gozluklu, E. Gurses, and D. Coker, “Experimental and computational study of the

damage process in CFRP composite beams under low-velocity impact,” Compos. Part A Appl. Sci.

Manuf., vol. 92, pp. 167–182, 2017.

[135] K. Imielińska and L. Guillaumat, “The effect of water immersion ageing on low-velocity impact

behaviour of woven aramid-glass fibre/epoxy composites,” Compos. Sci. Technol., vol. 64, no. 13–

14, pp. 2271–2278, 2004.

[136] T. Gómez-del Río, R. Zaera, E. Barbero, and C. Navarro, “Damage in CFRPs due to low velocity

impact at low temperature,” Compos. Part B Eng., vol. 36, no. 1, pp. 41–50, 2005.

[137] H. Singh, K. K. Namala, and P. Mahajan, “A damage evolution study of E-glass / epoxy composite

under low velocity impact,” Compos. Part B, vol. 76, pp. 235–248, 2015.

[138] S. Long, X. Yao, and X. Zhang, “Delamination prediction in composite laminates under low-

velocity impact,” Compos. Struct., vol. 132, pp. 290–298, 2015.

[139] D. Feng and F. Aymerich, “Finite element modelling of damage induced by low-velocity impact

Page 152: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

132

on composite laminates,” Compos. Struct., vol. 108, no. 1, pp. 161–171, 2014.

[140] W. Tan, B. G. Falzon, L. N. S. Chiu, and M. Price, “Predicting low velocity impact damage and

Compression-After-Impact (CAI) behaviour of composite laminates,” Compos. Part A Appl. Sci.

Manuf., vol. 71, pp. 212–226, 2015.

[141] B. B. Liao and P. F. Liu, “Finite element analysis of dynamic progressive failure of plastic

composite laminates under low velocity impact,” Compos. Struct., vol. 159, pp. 567–578, 2017.

[142] C. S. Lopes, S. Sádaba, C. González, J. Llorca, and P. P. Camanho, “Physically-sound simulation

of low-velocity impact on fiber reinforced laminates,” Int. J. Impact Eng., vol. 92, pp. 3–17, Jun.

2016.

[143] J. Marzbanrad, M. Alijanpour, and M. S. Kiasat, “Design and analysis of an automotive bumper

beam in low-speed frontal crashes,” Thin-Walled Struct., vol. 47, no. 8–9, pp. 902–911, 2009.

[144] R. Hosseinzadeh, M. M. Shokrieh, and L. B. Lessard, “Parametric study of automotive composite

bumper beams subjected to low-velocity impacts,” Compos. Struct., vol. 68, no. 4, pp. 419–427,

2005.

[145] C. V. Singh and R. Talreja, “A synergistic damage mechanics approach to mechanical response of

composite laminates with ply cracks,” J. Compos. Mater., vol. 47, pp. 2475–2501, 2012.

[146] Y. Zuo, J. Montesano, and C. V. Singh, “Assessing progressive failure in long wind turbine blades

under quasi-static and cyclic loads,” Renew. Energy, vol. 119, pp. 754–766, 2018.

[147] J. Montesano, H. Chu, and C. V. Singh, “Development of a physics-based multi-scale progressive

damage model for assessing the durability of wind turbine blades,” Compos. Struct., vol. 141, no.

February, pp. 50–62, 2016.

[148] I. G. García, V. Mantič, and E. Graciani, “Debonding at the fibre-matrix interface under remote

transverse tension. One debond or two symmetric debonds?,” Eur. J. Mech. A/Solids, vol. 53, pp.

75–88, 2015.

[149] B. Budiansky, A. G. Evans, and J. W. Hutchinson, “Fiber-matrix debonding effects on cracking in

Page 153: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

133

aligned fiber ceramic composites,” Int. J. Solids Struct., vol. 32, no. 3–4, pp. 315–328, 1995.

[150] I. J. Beyerlein and S. L. Phoenix, “Stress profiles and energy release rates around fiber breaks in a

lamina with propagating zones of matrix yielding and debonding,” Compos. Sci. Technol., vol. 57,

no. 8, pp. 869–885, 1997.

