EFFICIENT LOAD DISTRIBUTION ANALYSIS AND · A.2.2 Matrix Inversion (Gauss-Jordan Elimination) A-2 A.2.3 Gauss-Elimination with Back Substitution A-4 A.3 Number of Scalar Operations

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

  • EFFICIENT LOAD DISTRIBUTION ANALYSIS AND

    STRENGTH PREDICTION OF BOLTED COMPOSITE

    JOINTS AT VARIOUS LOADING RATES

    Philip Anthony Sharos

    Thesis submitted in fulfilment of the requirements for the Degree of Doctor of Philosophy at the Faculty of Science and Engineering,

    University of Limerick

    Candidate Supervisor: Dr. Conor T. McCarthy

    Submitted to the University of Limerick: May 2016

  • i

    ABSTRACT

    Mechanical fasteners are used extensively in composite aircraft structures, offering a

    cost-effective, reliable and repeatable joining method while also facilitating

    disassembly for maintenance and repair. However, current design practices rely

    extensively on experimentation and detailed numerical analyses which are time

    consuming and expensive and thus significantly increase development and production

    costs. Although three-dimensional finite element analysis can be used in-lieu of more

    expensive experimental tests, current state-of-the-art computing technology cannot

    facilitate analysis of large-scale sub-structures (e.g. panels, wing spars, etc.) with

    potentially thousands of fasteners. The work conducted herein addresses this issue

    through the development of highly efficient analysis methodologies for bolted

    composite joints loaded at various rates.

    Several highly efficient numerical methods are developed that can account for

    laminate-laminate friction, bolt-hole clearance, non-linear material behaviour, bolt

    failure and loading-rate effects. Validated against quasi-static and high-rate

    experimental data, these methods have significantly advanced the state-of-the-art, due

    to their ability to accurately predict the response of multi-row, multi-column joints in

    seconds. Such an increase in efficiency is owed to the implementation of a series of

    parameterised functions which model the local mechanical behaviour of the fastener

    and fastened material. This approach allows the statistical variations arising from

    manufacturing and material processing to be easily accounted for, a feat that is

    impractical using traditional modelling techniques and near-impossible

    experimentally. Furthermore, these methods have been developed into user-friendly,

    stand-alone joint analysis tools and a user-defined finite element which is

    straightforward to implement in commercial finite element software.

    The developed analysis methods were used to investigate the effects of loading rate

    and relative spacing between fasteners in multi-bolt joints. Although loading rate

    effects primarily manifest as changes in joint failure loads and energy absorption, the

    propagation of stress waves also leads to variations in load distribution of up to 1.7%

    in the joints considered. Furthermore, it was found that a strain shielded region exists

    in all multi-fastener arrays and consequently any bolts positioned within this region

    carry a reduced proportion of load. When load is evenly distributed across the joint,

    increased failure initiation loads are to be expected, however, the most effective

    method of augmenting joint strength is through the use of additional fasteners.

    Additionally, the use of a load-path visualisation methodology gave novel insight into

    the role of fibre orientation on load transfer in bolted multi-directional laminates. It

    was found that load is transferred in a trajectory that is strongly dependent on fibre

    orientation, which is important in the distribution of load in multi-fastener joints.

  • iii

    DECLARATION

    The substance of this thesis is the original work of the author and due reference and

    acknowledgement has been made, where necessary, to the work of others. No part of

    this thesis has already been submitted and is being concurrently submitted in

    candidature for any other degree.

    Candidate Supervisor

    Philip A. Sharos Dr. Conor McCarthy

    Signature Signature

    Date Date

    Examination Board

    Chairman: Dr. Jeremy Robinson

    External Examiner: Prof. Carlos Santiuste

    Internal Examiner: Prof. Michael McCarthy

  • v

    ACKNOWLEDGEMENTS

    First and foremost, I would like to express my utmost gratitude to my supervisor,

    Dr. Conor McCarthy, for his guidance, encouragement and patience throughout my

    studies in UL. His knowledge in finite element modelling and in the mechanics of

    composite bolted joints was invaluable. I am grateful to have had the opportunity to

    work under his supervision.

    I would like to gratefully acknowledge the financial support provided to be by the Irish

    Research Council (IRC) and the Donegal County Council.

    I would also like to thank Dr. Brian Egan and Dr. Paddy Gray for making their

    experimental data available to me (work that was frequently referenced throughout

    this thesis). Without access to this data, much of my work would not have been

    possible. Also sincere thank you to Dr. Ronan O’Higgins who kindly answered my

    numerous questions.

    A special thanks to the friends I have made during my time in UL and those who have

    helped me along the way including (in no particular order!) Brian, Donal, Rob, Ned,

    Brian, Rory, Niall, Jon, John, Kev, James and my good friends in the Skydive Club.

    A quick shout out to Mark, I’m sorry Dan and I undoubtedly ruined your beloved

    Counter-Strike for you! And to Dan, with whom I’ve shared too many Chinese

    takeaways with during all those late nights in the office. Also, I would be remiss if I

    did not mention my good friend Chrispy for keeping me relatively sane with the

    occasional Friday night pint(s).

    Finally I would like to thank my family for their unconditional love and support, not

    just during this process but throughout my life.

  • vii

    CONTENTS

    List of Figures xi

    List of Tables xxi

    Nomenclature xxiii

    Chapter 1: Introduction

    1.1 Background and Motivation 2

    1.2 Objectives 6

    1.3 Thesis Outline 7

    Chapter 2: Literature Review

    2.1 Introduction 10

    2.2 Mechanical Behaviour of Bolted Joints 10

    2.2.1 Terminology 10

    2.2.2 Failure Modes of Bolted Joints 12

    2.2.3 Material and Lay-up 15

    2.2.4 Geometry 16

    2.2.5 Fastener Type 17

    2.2.6 Bolt-hole Clearance 18

    2.2.7 Bolt Torque and Lateral Constraint 20

    2.2.8 Loading Rate 24

    2.2.9 Multiple Fasteners 29

    2.3 Detailed Finite Element Modelling of Bolted Joints 33

    2.3.1 Simulation of Elastic Joint Behaviour 33

    2.3.2 Simulation of Joint Failure 35

    2.4 Efficient Analysis Methods 39

    2.4.1 Closed-Form Approaches 39

    2.4.2 Finite Element Analysis 43

    2.5 Discussion of Literature Review 49

  • viii

    2.5.1 Mechanical Response of Bolted Joints 49

    2.5.2 Detailed Finite Element Analysis 51

    2.5.3 Efficient Analysis Techniques 51

    Chapter 3: Analytical Model for Strength Prediction of Joints Under

    Quasi-Static and High-Speed Loading

    3.1 Introduction 54

    3.2 Problem Description 54

    3.3 Model Development 55

    3.3.1 Undamaged Joint Response 58

    3.3.2 Damaged Joint Response and Rate Effects 61

    3.4 Model Validation 68

    3.4.1 Single Fastener Results 70

    3.4.2 Multi-Fastener Results 73

    3.5 Conclusions 76

    Chapter 4: Development and Validation of Finite Element Code for

    Multi-Fastener Joints

    4.1 Introduction 78

    4.2 Computational Efficiency of FE Solution Methods 78

    4.3 Mesh Generation and Element Stiffness Matrices 82

    4.4 Direct Solution Method 84

    4.4.1 Fastener Element Behaviour 84

    4.4.2 Boundary Conditions and Loading 86

    4.5 Virtual Fastener Method 87

    4.5.1 Fastener Reaction Forces 88

    4.5.2 Displacement Field Approximation 89

    4.5.3 Summary of Virtual Fastener Method 93

    4.6 Model Validation 94

    4.6.1 Statistical Variation of Fastener Properties 94

  • ix

    4.6.2 Effect of Missing Fasteners 101

    4.6.3 Computational Efficiency 103

    4.7 Graphical User Interface 106

    4.8 Conclusions 109

    Chapter 5: User-Defined Finite Element for Bolted Joints in Explicit

    FE Solvers

    5.1 Introduction 112

    5.2 Model Development 114

    5.2.1 Overview 114

    5.2.2 User Element Behaviour 116

    5.2.3 Composite Material Modelling 124

    5.3 Model Validation 126

    5.3.1 Single Bolt Results 126

    5.3.2 Multi-fastener Quasi-static Results 129

    5.3.3 Dynamic Results 135

    5.4 Effect of velocity on load-distribution 137

    5.5 Conclusions 143

    Chapter 6: Analysis of Load Distribution in Multi-Fastener Joints

    6.1 Introduction 146

    6.2 Load Path Study 146

    6.2.1 Theoretical Overview 146

    6.2.2 Effect of Fibre Orientation 149

    6.2.3 Missing Fasteners 156

    6.3 Fastener Positioning 161

    6.4 Conclusions 167

    Chapter 7: Conclusions

    7.1 Introduction 170

  • x

    7.2 Notes on the Mechanical Response of Bolted Joints 170

    7.2.1 Variations in Individual Bolt Response in Multi-Fastener Joints 170

    7.2.2 Effects of High-Rate Loading 171

    7.2.3 Effect of Fastener Position 172

    7.2.4 Effects of Material and Layup 172

    7.2.5 Guidelines for Design 173

    7.3 Highly Efficient Numerical Methods 174

    7.4 Recommendations for Future Work 177

    Bibliography 179

    Appendix A: Approximation of Computational Time Required for

    Finite Element Solver Methodologies

    A.1 Overview A-1

    A.2 Approximation for Common Matrix Operations A-1

    A.2.1 Matrix Multiplication A-2

    A.2.2 Matrix Inversion (Gauss-Jordan Elimination) A-2

    A.2.3 Gauss-Elimination with Back Substitution A-4

    A.3 Number of Scalar Operations for FE Solution Methods A-5

    A.3.1 Direct Solution Method A-5

    A.3.2 Virtual Fastener Method (VFM) A-5

    A.3.3 Explicit Method A-6

    A.4 Notes on the Efficiency of the VUEL Model A-7

    Appendix B: Input File Configuration and VUEL Subroutine for User-

    Defined Finite Element

    B.1 Introduction A-9

    B.2 ABAQUS Input File Configuration A-9

    B.3 MATLAB Script to Generate UEL Properties A-11

    B.4 VUEL FORTRAN Subroutine A-25

  • xi

    LIST OF FIGURES

    Chapter 1

    Figure 1.1 Boeing 787 showing breakdown of materials used (redrawn from

    Georgiadis et al. 2008)).

