Вы находитесь на странице: 1из 332

ADVANCED ANALYSIS AND

PROBABILISTIC-BASED DESIGN OF
SUPPORT SCAFFOLD SYSTEMS

BY

TAYAKORN CHANDRANGSU
B.Sc. (Civil Engineering), University of Wisconsin Madison, USA
M.Sc. (Structural Engineering), Stanford University, USA

A thesis submitted in fulfilment


of the requirements for the degree of Doctor of Philosophy
School of Civil Engineering
The University of Sydney
Australia

2010 Tayakorn Chandrangsu


CERTIFICATE OF ORIGINALITY

I certify that the work in this thesis has not previously been submitted for a degree nor
has it been submitted as part of requirements for a degree except as fully acknowledged
within the text.

I also certify that the thesis has been written by me. Any help that I have received in my
research work and the preparation of the thesis itself has been acknowledged. In
addition, I certify that all information sources and literature used are indicated in the
thesis.

Tayakorn Chandrangsu
Abstract i
______________________________________________________________________

ABSTRACT

Support scaffold systems are used to provide temporary underpinning in the


construction of bridges, buildings, and similar structures. They are used as platforms
while building formwork for reinforced concrete, and are usually heavily loaded under
the weight of wet poured concrete. This thesis presents comprehensive studies into the
direct design of steel support scaffold (Cuplok) systems by advanced geometric and
material nonlinear analysis.

The research program includes the collection and statistical evaluation of on-site data of
member out-of-straightness (member crookedness), frame out-of-plumb, and loading
eccentricity, as well as eighteen full-scale load tests of 3x3 bay subassemblies and a
comprehensive series of tests on Cuplok joints in various configurations to determine
statistical distributions for the strength, semi-rigid stiffness and looseness. Advanced
finite element models are calibrated using the full-scale subassembly tests, hence
providing statistical data for the modelling error. In addition, parametric studies are
performed to determine any factors affecting the system strength. Having procured the
statistical data from on-site survey and experimental tests in the laboratory for the main
random variables affecting the strength of support scaffold systems, Monte-Carlo
simulations using advanced analysis are carried out to obtain the statistical distributions
of system strength for a range of geometric configurations of support scaffold systems.
A trend is emerging from these studies suggesting that the failure mode is mainly
determined by the extension of the jacks at the top and bottom of the frame as well as
the lift height. The first-order second-moment and first-order reliability methods are
employed for calculating the system resistance factors for support scaffold systems in
various configurations, leading to the formulation of a design method for support
scaffold systems based on advanced analysis, which requires no recourse to a steel
design specification for checking member and connection strengths.
Acknowledgements ii
______________________________________________________________________

ACKNOWLEDGEMENTS

The research work presented in this thesis was undertaken under the supervision of
Professor Kim Rasmussen at the School of Civil Engineering, University of Sydney. I
wish to express my sincere gratitude to Professor Kim Rasmussen for his guidance,
precious comments and continuous support during the whole study. I am indebted to Dr.
Hao Zhang for many valuable discussions and suggestions on the reliability study as
part of this thesis. I wish to thank laboratory technicians, Mr. Garry Towell, Mr. Paul
Bustra, and Mr. Brett Jones, for their kind assistance in the experimental tests. I would
also like to thank Boral staff, Mr. Hamid Pourbozorgi, Mr. Shahram Ebrahimian, and
Mr. Ranji Premaratne, for their advice, support and involvement in this research project.
I also like to thank Boral Formwork & Scaffolding Pty. Ltd. for making scaffold
subassembly test data available.

During the period of this study, my colleagues at the school and in Sydney have given
me their sincere support, which is deeply appreciated. Special thanks are directed to Dr.
Niphan Yaiaroon, Dr. Benoit Gilbert, Mr. Safat Al-deen, Mr. Cao Pham, Mr. Binh
Nguyen, Dr. Sawekchai Tangaramvong, Dr. Kamlaitip Pattapong, Dr. Krit Chaimoon
and Dr. Nida Seelsaen.

This research was funded by the Australian Research Council as part of the ARC
Linkage Project LP0884156 between the University of Sydney and Boral Formwork
and Scaffolding Pty. Ltd. I would like to thank these organisations for their financial
support through the APA (Industry) and CASE supplementary scholarships.