[151] I. J. Beyerlein and S. L. Phoenix, “Stress concentrations around multiple fiber breaks in an elastic

matrix with local yielding or debonding using quadratic influence superposition,” J. Mech. Phys.

Solids, vol. 44, no. 12, pp. 1997–2036, 1996.

[152] H. C. Tsai, A. M. Arocho, and L. W. Gause, “Prediction of fiber-matrix interphase properties and

their influence on interface stress, displacement and fracture toughness of composite material,”

Mater. Sci. Eng. A, vol. 126, no. 1–2, pp. 295–304, 1990.

[153] B. Budiansky, J. W. Hutchinson, and A. G. Evans, “Matrix fracture in fiber-reinforced ceramics,”

J. Mech. Phys. Solids, vol. 34, no. 2, pp. 167–189, 1986.

[154] H. D. Wagner, J. A. Nairn, and M. Detassis, “Toughness of interfaces from initial fiber-matrix

debonding in a single fiber composite fragmentation test,” Appl. Compos. Mater. An Int. J. Sci.

Appl. Compos. Mater., vol. 2, no. 2, pp. 107–117, 1995.

[155] B. W. Kim and J. A. Nairn, “Experimental Verification of the Effects of Friction and Residual

Stress on the Analysis of Interfacial Debonding and Toughness in Single Fiber Composites,” J.

Mater. Sci., vol. 3972, pp. 3965–3972, 2002.

[156] D. Notta-Cuvier, F. Lauro, and B. Bennani, “Modelling of progressive fibre/matrix debonding in

short-fibre reinforced composites up to failure,” Int. J. Solids Struct., vol. 66, pp. 140–150, 2014.

[157] A. Pupurs and J. Varna, “3-D FEM Modeling of fiber/matrix interface debonding in UD composites

including surface effects,” IOP Conf. Ser. Mater. Sci. Eng., vol. 31, no. 1, p. 12001, 2012.

[158] I. G. García, M. Paggi, and V. Mantic;, “Fiber-size effects on the onset of fiber-matrix debonding

under transverse tension: A comparison between cohesive zone and finite fracture mechanics

models,” Eng. Fract. Mech., vol. 115, pp. 96–110, 2014.

Page 154: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

134

[159] A. Ghavami, A. Abedian, and M. Mondali, “Finite difference solution of steady state creep

deformations in a short fiber composite in presence of fiber/matrix debonding,” Mater. Des., vol.

31, no. 5, pp. 2616–2624, 2010.

[160] J. Segurado and J. Llorca, “A computational micromechanics study of the effect of interface

decohesion on the mechanical behavior of composites,” Acta Mater., vol. 53, no. 18, pp. 4931–

4942, 2005.

[161] V. Tvergaard, “Model studies of fibre breakage and debonding in a metal reinforced by short

fibres,” J. Mech. Phys. Solids, vol. 41, no. 8, pp. 1309–1326, 1993.

[162] V. Tvergaard, “Analysis of tensile properties for a whisker-reinforced metal-matrix composite,”

Acta Metall. Mater., vol. 38, no. 2, pp. 185–194, 1990.

[163] C. F. Niordson and V. Tvergaard, “Nonlocal plasticity effects on the tensile properties of a metal

matrix composite,” Eur. J. Mech. A/Solids, vol. 20, no. 4, pp. 601–613, 2001.

[164] H. Huang and R. Talreja, “Numerical simulation of matrix micro-cracking in short fiber reinforced

polymer composites: Initiation and propagation,” Compos. Sci. Technol., vol. 66, no. 15, pp. 2743–

2757, Dec. 2006.

[165] T. Okabe, T. Sasayama, and J. Koyanagi, “Micromechanical simulation of tensile failure of

discontinuous fiber-reinforced polymer matrix composites using Spring Element Model,” Compos.

Part A Appl. Sci. Manuf., vol. 56, pp. 64–71, 2014.