    2

    Figure 1.2 Examples of materials and process development acceleration

    using computational tools under the AIM-C project (a)

    traditional testing supported analysis approach (b) analysis

    approach supported by experience, testing and demonstration

    (redrawn from National Research Council (2004)).

    4

    Figure 1.3 Certification procedure for composite materials (redrawn from

    MAAXIMUS (2009)).

    5

    Chapter 2

    Figure 2.1 Definition of geometric parameters of bolted joints (redrawn

    from McCarthy (2003)).

    11

    Figure 2.2 Failure modes of mechanically fastened joints (redrawn from

    Niu (1992)).

    13

    Figure 2.3 Definition of (a) clearance and (b) contact angle. 19

    Figure 2.4 Effect of bolt torque on initial load-displacement curve of

    single-bolt, single-lap joint with clearance (redrawn from

    McCarthy et al. (2002)).

    21

    Figure 2.5 (a) Loads carried by various interactions in single-lap joint (b)

    Bolt and total friction load distribution in multi-bolt joint with

    variable hole clearance, C1=neat-fit, C4=240 µm clearance

    (redrawn from McCarthy et al. (2005)).

    23

    Figure 2.6 (a) Load-displacement behaviour of pinned and clamped

    laminated (redrawn from Kelly & Hallström (2004)); (b) Effect

    of clamping pressure on ultimate bearing strength of bolted

    joints (redrawn from Park (2001)).

    24

  • xii

    Figure 2.7 Quasi-static and high rate stress strain curves for: (a) Transverse

    compression and; (b) shear in unidirectional carbon/epoxy

    composites (redrawn from Hsiao & Daniel (1998)).

    26

    Figure 2.8 Comparison of quasi-static and dynamic bearing response of

    composite joints redrawn from (a) Ger at al. (1996); (b) Pearce et

    al. (2009).

    28

    Figure 2.9 Bearing-bypass loading within a multi-fastener joint (redrawn

    from Crews & Naik (1987)).

    30

    Figure 2.10 Determining bypass loads in a multi-bolt joint (redrawn from

    McCarthy et al. (2005)).

    31

    Figure 2.11 Loading of single lap joint with (a) no deformation; (b)

    deformation with rigid members; (c) deformation with elastic

    members (redrawn from Eriksson et al. (1995)).

    32

    Figure 2.12 Bolt load distribution in 3-bolt double-lap composite joint with

    (a) equal clearance at each hole (neat-fit); (b) 240µm clearance at

    hole 1, neat-fit at holes 2 & 3 (redrawn from McCarthy et al.

    (2005)).

    33

    Figure 2.13 Flowchart of the progressive damage model (redrawn from

    Tserpes et al. (2001)).

    36

    Figure 2.14 (a) Meshing strategy and material property assignments in the

    bolt-hole region and (b) cohesive element locations (redrawn

    from Egan et al. (2015)).

    39

    Figure 2.15 Figure 2.15 Spring element model for a three-bolt joint (redrawn

    from McCarthy et al. (2006)).

    41

    Figure 2.16 Geometry of bolted hybrid joint (redrawn from Caccese et al.

    (2007)).

    45

    Figure 2.17 Video sequences from bird strike on specimen versus simulation

    results with rivet failure law: (a) t = 4.5ms; (b) t = 10ms

    (redrawn from McCarthy et al. (2004b)).

    46

  • xiii

    Figure 2.18 2.18 PLINK model load-displacement behaviour (redrawn from

    Gunnion et al. (2006)).

    47

    Chapter 3

    Figure 3.1 Typical load-displacement curve for single-bolt, single-lap joint:

    (a) complete load displacement response; (b) enlarged view of

    undamaged loading behaviour.

    55

    Figure 3.2 Three-bolt, single-lap joint: (a) illustration; (b) corresponding

    spring/mass system (redrawn from McCarthy et al. (2006).

    56

    Figure 3.3 Contact area for calculating Ks in highly torqued joint. 58

    Figure 3.4 Derivation of spring stiffness term for tapered laminate section:

    (a) Equivalent spring model for single bolt joint; (b) Enlarged

    view of taper section.

    60

    Figure 3.5 Possible methods of inserting plies to taper up laminate and

    definition of ply drop off ratio in relation to laminate geometry.

    61

    Figure 3.6 Experimental results of a single-bolt joint tested quasi-statically

    and at loading rates of 5 m/s and 10 m/s, (Redrawn from Egan

    et al. (2013))

    62

    Figure 3.7 Illustration of critical damage variables used in analytical model. 63

    Figure 3.8 Comparison between conic and cubic damage approximation

    functions for a single bolt joint: (a) Low energy case; (b) high-

    energy case.

    64

    Figure 3.9 Use of conic curves in the prediction of the joint non-linear

    stress strain behaviour: (a) elliptical curve and (b) enlarged view

    of hyperbolic curve.

    67

    Figure 3.10 (a) Single-bolt geometry, and (b) multi-bolt geometry. See Table

    3.1 for values of e, w, p, L and g.

    68

    Figure 3.11 Comparison of experimental and model load-displacement

    curves for single-bolt (SB) joints (experiments redrawn from

    Egan et al. (2013) and Gray et al. (2014a), QS = quasi-static, A, C

    71

  • xiv

    and E correspond to the layups used in joint, FF=fastener

    failure, FP=fastener pull-through).

    Figure 3.12 Comparison of experimental and model load-displacement

    curves for single-both E-laminate joints accounting for residual

    fastener strength: (a) 5m/s, and (b) 10m/s.

    72

    Figure 3.13 (a) Fastener numbering in single-column joint. (b) & (c) Quasi-

    static load displacement curve of C-layup and E-layup joints

    with equal fastener properties, respectively. (d) & (e) Individual

    fastener load displacement curves in C-layup and E-layup joints

    with equal fastener properties, respectively. (f) & (g) Individual

    fastener load displacement curves in C-layup and E-layup joints,

    respectively, with different values of energy absorption assigned

    to each fastener (values given in Table 3.5).

    74

    Figure 3.14 Comparison of experimental and model load-displacement

    curves for multi-bolt joints (experiments redrawn from Egan et

    al. (2013) and Gray et al. (2014b), QS = quasi-static, A, C and E

    correspond to the layups used in joint).

    75

    Chapter 4

    Figure 4.1 Comparison of normalised CPU times and number of

    mathematical operations required for basic matrix operations

    (GEBS – gauss elimination with back substitution).

    80

    Figure 4.2 Comparison of theoretical computational times for VFM and

    direct solution method (a) times normalized w.r.t the direct

    method; (b) Ratio between direct and VFM CPU times.

    81

    Figure 4.3 Disadvantages of paving method (a) interference of opposing

    elements (b) element size difference between opposing fronts:

    (redrawn from Lee et al. (2003)).

    83

    Figure 4.4 Mesh post-processing operation to account for specific fastener

    locations.

    84

  • xv

    Figure 4.5 (a) Integration of fastener element with composite laminate

    mesh; (b) Deflection of fastener element due to bearing loads.

    85

    Figure 4.6 Boundary conditions used and example finite element mesh for

    the direct solution method.

    86

    Figure 4.7 Loads and boundary conditions used for Virtual Fastener

    Method (VFM). (Note that XJ symbolises the rigid body motion

    applied to Laminate 2 and is not used as a boundary condition

    in the solution of Eq. (4.6)).

    87

    Figure 4.8 Bearing load applied to composite for VFM: (a) Superimposed

    view of virtual fastener element and bolt-hole interaction; (b)

    Components of bearing load applied to model.

    88

    Figure 4.9 Direction of force exerted on composite material by fastener

    and relative position of virtual fastener nodes: (a) no

    displacement applied (virtual fastener represented in black), (b)

    displacement at Node 1 greater than Node 2 (virtual fastener

    represented in green) and (c) displacement at Node 2 greater

    than Node 1 (virtual fastener represented in red).

    89

    Figure 4.10 (a) Displacement approximation technique for VFM applied to a

    6 bolt joint; (b) enlarged view illustrating inter-fastener effects

    on convergence.

    92

    Figure 4.11 Graphical Representation of Forced Convergence Procedure. 92

    Figure 4.12 Flow chart of Virtual Fastener Method. 93

    Figure 4.13 (a) Single column geometry (SC); (b) Double column geometry

    (DC) and (c) Triple column geometry (TC). All dimensions are

    in mm.

    94

    Figure 4.14 Student’s t-distribution for sample size of 3 (a) Probability

    density function (PDF) and (b) Cumulative distribution function

    (CDF).

    95

    Figure 4.15 Comparison of experimental, direct method and VFM load-

    displacement curves for C-layup, single column (SC) joint

    configuration (QS=quasi-static, OC=offset correction).