Finally, I would like to thank my parents, Dr. Karoon and Mrs. Sirikwan Chandrangsu,
my sisters, Dr. Kulsiri Chandrangsu-Ferrand and Ms. Chandrapa Chandrangsu, and my
brother-in-law, Dr. David Ferrand, for their greatest love, support and motivation in
every step of my life. I thank Ms. Panchita Noppisanvong for her encouragement and
support throughout my PhD study.
Table of Contents iii
______________________________________________________________________

TABLE OF CONTENTS

ABSTRACT.......................................................................................................................i
ACKNOWLEDGEMENTS ..............................................................................................ii
LIST OF FIGURES .........................................................................................................vi
LIST OF TABLES .........................................................................................................xiii
ABBREVIATIONS ........................................................................................................xv
PRINCIPAL NOTATIONS ...........................................................................................xvi
CHAPTER 1 - INTRODUCTION ....................................................................................1
1.1 Problem Statement ..................................................................................................1
1.2 Scope and Objectives ..............................................................................................4
1.3 Outline of Thesis .....................................................................................................5
1.4 List of Publications .................................................................................................8
CHAPTER 2 - LITERATURE REVIEW .......................................................................11
2.1 Introduction ...........................................................................................................11
2.2 Scaffold Systems...................................................................................................11
2.2.1 Scaffold Configurations .................................................................................12
2.2.2 Materials.........................................................................................................16
2.3 Collapses of the Scaffolds.....................................................................................17
2.3.1 Construction Stages........................................................................................17
2.3.2 Method of Determining Causes of the Collapses...........................................17
2.3.3 Main Causes of the Collapses ........................................................................17
2.4 Analysis and Modelling of Scaffold Systems .......................................................18
2.4.1 Nonlinear Structural Analysis ........................................................................18
2.4.2 Three-Dimensional Model vs. Two-Dimensional Model ..............................19
2.4.3 Load Combinations and Load Paths ..............................................................19
2.4.4 Initial Imperfections .......................................................................................21
2.4.5 Joint Modelling and Boundary Conditions ....................................................22
2.4.6 Suggestions ....................................................................................................26
2.5 Ultimate Load of Scaffold Systems ......................................................................27
2.5.1 Parametric Studies..........................................................................................27
2.5.2 Failure Modes.................................................................................................29
2.5.3 Simplified Equations......................................................................................31
2.6 Design of Scaffold Systems ..................................................................................35
2.6.1 British Standards ............................................................................................35
2.6.2 Australian Standards ......................................................................................37
2.6.3 Effective Lengths ...........................................................................................39
2.6.4 Bracing Systems.............................................................................................39
2.6.5 Design by Advanced Structural Analysis ......................................................40
2.6.6 Reliability Analysis........................................................................................41
2.6.7 Safety in Construction of Scaffold Systems...................................................47
2.7 Conclusions ...........................................................................................................48
CHAPTER 3 - GEOMETRIC IMPERFECTIONS AND LOADING ECCENTRICITY
OF SUPPORT SCAFFOLD SYSTEMS.........................................................................49
3.1 Introduction ...........................................................................................................49
3.2 Methods of Procurement .......................................................................................51
3.3 Survey Results of Geometric Imperfections and Loading Eccentricity................54
3.4 Discussion .............................................................................................................56
Table of Contents iv
______________________________________________________________________

3.5 Conclusions ...........................................................................................................60