[166] T. Sasayama, T. Okabe, Y. Aoyagi, and M. Nishikawa, “Prediction of failure properties of

injection-molded short glass fiber-reinforced polyamide 6,6,” Compos. Part A Appl. Sci. Manuf.,

vol. 52, pp. 45–54, 2013.

[167] T. Okabe, M. Nishikawa, and N. Takeda, “Micromechanics on the Rate-dependent Fracture of

Discontinuous Fiber-reinforced Plastics,” Int. J. Damage Mech., vol. 19, no. 3, pp. 339–360, 2010.

[168] S. Kammoun, I. Doghri, L. Brassart, and L. Delannay, “Micromechanical modeling of the

progressive failure in short glass-fiber reinforced thermoplastics - First Pseudo-Grain Damage

Page 155: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

135

model,” Compos. Part A Appl. Sci. Manuf., vol. 73, pp. 166–175, 2015.

[169] H. W. Wang, H. W. Zhou, H. W. Ji, and X. C. Zhang, “Application of extended finite element

method in damage progress simulation of fiber reinforced composites,” Mater. Des., vol. 55, pp.

191–196, 2014.

[170] M. G. Pike and C. Oskay, “Three-Dimensional Modeling of Short Fiber-Reinforced Composites

with Extended Finite-Element Method,” J. Eng. Mech., vol. 142, no. 2000, pp. 1–12, 2015.

[171] J. Koyanagi, Y. Sato, T. Sasayama, T. Okabe, and S. Yoneyama, “Numerical simulation of strain-

rate dependent transition of transverse tensile failure mode in fiber-reinforced composites,”

Compos. Part A Appl. Sci. Manuf., vol. 56, pp. 136–142, 2014.

[172] Z. Li, S. Ghosh, N. Getinet, and D. J. O’Brien, “Micromechanical modeling and characterization

of damage evolution in glass fiber epoxy matrix composites,” Mech. Mater., vol. 99, pp. 37–52,

2016.

[173] Z. Tang, C. Wang, and Y. Yu, “Failure response of fiber-epoxy unidirectional laminate under

transverse tensile/compressive loading using finite-volume micromechanics,” Compos. Part B

Eng., vol. 79, pp. 331–341, 2015.

[174] C. González and J. LLorca, “Mechanical behavior of unidirectional fiber-reinforced polymers

under transverse compression: Microscopic mechanisms and modeling,” Compos. Sci. Technol.,

vol. 67, no. 13, pp. 2795–2806, 2007.

[175] S. A. Elnekhaily and R. Talreja, “Damage initiation in unidirectional fiber composites with different

degrees of nonuniform fiber distribution,” Compos. Sci. Technol., vol. 155, pp. 22–32, 2018.

[176] M. Nishikawa, T. Okabe, and N. Takeda, “Effect of the Microstructure on the Fracture Mode of

Short-Fiber Reinforced Plastic Composites,” J. Solid Mech. Mater. Eng., vol. 3, no. 8, pp. 998–

1009, 2009.

[177] J. Varna, E. Spārniņš, R. Joffe, K. Nättinen, and J. Lampinen, “Time-dependent behavior of

flax/starch composites,” Mech. Time-Dependent Mater., vol. 16, no. 1, pp. 47–70, Oct. 2011.

Page 156: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

136

[178] M. Shirinbayan, J. Fitoussi, M. Bocquet, F. Meraghni, B. Surowiec, and A. Tcharkhtchi, “Multi-

scale experimental investigation of the viscous nature of damage in Advanced Sheet Molding

Compound (A-SMC) submitted to high strain rates,” Compos. Part B Eng., vol. 115, pp. 1–11,

2017.

[179] B. N. Nguyen, B. J. Tucker, and M. A. Khaleel, “A Mechanistic Approach to Matrix Cracking

Coupled With Fiber – Matrix Debonding in Short-Fiber Composites,” J. Eng. Mater. Technol., vol.

127, pp. 337–350, 2005.

[180] J. Montesano and C. V Singh, “Predicting evolution of ply cracks in composite laminates subjected

to biaxial loading,” Compos. Part B Eng., vol. 75, pp. 264–273, 2015.