    96

  • xvi

    Figure 4.16 Comparison of damage response resulting from randomly

    generated fastener properties using the statistical approach.

    97

    Figure 4.17 Load-displacement curves for multi-fastener joints: (a) C-layup

    SC; (a) E-layup SC; (c) C-layup DC; (d) E-layup DC; (e) C-layup

    TC. (R) indicates that random fastener properties and (E)

    indicates equal properties were used.

    99

    Figure 4.18 (a) Load-displacement curves for single-bolt C-layup joint

    illustrating effect of failure mode on joint strength; (b) Failed SC

    C-Layup specimen.

    100

    Figure 4.19 Load-displacement curves for C-Layup joints with corrected

    failure loads; (a) C-Layup SC, (b) C-Layup DC, (c) C-Layup TC.

    Numerical curves appended (R) and (E) denote that randomly

    generated and equal fastener damage properties were used,

    respectively.

    101

    Figure 4.20 Load-displacement curves of C-layup TC joint with missing

    fasteners: (a) MF-9 configuration; (b) MF-8 configuration; (c)

    MF-5 configuration; (d) MF-6 configuration. Numerical curves

    appended (R) and (E) denote that randomly generated and equal

    fastener damage properties were used, respectively.

    102

    Figure 4.21 Loads carried by individual fasteners in MF-9 Joint: (a) Bolt

    numbers and positions; (b) Experimental data (redrawn from

    Gray et al (2011)); (c) Direct Method with equal fastener

    properties; (d) Direct Method with random fastener properties;

    (e) VFM with equal fastener properties.

    103

    Figure 4.22 CPU Times required for VFM and direct method using constant

    displacement increment (2.5µm).

    105

    Figure 4.23 Effect of displacement step size on the accuracy of the VFM

    and direct methods.

    105

    Figure 4.24 Dialogue boxes to define composite and bolt material

    properties.

    107

  • xvii

    Figure 4.25 Sections of GUI to define geometry and mesh; (a) laminate

    dimensions, seeding and bolt positions, (b) graphic to assist in

    definition of dimensions, (c) bolt geometry.

    108

    Figure 4.26 Finite element analysis solver configuration panel. 108

    Chapter 5

    Figure 5.1 Comparison of operational times for Direct, Explicit and Virtual

    Fastener (VFM) methods: (a) normalised with respect to Direct

    method and (b) normalised with respect to Virtual Fastener

    method.

    113

    Figure 5.2 Joint finite element model (a) boundary conditions and

    constraints and (b) contact surfaces and user-element locations.

    114

    Figure 5.3 Acceleration and velocity time histories (a) smooth step; (b)

    constant velocity.

    115

    Figure 5.4 Method for coupling user-element with laminate mesh. 116

    Figure 5.5 Simplification of joint region and user-element DOFs. 117

    Figure 5.6 a) Bearing load acting on composite material (b) bending

    moment in single-shear joint.

    118

    Figure 5.7 (a) Variation of centre of contact pressure with joint load (b)

    countersunk (CSK) and non-countersunk (NCSK) contact

    surfaces in a single-bolt, single-lap joint.

    119

    Figure 5.8 Equivalent spring stiffness model for out-of-plane loading of

    joint region.

    120

    Figure 5.9 Coupling constraint stiffness correction. 121

    Figure 5.10 Assumptions regarding recoverable elastic energy and unloading

    of joint element.

    123

    Figure 5.11 Effect of stable time increment on (a) CPU time and (b) Load-

    displacement response.

    124

  • xviii

    Figure 5.12 Figure 5.12 Modelling of composite laminate using layered,

    conventional shell elements.

    124

    Figure 5.13 Effect of mass-proportional damping on: (a) composite laminate

    loaded with a ramp force time-history and (b) load-displacement

    response of a single-bolt joint user-element model.

    126

    Figure 5.14 User-element, single-bolt (SB) joint results: (a) deformed finite

    element mesh; (b) load-displacement curve for and (c) out-of-

    plane displacement for C-Layup; (d) load-displacement curve for

    and (e) out-of-plane displacement for E-Layup. Data series

    appended 35% and 50% refer to eccentricity of loading assumed

    as a percentage of laminate thickness.

    128

    Figure 5.15 Internal and kinetic energy in single-bolt E-layup joint. 129

    Figure 5.16 User-element results for a single-column (SC) joint: (a) Bolt

    numbering convention and virtual strain gauge positions; (b)

    Joint load-displacement curve; (c) Comparison of strains output

    from virtual strain gauges; (d) Joint load-distribution calculate

    from surface strains; (e) Comparison of bolt load distribution

    calculated from surface strains on top (TOP) and bottom (BTM)

    laminates with force applied through VUEL element

    (APPLIED).

    131

    Figure 5.17 Fastener loads in quasi-static SC C-Lam joint: (a) Load

    distribution for joint loaded to ultimate failure load, (b) Detailed

    view of load distribution for elastic loading of the joint.

    132

    Figure 5.18 User-element results for a triple-column (TC) joint: (a) Bolt

    numbering convention and virtual strain gauge positions; (b)

    Joint load-displacement curve; (c) Load distribution from load

    applied through the user-element ; (d) Joint load-distribution

    calculated from surface strains (experimental); (e) Joint load-

    distribution calculate from surface strains (FE-Model).

    133

    Figure 5.19 Comparison of surface strains resolved at virtual strain gauge

    positions in QS-TC C-Lam joint; (a) between Bolt 1 and 4; (b)

    134

  • xix

    between Bolt 2 and 5; (c) between Bolt 3 and 6; (d) between

    Bolt 4 and 7; (e) between Bolt 5 and 8; (f) between Bolt 6 and 9.

    Figure 5.20 Comparison of raw and filtered SB C-Lam 10m/s load-

    displacement curves (a) from VUEL model; (b) redrawn from

    Egan at al. (2015).

    136

    Figure 5.21 Comparison of raw and filtered load-displacement curves

    obtained from VUEL model and experimentation by Egan et al.

    (2013).

    136

    Figure 5.22 Load distribution in 9-bolt joint: (a) bolt, row and column

    numbering; (b) QS load distribution history; (c) QS load

    distribution at 25kN joint load; (d) 5m/s load distribution

    history; (e) 5m/s load distribution at 25kN joint load; (f) 10m/s

    load distribution history; (g) 10m/s load distribution at 25kN

    joint load.

    139

    Figure 5.23 Propagation of elastic stress wave in C-laminate material arising

    from end-displacement loading of 10m/s.

    140

    Figure 5.24 Time dependent load distribution due to dynamic effects in

    triple-column, C-Layup joint loaded at 10m/s: (a) Greatest load

    carried by row 1, (b) Load distribution similar to baseline case,

    (c) Greatest load carried by row 3.

    141

    Chapter 6

    Figure 6.1 (a) Force “stream tube” and (b) Construction of force

    components.

    148

    Figure 6.2 Effect of fibre orientation on the trajectory of bearing load

    paths (a) 0°; (b) 30°; (c) 45°; (d) 60°; (e) 90°.

    150

    Figure 6.3 Effect of material orientation on load trajectories in multi-

    fastener joints: (a) 0°; (b) 90°; (c) 45°; (d) -45°.

    153

    Figure 6.4 Effect of fibre-orientation on load distribution in a multi-pin

    joint.

    153

  • xx

    Figure 6.5 Effect of stiffness and anisotropy on load distribution (o) –

    orthotropic, (i) – isotropic.

    155

    Figure 6.6 Comparison of load paths from user-element model and 2D

    FEA.

    156

    Figure 6.7 Load-displacement curves and fastener positions in missing-

    fastener cases: (a) MF5, (b) MF6, (c) MF8, (d) MF9.

    158

    Figure 6.8 Load distribution from VUEL model: (a) MF5; (b) MF6; (c)

    MF8 and (d) MF8, Difference in pin load from 2D FEA.

    160

    Figure 6.9 Patterns used in fastener position study. 162

    Figure 6.10 Load distribution in 9-bolt joint configurations. 163

    Figure 6.11 Load distribution in 8-bolt joint configurations. 164

    Figure 6.12 Distribution of load in 9-bolt joint with fasteners in (a) grid

    arrangement and (b) circular arrangement. Relative bolt

    positions for (c) grid arrangement and (d) circular arrangement.

    166

    Chapter 7

    Figure 7.1 Application of the developed numerical methods to the

    certification procedure for composite materials.

    176

    Appendices

    Figure A.1 Parallelisation benchmark tests for VUEL model. A-8

    Figure B.1 Parameters defining load displacement behaviour of a single-

    bolt, single-lap joint.

    A-10

  • xxi

    LIST OF TABLES

    Chapter 3

    Table 3.1 Joint Geometries (QS=quasi-static, HS=high speed), All

    dimensions are in mm.

    69

    Table 3.2 Orientations of layups tested. 69

    Table 3.3 Averaged key damage variables determined from experimental

    data for C and E laminates loaded quasi-statically and at

    velocities of 5 and 10 m/s.

    69

    Table 3.4 Final (catastrophic) failure mode for single bolt cases (FF =

    fastener failure, FP = fastener pull through).

    70

    Table 3.5 Energies associated with non-linear damage (E1) assigned to

    each bolt when varying fastener properties were applied to the

    quasi-static model

    76

    Chapter 4

    Table 4.1 Number of operations required and complexity of Direct and

    Virtual Fastener solution methods.

    82

    Table 4.2 Mean values (µ) and standard deviations (σ) of damage variables

    from single bolt tests (FF=fastener failure, FP=Fastener Pull

    Through).

    96

    Table 4.3 CPU run-times for Direct Method (DM) and Virtual Fastener

    Method (VFM) using a single core on a quad-core, 32GB RAM

    computer.

    106

    Chapter 5

    Table 5.1 Operations required and complexity of Direct, Virtual Fastener

    and Explicit solution methods.

    113

    Table 5.2 Inertia values for nodes in user-element for composite layups

    considered.

    118

  • xxii

    Table 5.3 Effect of loading velocity on load distribution in TC C-laminate

    joint.

    142

    Chapter 6

    Table 6.1 Initial and ultimate failure loads of multi-fastener configurations. 165

  • xxiii

    NOMENCLATURE

    Upper Case

    A Constants for cubic curve fit

    Ac Projected area of countersunk fastener head onto the shear plane

    B (Chapter 3) Constants for conic curve fit

    B (Chapter 4) Element strain matrix

    C Clearance

    D Vector of known control variables

    Dde Energy dissipated through damage

    DOF Degree of Freedom

    E Young’s Modulus

    Ec Energy dissipated through viscoelasticity and creep

    F Force

    FBi Load carried by bolt i

    G Shear modulus

    I Inertia matrix

    Id Number of increments required for direct and VFM analyses

    IE Internal energy

    Ie Number of increments required for explicit analysis

    Ixx, Iyy, Ixy Moments of Inertia

    K Stiffness matrix

    KB, KBi, Kij Spring stiffness of joint member (i.e. bolt and laminate regions)

    Kc Spring stiffness of clamped region in pull-through

    Kcpl Equivalent spring stiffness of coupling region

    KE Spring stiffness of fastener when loaded via elastic bolt-hole contact

    Ke (Chapter 4) Element stiffness matrix

    Ke (Chapter 5) Kinetic Energy

    KS Spring stiffness of fastener when loaded is reacted through friction

    KQLS Spring stiffness value during quasi-linear unloading of joint

    L Length of laminate

    Mx, My Moment

    P Joint Load

    PF Ultimate joint load

    PFRIC Maximum Joint load reacted through friction

    Pi Joint initial failure load

    RB Residual force arising from the approximated displacement field

    RC Coupling radius

    T Transformation matrix

    TDM, TVFM Computational times for direct method and VFM

  • xxiv

    TF Force threshold for convergence in VFM

    U (Chapter 3) Vector of unknown variables

    U (Chapter 4) Displacement field

    V, Vx, Vy, Vz Stress pointing vector

    Ve Element volume

    VFM Virtual Fastener Method

    W, Wx, Wy Weight factor for first VFM convergence iteration

    XBi Relative displacement of nodes in i-th fastener spring element

    XJ End-displacement of joint

    XPF Joint displacement at ultimate failure load

    XPi Joint displacement at initial failure load

    XP0 Joint displacement at zero load

    Zp Z-coordinate of centre of contact pressure

    Lower Case

    a Constant for conic curve fit

    c (Chapter 3) Constant for conic curve fit

    c (Chapter 5) Damping matrix

    cd Dilatational wave speed

    d, dh Hole diameter

    dH Fastener head diameter

    e Edge distance

    e (Chapter 5) Distance from shear plane

    fNL Non-linear damage function

    k Force-per-hole factor

    m Mass

    n Number of DOF’s in analysis

    o Unit vector defining fibre orientation

    p, pr, pc Bolt pitch

    t Laminate thickness

    u Vector of nodal deflection’s

    v Vector of nodal velocities

    w Laminate width

    x Nodal displacement

    xB Displacement of fastener node in X-direction

    yB Displacement of fastener node in Y-direction

    Greek

    α Mass proportional damping

    β (Chapter 3) Fraction of bending moment reacted by fastener head

    β (Chapter 5) Stiffness proportional damping

  • xxv

    Δt Stable time increment

    εb Bearing strain

    η Convergence increment scaling factor

    θ Angle of tangent to conic curve

    λ Lamé parameter

    μ Shear modulus

    ν Poisson’s ratio

    ρ Density

    σ Stress tensor

    σb Bearing stress

    τ Shear stress

    φ Deviation between local stress trajectory and fibre orientation

  • Chapter 1

    INTRODUCTION

  • Chapter 1

    2

    1.1 Background and Motivation

    Composites, specifically fibre-reinforced plastics (FRPs), offer a number advantages over

    conventional materials such as steel and aluminium alloys. Such advantages include

    increased strength-to-weight ratio, corrosion resistance and fatigue life (Niu 1992),

    making FRP’s an ideal choice in the manufacture of aerospace and automotive structures.

    Consequently composites have seen increased use in light aircraft, military fighters and

    helicopters. However, in commercial aviation the introduction of these materials had been

    gradual, owing to the high safety standards and the conservative nature of the industry.

    One of the first applications of composites in civil aircraft primary structures was in the

    1980’s on the 737 horizontal stabiliser (Roeseler et al. 2007). Since then, the use of

    composites in aircraft structures have continued to increase with the latest generation of

    commercial aircraft, the Boeing 787 and Airbus A350, consisting of over 50% composite

    materials by weight (Boeing 2006; EADS 2009). The extensive use of composites in the

    Boeing 787 is illustrated in Figure 1.1. Additionally, composite materials have seen

    increased use in the automotive industry. For example, carbon fibre reinforced plastic

    (CFRP) was used extensively in the manufacture of the Lexus LFA (Toho Tenax 2010),

    and in the electric BMW i3 (BMW 2013) and Tesla Roadster (JEC Composites 2011),

    both of which feature a 100% CFRP passenger cell. Furthermore, the BMW i3 (formerly

    known as the MCV) was the first example of a mass produced carbon fibre based car

    (JEC Composites 2011; BMW 2013).

    Figure 1.1 Boeing 787 showing breakdown of materials used (redrawn from Georgiadis et al. (2008)).

    In contrast to composite materials, the use of mechanical fasteners is a mature technology

    and is used to great extent in traditional aircraft structures. This can be seen in the Boeing

    747, 767 and 777 which have an estimated 3 million, 1.8 million and 1 million mechanical

  • Chapter 1

    3

    fasteners, respectively (Wallace 2008). Although the manufacturing processes associated

    with FRPs facilitate an overall reduction in the number of fasteners required,

    mechanically fastened joints are still prominent in composite structures. The

    predominantly CFRP fuselage of the 787 uses only an estimated 40k-50k fewer fasteners

    (Walz 2006) when compared to an aluminium structure of similar size (e.g. 767). This

    continued use of bolted joints arises from a number of advantages being associated with

    the method, namely less sensitivity to environmental conditions and surface preparation

    and the ability to disassemble the joint or detach components for maintenance, repair or

    material recycling (Niu 1992).

    Bolted joints represent potential weak points in a structure and can limit its overall load

    carrying capacity. Therefore it is crucial for design engineers to understand and predict

    the behaviour the joint. Joint efficiency is a measure of the strength of the joint compared

    to that of the parent material, and in metals values of 70-80% are typical. However, the

    efficiency of composite joints is much lower (40-50%), and so optimising the joint design

    is essential in order to realise the maximum potential benefits of composites. Some factors

    contributing to the lower efficiency in composite joints are the brittle nature of the

    material leading to less stress relief around load holes and anisotropy leading to higher

    stress concentration factors. In recent years much focus has been placed on the

    development of numerical methods to replace time consuming and expensive

    experimental tests. The EU FP5 project, BOJCAS – Bolted Joints in Composite Aircraft

    Structures was focused on developing advanced numerical methods for bolted joints

    (McCarthy 2001). A key technology which resulted from this project was the three-

    dimensional finite element (3D-FE) modelling of composite bolted joints. These models

    were able to accurately capture the full mechanical behaviour of the joint including effects

    of variables such as bolt torque and clearance. Furthermore, using the developed methods,

    it was possible to predict damage in the joint on the mesoscale (ply-level). However, a

    limitation of this approach was the time required to complete the analyses making 3D-FE

    infeasible for large scale studies.

    In a report by the National Research Council (2004) to the US department of defence, a

    number of recommendations were made regarding the cost-effective, efficient

    implementation of new technologies and materials. One such recommendation was a shift

    in design philosophy from an analysis-supported testing-based approach to a testing-

    supported analysis-based approach with emphasis on efficient modelling and focused

  • Chapter 1

    4

    testing. The Defence Advanced Research Projects Agency (DARPA) program on

    Accelerated Insertion of Materials - Composites (AIM-C) was used as an example of the

    improvements in efficiency that could be made when this philosophy was adopted. Figure

    1.2 summarises the achievements of the AIM-C project in terms of successful model

    integration as part of the design process. A significant reduction in the time required to

    introduce new technology was noted, owed in part to the replacement of a 6-month

    experimental series with 2- to 3-day modelling-based activities. Thus, through the use of

    robust and high-fidelity numerical models, the time required to implement a new

    technology or design can be significantly reduced. Although the AIM-C project was

    focused on the introduction of new materials in the design process, the philosophy can be

    applied to any new technology including the design of bolted composite structures.

    Figure 1.2 Examples of materials and process development acceleration using computational tools under

    the AIM-C project (a) traditional testing supported analysis approach (b) analysis approach supported by

    experience, testing and demonstration (redrawn from National Research Council (2004)).

    The EU FP7 project MAAXIMUS – More Affordable Aircraft through Extended,

    Integrated and Mature Numerical sizing was aimed at the efficient development and right-

    first-time validation of highly optimised composite fuselage structures (MAAXIMUS

    2009). Although one of the aims of the project was the reduction of the amount of

    mechanical fastening required, a significant portion of the project was focused on bolted

    joint design. This is due to the necessity of bolted joints for inspection and repair purposes

    throughout the life-cycle of an aircraft.