[181] V. Vinogradov and Z. Hashin, “Probabilistic energy based model for prediction of transverse

cracking in cross-ply laminates,” Int. J. Solids Struct., vol. 42, no. 2, pp. 365–392, 2005.

[182] C. Zhang, J., Fan, J., & Soutis, “Analysis of multiple matrix cracking in [+/-θm/90n]s composite

laminates. part 2: Development of transverse ply cracks,” Composites, vol. 23, no. 5, pp. 299–304,

1992.

[183] L. J. Zapas and J. M. Crissman, “Creep and recovery behaviour of ultra-high molecular weight

polyethylene in the region of small uniaxial deformations,” Polymer (Guildf)., vol. 25, no. 1, pp.

57–62, 1984.

[184] M. E. Tuttle and H. F. Brinson, “Prediction of the long-term creep compliance of general composite

laminates,” Exp. Mech., vol. 26, no. 1, pp. 89–102, 1986.

[185] A. Asadi and J. Raghavan, “Model for prediction of simultaneous time-dependent damage

evolution in multiple plies of multidirectional polymer composite laminates and its influence on

creep,” Compos. Part B Eng., vol. 79, pp. 359–373, 2015.

[186] T. Berton, J. Montesano, and C. V. Singh, “Development of a synergistic damage mechanics model

to predict evolution of ply cracking and stiffness changes in multidirectional composite laminates

under creep,” Int. J. Damage Mech., vol. 25, no. 7, pp. 1060–1078, 2016.

Page 157: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

137

[187] M. Megnis and J. Varna, “Micromechanics based modeling of nonlinear viscoplastic response of

unidirectional composite,” Compos. Sci. Technol., vol. 63, no. 1, pp. 19–31, 2003.

[188] L. F. Sánchez-Heres, J. W. Ringsberg, and E. Johnson, “Effects of Matrix Cracking on the

Estimation of Operational Limits of Frp Laminates,” in 15th European Conference on Composite

Materials, 2012, no. June, pp. 1–8.

[189] M. M. Shokrieh and M. J. Omidi, “Tension behavior of unidirectional glass/epoxy composites

under different strain rates,” Compos. Struct., vol. 88, no. 4, pp. 595–601, 2009.

[190] “ANSYS v15.0 User Manual.” ANSYS, Inc., Cannonsburg, PA, 2011.

[191] P. A. Carraro, E. Novello, M. Quaresimin, and M. Zappalorto, “Delamination onset in symmetric

cross-ply laminates under static loads: Theory, numerics and experiments,” Compos. Struct., vol.

176, pp. 420–432, 2017.

[192] J. Montesano, B. McCleave, and C. V. Singh, “Prediction of ply crack evolution and stiffness

degradation in multidirectional symmetric laminates under multiaxial stress states,” Compos. Part

B Eng., vol. 133, pp. 53–67, 2018.

[193] C. L. Tucker and E. Liang, “Stiffness predictions for unidirectional short- fiber composites : Review

and evaluation,” vol. 59, 1999.

[194] A. Needleman, “Some issues in cohesive surface modeling,” in Procedia IUTAM, 2013, vol. 10,

pp. 221–246.

[195] M. Elices, G. V. Guinea, J. Gómez, and J. Planas, “The cohesive zone model: advantages,

limitations and challenges,” Eng. Fract. Mech., vol. 69, no. 2, pp. 137–163, 2002.

[196] C. Popelar, C. Popelar, and V. Kenner, “Viscoelastic material characterization and modeling for

polyethylene,” Polym. Eng. Sci., vol. 30, no. 10, 1990.

[197] F. Alavi, A. H. Behravesh, and M. Mirzaei, “Mixed-mode cohesive zone modeling and damage

prediction of irregular-shaped interfaces in wood–plastic composites,” Compos. Interfaces, vol. 22,

no. 7, pp. 651–662, 2015.