  • Chapter 1

    5

    Figure 1.3 Certification procedure for composite materials (redrawn from MAAXIMUS (2009)).

    The typical experiments required in the certification procedure of composite aircraft

    structures are illustrated in Figure 1.3. One of the aims of the BOJCAS project was to

    reduce the quantity of experimental tests required during the certification phase through

    the use of numerical models (McCarthy 2001). However, given the complexity and

    resources required for 3D-FE analyses, this approach was only feasible for coupons and

    small sub-assemblies (Levels 1-2 in Figure 1.3). The numerical methods developed

    through the MAAXIMUS project allow for simulations of larger structural components

    (Levels 3-5) under certain loading conditions. A highly efficient joint model had been

    developed for analysis of large composite structures (Gray and McCarthy 2011), however

    it was subject to a number of limitations, specifically a high dependence on experimental

    data for calibration and an inability to model dynamic problems. The latter is an important

    consideration as some critical load cases in aircraft design are dynamic, i.e. bird-strike

    and tyre-debris impact.

  • Chapter 1

    6

    1.2 Objectives

    The primary objective of this thesis is the development of a highly efficient numerical

    method for modelling composite bolted joints at static and dynamic loading rates. For

    impact and crash analysis the behaviour of the joint beyond initial failure and up to

    complete separation is required (Gunnion et al. 2006), and thus any method developed

    must be able to account for this. Furthermore, to achieve maximum effectiveness for use

    in industry the method developed should also be applicable to the testing-supported

    analysis-based design philosophy. This requires highly accurate numerical models which

    allow the joint design variables (such as clearance, bolt torque, strength, etc.) to be easily

    varied. A certain degree of experimentation will always be necessary as part of the design

    and certification process of aircraft. However, it is envisioned that results from coupon

    testing (Level 1 in Figure 1.3) be used only to calibrate highly efficient numerical models

    which are then used in lieu of experiments in Levels 2-5. Therefore, regarding the

    development of an efficient analysis tool, the primary objectives are as follows:

    1) Investigate and compare numerical solution procedures in terms of computational

    efficiency. The developed model should use the most efficient solution

    methodology available.

    2) Develop a representative model for bolted joints capable of capturing any

    dynamic loading effects and modelling the joint to complete separation.

    3) Integrate a representative model for bolted joint in commercial FE software to

    provide a robust, user-friendly tool for industry.

    Additionally, there are a number of factors regarding the behaviour of bolted joints that,

    due to limitations of experimental methods and traditional modelling techniques, have

    not been fully investigated. Once a highly efficient and accurate joint model has been

    developed, the following secondary objectives are set:

    1) Investigate the effect of statistical variations of fastener properties in multi-bolt

    joints.

    2) Investigate factors affecting load distribution in multi-fastener joints. Parameters

    to consider are material properties and layup, loading velocity and the relative

    spacing between fasteners.

    3) Based on the findings of numerical studies, propose a series of guidelines for the

    design of multi-fastener composite joints.

  • Chapter 1

    7

    1.3 Thesis Outline

    In Chapter 2 a review of relevant publications was conducted. Firstly, the literature

    pertaining to joint design variables and how these effect the mechanical response of the

    joint was discussed. Such variables include bolt torque, clearance, composite layup and

    loading rate. Secondly literature relating to detailed finite element modelling of bolted

    joints was reviewed. The purpose of this was to determine the capabilities and limitations

    of using traditional modelling techniques. Finally, a review was conducted on efficient

    numerical modelling strategies to determine the current state-of-the-art. Areas of focus in

    this section include purpose finite elements and analytical models.

    Chapter 3 presents a semi-empirical method for approximating the damaged response of

    shear-loaded composite bolted joints. Using a novel application of a conic function, the

    damaged load-displacement behaviour of the joint was approximated knowing only the

    initial and ultimate failure loads and energy absorbed. When applied in a simple

    equivalent spring model accurate load-displacement curves to failure of single-row,

    multi-fastener joint configurations were obtained. This method was validated against

    experimental data and was found to provide accurate results for both static and dynamic

    loading conditions.

    In Chapter 4, the analytical model presented in Chapter 3 was applied to a custom, two-

    dimensional finite element code with a novel highly efficient solver methodology. This

    allowed for the highly efficient analysis of two-dimensional fastener arrays. Using the

    parameterised damage approximation function it was possible to vary the damaged

    response at each fastener location. Using this approach, the statistical variations in energy

    absorption and failure loads that were observed experimentally were applied to the finite

    element model to investigate what effect this had in multi-fastener joints.

    Chapter 5 presents the development of a highly efficient user-defined finite element for

    modelling the three-dimensional behaviour of single-shear joints. The element was easily

    implemented in commercial FE software and was validated against experimental data and

    results from detailed three-dimensional FE analyses. The element is capable of capturing

    the full non-linear shear load-displacement behaviour of the joint, in addition to

    accounting for the stiffness of the clamped region when loaded in the transverse direction.

    The element was developed specifically for use in an explicit FE solver, making it

    particularly suited for dynamic problems.

  • Chapter 1

    8

    In Chapter 6, the effect of fibre orientation and material anisotropy on load distribution

    in multi-fastener joints was investigated. Using a novel qualitative visualisation method

    trajectories of bearing load were illustrated. The user-element developed in Chapter 5 was

    also applied in this chapter to investigate the effects of missing fasteners and relative bolt

    spacing on load distribution in multi-fastener joints.

    Finally, in Chapter 7 concluding remarks and recommendations for future work are

    presented. The methods developed in this thesis were discussed in the context of their

    potential applications in relation to the certification of composite aircraft. The findings

    made in Chapters 3, 4, 5 and 6 regarding the mechanical response of bolted joints are

    summarised and a number of design recommendations are presented for multi-fastener

    joints.

  • Chapter 2

    LITERATURE REVIEW

  • Chapter 2

    10

    2.1 Introduction

    Bolted joints are an important consideration in the design of both composite and

    metallic structures. Consequently, extensive research has been carried out on this topic

    over the last number of decades. Mechanical behaviour, development of experimental

    techniques and the development of numerical and analytical design tools are aspects

    which have been investigated.

    This literature review is divided into three main parts. Firstly, publications pertaining

    to the mechanical response of bolted joints are discussed. Following an introduction

    to the terminology related to bolted joints, a description of common failure modes is

    provided. Publications which discuss the effects of design variables such as geometry,

    composite layup, clearance and loading rate are then reviewed. The second section

    provides a brief discussion on the capabilities of three-dimensional finite element (3D-

    FE) analyses of bolted joints. Specifically, the ability to capture both the undamaged

    response of the joint in addition to the prediction and simulation of joint failure was

    examined. Finally, efficient modelling strategies for bolted joints are reviewed.

    Analytical models in addition to highly simplified global modelling techniques were

    discussed. To supplement this review of literature, the main points from each section

    are summarised at the end of the chapter.

    2.2 Mechanical Behaviour of Bolted Joints

    Due to the interaction of multiple components and a significant number of variables

    associated with the problem, predicting the behaviour of bolted joints is a non-trivial

    task. A large number of experimental and numerical studies have been carried out to

    investigate the influence of various factors on the response of composite bolted joints.

    This section provides a summary of these studies in addition to an introduction to the

    terminology relevant to bolted joints.

    2.2.1 Terminology

    In this section, terms pertaining to the geometric configuration, loading, and failure of

    bolted joints are presented. Figure 2.1 illustrates a number of commonly used

    geometric parameters. The length of the laminate, L, consists of two regions, the

    overlap and the non-overlap region. The overlap region is the section of laminate in

    contact with another laminate, and this interface is referred to as the shear plane. It is

  • Chapter 2

    11

    possible for the joint to have more than one shear plane, as is the case for the double-

    lap joint in Figure 2.1. However, the single shear bearing configuration is

    representative of most aircraft bolted joint applications (MIL-HDBK-17 2002). This

    is due to the bending and shear loads which are induced on the fastener, while the

    double-lap joint induces mostly shear loads. Hole diameter, width and edge distance

    are defined as d, w, and e respectively while laminate thickness is denoted by the

    variable, t. In multi-fastener joints, two additional parameters are used to define the

    position of the bolts relative to each other. These are the distances between the hole-

    centres along the loading direction, pr (row pitch), and transverse to the loading

    direction, pc (column pitch). For the entirety of this thesis, “rows” refer to fasteners

    groups aligned with the main loading direction while “columns” are oriented

    transverse. To assist in the explanation of failure modes in bolted joints, failure planes

    have been illustrated in Figure 2.1, these include the bearing plane, net-tension plane

    and shear-out plane. Failure modes in bolted joints will be discussed in detail in

    Section 2.2.2.

    Figure 2.1 Definition of geometric parameters of bolted joints (redrawn from McCarthy (2003)).

    A number of additional terms are used to describe the strength and efficiency of joints.

    Bolted joints are often analysed in terms of load-displacement curves or bearing stress-

    strain curves. The displacement undergone by the joint is typically taken as the

    crosshead displacement of the test machine, unless extensometers are fitted to the

    specimen. The total load carried by the joint is taken from a load-cell affixed to one of

    the crossheads. There is no standard method to determine the loads carried by

    individual bolts, however some methods include the use of instrumented fasteners

  • Chapter 2

    12

    (Lawlor 2004; Ekh and Schön 2006) and from the measurement of surface strains

    (Starikov and Schön 2002; Lawlor et al. 2005). These will be discussed in more depth

    in Section 2.2.9.

    To determine bearing strength, bearing stress-strain curves may be derived. From

    MIL-HDBK-17 (MIL-HDBK-17 2002) relating to the characterisation of polymer-

    matrix composites and from ASTM Standard D5961/D5961M-08 (ASTM 2008) on

    testing the bearing response of composite materials, in a single bolt joint, bearing stress

    is defined as:

    dt

    Pb (Eq 2.1)

    Where P is the load carried by the joint. Bearing strain is defined as the ratio of the

    deformation of the bearing hole in the direction of the applied force to the pin diameter

    is defined as:

    dk

    b21

    (Eq 2.2)

    Where δ1 and δ2 are the displacements measured at extensometers 1 and 2 respectively,

    and k is the “force-per-hole factor”, set to either 1.0 for single-lap or 2.0 for double-

    lap joints. A final term used in the discussion of bolted joints is the joint efficiency

    which is defined by Eq. (2.3) (Hart-Smith 2003). This is the ratio of joint strength to

    that of the parent base material.

    StrengthLaminate notched-Un

    StrengthJoint Efficency Joint (Eq 2.3)

    2.2.2 Failure Modes of Bolted Joints

    When designing a joint, or any structure, it is important to consider the modes of

    failure which it can undergo. The six main failure modes in bolted joints, are illustrated

    in Figure 2.2. It is important to note that the occurrence these failure modes are not

    random and have been found to depend strongly on geometry and layup. The

    dependence of failure modes on these parameters is the subject of Sections 2.2.3 to

    2.2.7. However, in this section a brief description of each failure mode is provided in

    addition to a short discussion on their desirability and the micromechanical

    mechanisms leading to failure.

  • Chapter 2

    13

    Bearing failure is a progressive failure mode that occurs in the material immediately

    adjacent to the points of contact between the fastener and the laminate. It is thought to

    occur when the ratio of d/w is low or the ratio of by-pass to bearing load is small.

    Bearing failure was also found to be strongly affected by clamping pressure (Eriksson

    et al. 1995). A characteristic of this failure mode is the non-linear load displacement

    behaviour before final failure which arises from the accumulation of damage.

    Camanho et al. (1998) found bearing failure to initiate with localised delamination.

    As load increases matrix cracking occurs in the 90° and 45° plies with these cracks

    providing preferential starting points for other damage mechanisms. Fibre micro-

    buckling was also observed to occur in 0° plies and appeared to be related to the

    presence of matrix cracks in the 45° plies.

    Figure 2.2 Failure modes of mechanically fastened joints (redrawn from Niu (1992)).

    Net-tension failure is a catastrophic failure mode which occurs without significant

    warning. This mode of failure is most likely to occur when the ratio of d/w is high, or

    when the ratio of by-pass to bearing load is high (Eriksson et al. 1995). Camanho et

    al. (1998) found no significant damage in joints designed to fail by this mode until

  • Chapter 2

    14

    90% of the ultimate failure load was carried. It was at this load that delamination in

    the 90°/45° interfaces began to occur. Additionally matrix cracking in the off-axis

    plies was present and lead to non-linearity in the load-displacement curve.

    Shear-out failure is another catastrophic mode that provides little warning before the

    complete failure of the joint. Typically characterised by a linear load-displacement

    relationship until final failure (Camanho et al. 1998), shear-out is prone to occur in

    highly orthotropic laminates and those with short end distances (Eriksson et al. 1995).

    Similar to net-tension failure, Camanho et al. (1998) found damage to occur as late as

    90% of the ultimate failure load, when at this stage crushing and delamination became

    noticeable. Cleavage failure is a combination of the net-tension and shear-out modes

    and occurs due to the proximity of the hole to the end of the specimen (Hart-Smith

    2003; MIL-HDBK-17 2002). This type of failure often initiates at the end of the

    specimen rather than adjacent to the fastener (MIL-HDBK-17 2002).

    Fastener pull-through is an out-of-plane failure mode which is characterised by a linear

    load-displacement response up to the ultimate load (Waters and Williams 1985;

    Banbury and Kelly 1999; Kelly and Hallström 2005). Through-thickness failure at the

    edge of the fastener head and subsequent in-plane delamination are the damage

    mechanisms characterised by this mode of failure (Banbury and Kelly 1999). Fastener

    pull-through is highly dependent on the fastener, being frequently associated with

    countersunk joints (MIL-HDBK-17 2002). Resistance to pull-through failure can be

    increased by increasing the fastener head diameter (Banbury and Kelly 1999).

    Fastener failure is seen as a premature mode of failure and is generally considered

    undesirable. McCarthy et al. (2002) observed joints which initially failed in bearing

    ultimately failed due to fastener failure. Egan et al. (2013) also observed bolt failure

    in single-lap, countersunk joints which was attributed to a large bending moment

    reacted by the fastener head as a result of the thickness of the laminate.

    Of the failure modes illustrated in Figure 2.2 only bearing and net-tension are

    considered desirable, as all other modes result in premature failure (Hart-Smith 2003).

    It was found by Hart-Smith (2003) that joints designed to fail via net tension tend to

    carry the greatest loads. Despite this, joints are typically designed to initially fail in

    bearing as this is a progressive mode, allowing damage to be detected before

    catastrophic failure of the joint occurs.

  • Chapter 2

    15

    2.2.3 Material and Lay-up

    A well-known characteristic of composite bolted joints is a lower structural efficiency

    when compared to their metallic counterparts. For comparison, optimal designs of

    composite joints rarely exceed a structural efficiency of 40%. However joints

    manufactured from ductile metals such as aluminium can reach in excess of 60%. The

    reduction in efficiency is generally attributed to higher stress concentrations associated

    with orthotropic materials (Collings 1977; Hart-Smith 2003). When only structural

    efficiency is considered, there is a clear advantage to using conventional materials.

    However, Collings (1977) compared the ratios of specific strengths of CFRP and

    metallic joints and it was observed that composites offer a potential advantage of 32-

    52% when compared to aluminium alloys and 79-107% for steel. Thus when selecting

    composites over conventional materials, structural efficiency in the joint region is

    sacrificed for a superior strength-to-weight ratio.

    Fibre orientation in multidirectional laminates has been found to play an important

    role in determining the failure mode in bolted joints. Collings (1977) found that large

    numbers of ±45° plies could reduce the stress concentration around the hole by

    imparting a degree of softening in the joint. Arnold et al. (1990) also emphasised the

    importance of 45° plies as a mechanism of diffusing loads around the bolt hole, while

    Eriksson (1990) found laminates that used higher proportions of ±45° plies yielded

    higher bearing strength. Experiments from Collings (1977) suggest that the best tensile

    performance is achieved when using 30-50% 45° plies in [0°/ ±45°] laminates.

    Furthermore, Kelly and Hallström (2004) observed that a high percentage of ±45°

    acted to inhibit pure shear-out failure in the joint sections tested. From a review by

    Hart-Smith (2003), it was noted that shear-out and cleavage failures can arise, in part,

    due to a highly orthotropic laminate pattern or if there is insufficient dispersion of the

    differently oriented plies. It was recommended that fibre patterns be fully dispersed to

    maximise resin interfaces between changes in direction of layers of fibres.

    Orientations in a multidirectional laminate should consist of at least 12.5% of plies in

    each of the four directions: 0°, ±45° and 90° (Arnold et al. 1990; Hart-Smith 2003).

    Both Wang et al. (1996) and Park (2001) found joints with 90° plies on the surface of

    the laminate had greater delamination bearing strengths than those with 0°. This

    resulted from the tendency of 0° plies located on the surface to fail by breaking away

    from the laminate due to bearing loads (Wang et al. 1996).

  • Chapter 2

    16

    A benefit of the use of multidirectional laminates is the ability to tailor the orthotropic

    material properties to suit the loading conditions that the structure is expected to

    experience. This primarily consists of orientating a large proportion of 0° in the

    primary loading direction. Although experiments from Pakdil et al. (2007) found that

    bearing strength was maximum in zero-dominated layups, evidence from the majority

    of sources reviewed suggest that a quasi-isotopic layup is preferential to optimise the

    bearing response of laminates (Collings 1977; Hart-Smith 2003). Furthermore, Hart-

    Smith (2003) noted that joint strength varies less with the percentage of 0° plies than

    does the notched laminate strength.

    2.2.4 Geometry

    It is apparent that the performance of a joint is strongly governed by its geometry. This

    section reviews a number of published works which discussed the influence of various

    geometric ratios on ultimate bearing strengths and failure modes of bolted joints. Other

    important geometric parameters such as bolt-hole clearance and multiple fasteners are

    discussed in Sections 2.2.6 and 2.2.9 respectively.

    The effect of the w/d and e/d ratios was investigated by Cooper and Turvey (1995),

    Collings (1977), Hart-Smith (2003) , Kelly and Hallström (2004) and Pakdil et al.

    (2007). All sources found that increasing the w/d ratio resulted in a change in failure

    mode from net-tension to the more progressive bearing failure. However, the w/d ratio

    at which this transition occurs was observed to vary with layup and clamping pressure.

    Collings (1977) found the transition to bearing failure occurred at w/d ratios between

    3.5-4 in layups consisting of [0°/±45°], [0°/±60°] and [0°/90°] plies but at 7 for ±45°

    laminates. Kelly and Hallström (2004) found the critical w/d ratio for the transition to

    bearing failure was lower in pinned joints (w/d = 2) compared to clamped joints (w/d

    = 3). This was attributed to the lateral support at the edge of the hole inhibiting through

    thickness expansion.