Page 158: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

138

[198] R. C. Neagu, “Stiffness Contribution of Various Wood Fibers to Composite Materials,” J. Compos.

Mater., vol. 40, no. 8, pp. 663–699, Jul. 2005.

[199] A. Krairi and I. Doghri, “A thermodynamically-based constitutive model for thermoplastic

polymers coupling viscoelasticity, viscoplasticity and ductile damage,” Int. J. Plast., vol. 60, pp.

163–181, 2014.

[200] Y. Pan and A. Pelegri, “Influence of matrix plasticity and residual thermal stress on interfacial

debonding of a single fiber composite,” J. Mech. Mater. Struct., vol. 5, no. 1, pp. 129–142, 2010.

[201] J. Bergström, “Plasticity Models,” in Mechanics of Solid Polymers, 2015, pp. 353–369.

[202] A. Turon, C. G. Dávila, P. P. Camanho, and J. Costa, “An engineering solution for mesh size effects

in the simulation of delamination using cohesive zone models,” Eng. Fract. Mech., vol. 74, no. 10,

pp. 1665–1682, 2007.

[203] M. Smith, “ABAQUS/Standard User’s Manual, Version 6.9.” Simulia, Providence, 2009.

[204] M. Jerabek, Z. Major, K. Renner, J. Móczó, B. Pukánszky, and R. W. Lang, “Filler/matrix-

debonding and micro-mechanisms of deformation in particulate filled polypropylene composites

under tension,” Polymer (Guildf)., vol. 51, no. 9, pp. 2040–2048, 2010.

[205] S. Armioun, S. Panthapulakkal, J. Scheel, J. Tjong, and M. Sain, “Biopolyamide hybrid composites

for high performance applications,” J. Appl. Polym. Sci., vol. 43595, p. n/a-n/a, 2016.

[206] S. Armioun, S. Panthapulakkal, J. Scheel, J. Tjong, and M. Sain, “Sustainable and lightweight

biopolyamide hybrid composites for greener auto parts,” Can. J. Chem. Eng., no. August, 2016.

[207] N. Obaid, M. T. Kortschot, and M. Sain, “Predicting the stress relaxation behavior of glass-fiber

reinforced polypropylene composites,” Compos. Sci. Technol., vol. 161, no. April, pp. 85–91, 2018.

[208] N. Obaid, M. T. Kortschot, and M. Sain, “Modeling and predicting the stress relaxation of

composites with short and randomly oriented fibers,” Materials (Basel)., vol. 10, no. 10, 2017.

[209] V. Placet, O. Cissé, and M. Lamine Boubakar, “Nonlinear tensile behaviour of elementary hemp

fibres. Part I: Investigation of the possible origins using repeated progressive loading with in situ

Page 159: COMPUTATIONAL MODELLING OF PROGRESSIVE DAMAGE IN ... · Finally, I want to thank my friends: Sami, Jean-Nicolas, Mark, Matt, Mireille, Thomas, Mohit, Shwetank, Kulbir, Nav, Sean,

139

microscopic observations,” Compos. Part A Appl. Sci. Manuf., vol. 56, pp. 319–327, 2014.

[210] G. A. Schoeppner and S. Abrate, “Delamination threshold loads for low velocity impact on

composite laminates,” Compos. Part A Appl. Sci. Manuf., vol. 31, no. 9, pp. 903–915, 2000.

[211] F. Zeng, H. Xie, Q. Liu, F. Li, and W. Tan, “Design and optimization of a new composite bumper

beam in high-speed frontal crashes,” Struct. Multidiscip. Optim., vol. 53, no. 1, pp. 115–122, 2016.

[212] Y. Zuo, J. Montesano, and C. V. Singh, “Assessing progressive failure in long wind turbine blades

under quasi-static and cyclic loads,” Renew. Energy, vol. 119, pp. 754–766, 2018.

[213] J. Montesano, H. Chu, and C. V. Singh, “Development of a physics-based multi-scale progressive

damage model for assessing the durability of wind turbine blades,” Compos. Struct., vol. 141, no.

February, pp. 50–62, 2016.


Recommended