    Increasing the e/d ratio was also found to result in a change in failure mode, where

    shear-out failure occurred at lower ratios while larger ratios promoted bearing failure

    (Collings 1977; Cooper and Turvey 1995; Kelly and Hallström 2004). Again, the

    critical e/d ratio resulting in this change in failure mode was found to be dependent on

    layup and clamping pressure, with the larger e/d ratios required in clamped joints

    (Kelly and Hallström 2004). The orientations which resulted in the lowest ratios were

  • Chapter 2

    17

    [0°/±45°], [0°/±60°] requiring an e/d ratio of 2.5-3 to promote bearing failure (Collings

    1977). For the joints tested by Kelly and Hallström (2004), it was found that once

    sufficiently large e/d and w/d ratios were used so that bearing failure occurred, further

    increase in these ratios resulted in very little change in the ultimate strength of the

    joint. This suggests that the ultimate strength of the joint is governed by the failure

    mode. Similar findings were made by Cooper and Turvey (1995).

    The effect of laminate thickness was also discussed in literature. Eriksson (1990),

    Kelly and Hallström (2004), Collings (1977) and Egan et al. (2013) found an increase

    in ultimate bearing strength with increasing t/d ratio. Additionally Egan et al. (2013)

    noticed a change in failure mode from one of fastener pull through to fastener fracture

    with increasing thickness. This was attributed to a greater bending moment reacted by

    the fastener head in thicker layups. However, the use of too small a bolt diameter

    results in excessive bolt bending and a highly non-uniform bending stress distribution

    which tends to promote pull-through failure (Hart-Smith 2003). To further the

    argument of using larger diameter fasteners, both Eriksson (1990) and Wang et al.

    (1996) found an increase in ultimate bearing stress with increasing bolt diameter, using

    a constant w/d ratio.

    2.2.5 Fastener Type

    Few authors have directly compared the response of joints fastened with protruding

    head and countersunk bolts. McCarthy et al. (2002) noted a reduction in stiffness in

    countersunk joints when compared to their protruding head counterparts. Furthermore,

    a 25-35% reduction in bearing strength was observed which was thought to be due to

    the presence of high stress concentrations between the bolt shank and hole. Hart-Smith

    (2003) also noticed these trends, citing the reduced effective bearing area in

    countersunk joints as a contributing factor.

    The removal of material to accommodate a countersunk head also has a significant

    effect on the failure behaviour of the joint. Gunnion et al. (2006) found the primary

    difference between the responses of the two head types was that protruding head

    fasteners exhibited a sudden transition from the linear elastic to damage region. In

    comparison this transition was much more gradual in countersunk specimens. Pearce

    (2009) observed different failure modes for the two head types. Countersunk joints

    were observed to fail via bolt failure at quasi-static rates but via bearing failure at

  • Chapter 2

    18

    higher rates. In comparison, protruding head joints failed via net-tension and cleavage

    when loaded quasi-statically and bearing-cleavage at higher rates. Of course these

    observations are specific to the material and geometry used by Pearce (2009), but they

    highlight the effect of fastener type on the failure response of the joint. In general,

    countersunk joints are more prone to bearing failure and fastener pull-through (Niu

    1992).

    2.2.6 Bolt-hole Clearance

    In any manufacturing process, there is a trade-off between maximising quality and

    minimising process costs. As a consequence of this, statistical variations from nominal

    hole and bolt diameters are to be expected. The difference between these values results

    in bolt-hole clearance, which can significantly affect the contact angle (see Figure 2.3).

    Clearance can also be expressed as a percentage as shown in Figure 2.3 (a), and values

    of 1.2% are typical of aircraft joints (Crews and Naik 1987). Numerous experimental

    and numerical studies have been carried out to determine the influence of clearance

    on the performance of composite bolted joints.

    Naik and Crews (1986) carried out 2D finite element (FE) analysis to determine the

    effect of clearance in pin-loaded composite joints. The FE model developed was

    compared to the continuum analysis of Eshwar (1977), which allowed the contact

    angle to be predicted as a function of pin load, and excellent agreement was observed

    for the case of an infinite isotropic plate. Clearances ranging from 0-1.6% were

    investigated by Naik and Crews (1986). For 1.6% clearance, peak radial and tangential

    stresses were found to increase by 36% and 16% respectively over the neat-fit case.

    The contact angle was found to decrease by 30% and 16% greater hole elongation was

    observed. It was concluded that clearance in mechanically fastened joints should be

    considered in stress and strength analyses but may have little influence on joint

    stiffness.

    DiNicola and Fantle (1993) tested double shear pin joints with clearances ranging from

    neat-fit to 279µm in addition to carrying out FE analyses. It was found that bearing

    strength at four percent hole deformation depended strongly on bolt-hole clearance,

    with the 279µm clearance case exhibiting a 30% reduction in strength. As expected,

    the distribution of, and peak contact forces between the pin and hole was greatly

  • Chapter 2

    19

    influenced by hole oversize. However, ultimate bearing stress of the joint was not

    found to be dependent on clearance.

    Figure 2.3 Definition of (a) clearance and (b) contact angle.

    Lawlor et al. (2002a) carried out experimental studies on single-shear, single-bolt

    graphite/epoxy joints with quasi-isotropic and zero-dominated layups. Nominal

    clearances of 0, 1, 2 and 3% were investigated. It was found that for finger-tight,

    protruding head bolts in quasi-isotropic layups there was a delay in load take up, which

    is slightly larger than the nominal clearance value. Similar findings were made for

    countersunk fasteners, although the repeatability in load take up for various clearances

    were not as good as those observed for protruding head bolts. In highly torqued joints,

    the clearance was found to be taken up after the effects of static friction had been

    overcome. All specimens considered initially failed in bearing, however greater

    ultimate strain was observed with increasing clearance. It was believed this increase

    in ultimate strain occurred as larger clearances resulted in more concentrated loads on

    the laminate, thus resulting in extensive laminate damage. As the laminate absorbs

    more energy, there is this less remaining to be absorbed by plastic deformation of the

    bolt. Kelly and Hallström (2004) also observed extensive bearing damage in joints

    with larger bolt-hole clearance. Similar to DiNicola and Fantle (1993), Lawlor et al.

    (2002a) and McCarthy et al. (2002), Kelly and Hallström (2004) found the effect of

    clearance on the ultimate strength of the joint to be negligible. In contrast, Pierron et

    al. (2000), found ultimate failure loads to decrease by approximately 30% in pin-

  • Chapter 2

    20

    loaded, woven glass fibre epoxy joints. However, this reduction in strength

    corresponded to a clearance of 12.5%.

    In Lawlor et al. (2002a), a reduction in joint stiffness was observed when clearance

    was increased. Comparing the joint stiffness of the neat-fit and maximum clearance

    cases, a 10% decrease in stiffness was noted. McCarthy et al. (2002) found that the

    reduction in stiffness in 1% clearance joints with protruding head fasteners was

    statistically insignificant, however in countersunk joints a stiffness reduction of

    approximately 5% was noted for the same clearance value. For 2% clearance, a 7-10%

    reduction in stiffness was observed for protruding head fasteners and 3-4% for

    countersunk. A reduction in joint stiffness in single-bolt clearance fit joints was also

    observed by McCarthy and McCarthy (2003) in 3D-FE models and experimentally by

    Kelly and Hallström (2004). In contrast, Naik and Crews (1986) and Pierron et al.

    (2000) concluded that clearance had little effect on joint stiffness, however this was

    not quantified by either author.

    Clearance in double-lap, multi-bolt joints was investigated experimentally by Lawlor

    et al. (2005). Using strain gauges positioned on the laminate surface, the effects of

    clearance on bypass loads were measured. It was found that bolts with larger

    clearances do not initially carry any load, resulting in the remaining fasteners carrying

    the majority of the joint load. Once the distance between the bolt and hole has been

    taken up, larger clearance fasteners begin to contribute to the load carrying capacity

    of the joint and the load distribution started to even out. McCarthy and McCarthy

    (2003) found similar results in their numerical models. However, Lawlor et al. (2005)

    found this process was to be interrupted by the onset of material failures. Clearance

    was discovered to have a significant effect on the initial failure loads in multi-bolt

    joints, i.e. the load at which bearing failure occurs in one of the holes. The initial

    failure load in a 3-bolt joint was found to decrease by 25% with non-uniform

    clearances at each hole, when compared to the neat fit case. However, as in single-bolt

    joints, clearance was not found to affect the ultimate load carried in multi-bolt

    configurations.

    2.2.7 Bolt Torque and Lateral Constraint

    The clamping effect that arises due to bolt torque can have a significant influence on

    the joint response. A number of authors have investigated the effect of lateral

  • Chapter 2

    21

    constraint and clamping pressure in bolted joints. In the work carried out by McCarthy

    et al. (2002), the effect of bolt torque was investigated. Two different torque levels

    were considered, 0.5 Nm to represent “finger tight” conditions and 16 Nm as the

    recommended in-service torque. In “finger tight” joints a distinct “knee” was observed

    in the load-deflection curves of quasi-isotropic layups. Based on the findings of Park

    (2001), and Kelly and Hallström (2004), this was most likely due to the onset of

    bearing damage, which explains why this occurrence was most noticeable in

    countersunk joints. In zero-dominated layups and highly torqued joints this knee was

    not evident. In highly torqued joints, a distinct transition from an initial high slope to

    a lower slope was noticed, with the length of the transition region approximately equal

    to the clearance in the joint, as illustrated in Figure 2.4. The initially high stiffness of

    Slope 1 was believed to be dominated by static friction, and the lower slope during the

    transition region from kinetic friction. McCarthy et al. (2002) noted that the linearity

    of this high stiffness region was debatable, however assuming only static friction to

    be dominant McCarthy and Gray (2011) developed an analytical model for highly

    torqued, multi-bolt joints with clearance. Excellent correlation was observed between

    the analytical model, experiments and 3D-FE analyses. In Figure 2.4, Slope 2 was due

    to the stiffness of the untorqued joint in addition to kinetic friction.

    Figure 2.4 Effect of bolt torque on initial load-displacement curve of single-bolt, single-lap joint with

    clearance (redrawn from McCarthy et al. (2002)).

    Using 3D-FE analyses, McCarthy et al. (2005a) modelled the behaviour of single-lap,

    protruding head, composite bolted joints manufactured from HTA/6376. Bolt torque

    was simulated by applying orthotropic thermal expansion to one of the washers, and

    friction from a stick-slip model. This stick slip friction model attempted to capture the

  • Chapter 2

    22

    discontinuity in the idealised Coulomb model through a series of conditional

    statements.

    Schön (2004a) carried out a series of experiments on HTA/6376 and found the

    coefficient of friction for composite to composite contact varied between 0.65 and

    0.74. The same author (Schön 2004b), also found the coefficient of friction between

    composite and aluminium to have an initial value of 0.23 and peak value of 0.68 after

    wear in. Tsukizoe and Ohmae (1986) found that the coefficient of friction between

    mild steel and unidirectional carbon fibre-reinforced epoxy was dependent on the fibre

    volume fraction and varied between 0.1 and 0.3. Based on these publications,

    McCarthy et al. (2005a), used the following coefficients of friction: µ=0.7 for laminate

    to laminate, µ=0.3, for washer (steel)-to laminate and µ=0.1 for bolt (titanium) to

    laminate contact. Egan et al. (2015) used these same properties in the modelling of

    dynamically loaded countersunk M21E/IMA composite joints. The stick slip

    algorithm used in McCarthy et al. (2005a) required additional numerical parameters

    which were tuned to allow convergence of the finite element model and to provide

    accurate results.

    McCarthy et al. (2005a) found that Slope 1 in Figure 2.4 arose as the load carried by

    the joint was almost entirely reacted by friction between the laminates. During the

    transition region, friction between the washer and laminate which previously had little

    effect on the response of the joint now account for approximately one third of the load

    carried. For Slope 2, the load carried by friction between the laminates drops slightly

    with the greatest portion of the load reacted by the normal contact between the bolt

    and hole. Friction between the fastener and hole was found to have a negligible effect

    on the load displacement history. This is best illustrated in Figure 2.5 (a). It was also

    found that after the effects of friction have been overcome and the majority of the load

    was carried by the contact between the bolt and hole, the total friction load starts to

    reduce. The reason given for this is that for single lap joints, as a consequence of bolt

    rotation, the plates start to pull apart resulting in a loss in normal load and hence

    friction force between the plates. This effect was also found to be dependent on hole

    clearance and is illustrated in Figure 2.5 (b).

  • Chapter 2

    23

    Figure 2.5 (a) Loads carried by various interactions in single-lap joint (b) Bolt and total friction load

    distribution in multi-bolt joint with variable hole clearance, C1=neat-fit, C4=240 µm clearance

    (redrawn from McCarthy et al. (2005a)).

    In a study carried out by Park (2001), clamping pressure was found to supress the

    onset of delamination around the hole in bolted joints, and as a result delamination

    failure tended to occur outside the clamped region. As a consequence of the

    suppression of delamination, the bearing failure which was observed to be a

    catastrophic failure mode in pinned joints, was progressive in bolted joints. It was

    noted that the ultimate and delamination strengths of the bolted joint were nearly

    double that of the pinned joints. Similar results were obtained by Kelly and Hallström

    (2004) where a number of pinned and bolted composite joints were tested

    experimentally. It was found that the lateral constraint imposed on the composite

    laminate by the fastener tended to supress a “brooming” type failure at the edge of the

    hole observed in pinned joints. The effect of lateral constraint on joint strength is

    illustrated in Figure 2.6 (a). Eriksson (1990), Cooper and Turvey (1995), Khashaba et

    al. (2006) and Pakdil et al. (2007) also found the bearing strength of the joint to

    increase with bolt torque.

    In a study carried out by Park (2001), it was found that above a nominal clamping

    pressure, there was little increase in the ultimate bearing strength of bolted joints. This

    is illustrated in Figure 2.6 (b) for three different composite layups. Furthermore it was

    noted that at a clamping pressure of approximately 0.1% of the nominal bolt torque,

    the ultimate bearing strength was almost double compared to that for pinned joints.

    These results suggest that the increase in strength of bolted joints over pinned joints

    was primarily due to the lateral constraint of the fastener, rather than the increased

    friction between the laminates. This was also observed by Nassar et al. (2007) in

    double-lap, composite-aluminium joints. Further evidence of this can be seen in Figure

  • Chapter 2

    24

    2.5 (b) where McCarthy et al. (2005a) showed that the friction in the joint decreases

    as loading progresses. However, Park (2001) also noted that although increasing

    clamping pressure above a nominal value did not significantly change the ultimate

    bearing strength of the joint, the delamination bearing strength continued to increase

    with increasing clamping pressure. In contrast to Park (2001), McCarthy and Gray

    (2011) found that joints with larger values of torque carried greater ultimate loads.

    This was attributed to the onset of critical bearing damage occurring at higher load

    levels due to friction forces at the shear plane.

    Figure 2.6 (a) Load-displacement behaviour of pinned and clamped laminated (redrawn from Kelly

    and Hallström (2004)); (b) Effect of clamping pressure on ultimate bearing strength of bolted joints

    (redrawn from Park (2001)).

    2.2.8 Loading Rate

    A vast number of experimental studies have been carried out to investigate the strain

    rate effects in materials and mechanically fastened structures. This section provides a

    brief summary of the studies of relevance to the subject matter of this thesis. Firstly a

    summary on the findings of rate effects in the parent material, specifically the

    composite is given followed by experimental findings in bolted joints.

    The effect of strain rate on the mechanical behaviour of composite materials has been

    investigated by a number of authors using various experimental techniques. Using

    drop-tower testing and a servohydraulic test machine, Hsiao and Daniel (1998)

    investigated the rate effects on the compressive and shear behaviour of carbon/epoxy

    composites for strain rates up to 300s-1. A significant enhancement in matrix-

    dominated properties was observed for higher loading rates. Shear modulus and

    strength were found to increase by 18% and 80% respectively, while the transverse

  • Chapter 2

    25

    modulus and strength increased by 37% and 100%. The ultimate strain at failure for

    both shear and transverse loads appeared to be rate independent. Figure 2.7 illustrates

    the effects on strain rate observed by Hsiao and Daniel (1998) in matrix-dominated

    properties. In a similar study carried out by Hsiao et al. (1998), the variation of both

    strength and modulus with strain rate was investigated further. Between quasi-static

    and rates of 10s-1, a linear dependence on strength/modulus was observed with the log

    of strain rate. However, for strain rates above 10s-1, strength/modulus appeared to

    increase exponentially. When the longitudinal properties were investigated only a

    slight increase in the compressive modulus was reported, however strength and

    ultimate strain values increased by up to 79% and 74% respectively. For strain rates

    up to approximately 100s-1, only a linear variation in strength with the log of strain

    rate was observed.

    Using a Split Hopkinson Pressure Bar (SHPB), Koerber et al. (2010) investigated the

    strain rate effects on the transverse compression and in-plane shear properties of

    carbon-epoxy composites for strain rates up to 350s-1. The trends observed in terms of

    modulus and strength were comparable to those observed by Hsiao and Daniel (1998).

    Using SHPB testing Hosur et al. (2001) investigated the high strain rate compression

    response on unidirectional and cross-ply carbon epoxy laminates up to rates of 817s-1.

    For all cases tested, an increase in stiffness was observed when loading rate was

    increased from quasi-static to dynamic. However, in cross-ply laminates (loaded in

    plane and through thickness) and unidirectional laminates loaded along 0°, a decrease

    in stiffness was observed when strain rate was increased from 163s-1 to 817s-1. A

    decrease in peak stress was also noted in the case of cross-ply laminates loaded in-

    plane. Additionally, the peak stress in all cross-ply laminates loaded dynamically

    through the thickness was lower than the equivalent quasi-static case. However, the

    behaviour of the 90° plies was similar to that observed in Hsiao and Daniel (1998) and

    Koerber et al. (2010), in that both stiffness and strength were seen to increase with

    strain rate.

  • Chapter 2

    26

    Figure 2.7 Quasi-static and high rate stress strain curves for: (a) Transverse compression and; (b)

    shear in unidirectional carbon/epoxy composites (redrawn from Hsiao and Daniel (1998)).

    Groves et al. (1992) tested carbon/epoxy composites in tension, shear and

    compression up to rates of 3000s-1. Their findings supported those of Hsiao et al.

    (1998), specifically the exponential-like increase in stiffness and strength beyond

    strain rates of 10s-1. Several theories were postulated to explain the rate-dependent

    behaviour observed in polymer-matrix composites. Hsiao and Daniel (1998) proposed

    that the time-dependent nature of accumulating damage could be one source of the

    strain rate effects observed. At lower rates, damage accumulates more gradually, such

    that a well-defined non-linear region occurs near the end of the stress