CHAPTER 4 - CUPLOK JOINT TESTS .......................................................................61
4.1 Introduction ...........................................................................................................61
4.2 Test Setup..............................................................................................................63
4.3 Test Materials........................................................................................................66
4.4 Test Series .............................................................................................................67
4.5 Test Procedure.......................................................................................................69
4.6 Test Results ...........................................................................................................70
4.7 Discussion .............................................................................................................81
4.8 Conclusions ...........................................................................................................85
CHAPTER 5 - FORMWORK SUBASSEMBLY TESTS..............................................86
5.1 Introduction ...........................................................................................................86
5.2 Setup and Procedures ............................................................................................86
5.3 Test Configurations...............................................................................................89
5.4 Test Results ...........................................................................................................93
5.5 Conclusions ...........................................................................................................96
CHAPTER 6 - ADVANCED STRUCTURAL ANALYSIS MODELS OF SUPPORT
SCAFFOLD SYSTEMS .................................................................................................97
6.1 Introduction ...........................................................................................................97
6.1.1 Advanced Analysis ........................................................................................99
6.1.2 Previous Scaffold Models ............................................................................100
6.2 Finite Element Models ........................................................................................105
6.2.1 Spigot Joints .................................................................................................105
6.2.2 Semi-Rigid Standard-to-Ledger Connections ..............................................107
6.2.3 Brace Connections .......................................................................................109
6.2.4 Base Plate Eccentricity.................................................................................109
6.2.5 Loading Eccentricity ....................................................................................110
6.2.6 Geometric Imperfections..............................................................................111
6.2.7 Geometric and Material Nonlinearities........................................................112
6.2.8 Calibrations ..................................................................................................117
6.3 Discussion ...........................................................................................................128
6.4 Conclusions .........................................................................................................130
CHAPTER 7 - PARAMETRIC STUDIES OF SUPPORT SCAFFOLD SYSTEMS..131
7.1 Introduction .........................................................................................................131
7.2 Analysis Models for Parametric Studies .............................................................132
7.3 Loading Eccentricity ...........................................................................................137
7.4 Number of Lifts...................................................................................................139
7.5 Bracing Arrangements ........................................................................................141
7.6 Cuplok Joint Stiffness .........................................................................................142
7.7 Geometric Imperfections.....................................................................................144
7.8 Yield Stresses of the Main Components .............................................................147
7.9 Cross Sections of the Main Components ............................................................149
7.10 Discussions........................................................................................................150
7.11 Conclusions .......................................................................................................151
CHAPTER 8 - PROBABILISTIC ASSESSMENT OF SUPPORT SCAFFOLD
SYSTEMS.....................................................................................................................152
8.1 Introduction .........................................................................................................152
8.2 Analysis Models for Probabilistic Assessment ...................................................153
8.3 Uncertainties in Support Scaffold Systems.........................................................158
Table of Contents v
______________________________________________________________________

8.3.1 Cross-Sectional Area and Moment of Inertia...............................................158


8.3.2 Yield Stress ..................................................................................................159
8.3.3 Joint Stiffness ...............................................................................................159
8.3.4 Initial Geometric Imperfections ...................................................................161
8.3.5 Loading Eccentricity ....................................................................................163
8.4 Ultimate Strength of Support Scaffold Systems .................................................164
8.4.1 Monte Carlo Simulation Results ..................................................................164
8.4.2 Nominal Design Models ..............................................................................186
8.4.3 Modelling Uncertainty and Statistics of System Resistance........................188
8.5 Reliability Analysis with Design by Advanced Analysis ...................................191
8.5.1 Load Statistics ..............................................................................................194
8.5.2 System Resistance Factors ...........................................................................196
8.6 Conclusions .........................................................................................................200
CHAPTER 9 - DESIGN OF SUPPORT SCAFFOLD SYSTEMS ..............................201
9.1 Introduction .........................................................................................................201
9.2 Practical Advanced Analysis Models .................................................................202
9.2.1 Initial Geometric Imperfections ...................................................................203
9.2.2 Loading Eccentricity ....................................................................................205
9.2.3 Joint Stiffness ...............................................................................................206
9.2.4 Boundary Conditions ...................................................................................208
9.2.5 Material Model.............................................................................................208
9.3 Advanced Analysis and Design Procedures........................................................210
9.3.1 Load Combination........................................................................................210
9.3.2 Preliminary Design of Support Scaffold Systems........................................211
9.3.3 Modelling of Elements.................................................................................211
9.3.4 Determination of Load Increments ..............................................................211
9.3.5 Ultimate Strength Design Check..................................................................212
9.3.6 Adjustment of System Configuration...........................................................213
9.4 Design Comparisons ...........................................................................................213
9.5 Design Examples.................................................................................................214
9.5.1 Case Study 1.................................................................................................215
9.5.2 Case Study 2.................................................................................................218
9.6 Conclusions .........................................................................................................221
CHAPTER 10 - CONCLUSIONS ................................................................................223
10.1 Summary ...........................................................................................................223
10.2 Remarks ............................................................................................................224
10.3 Recommendations for Future Research ............................................................228
REFERENCES..............................................................................................................229
APPENDIX A.1 - OUT-OF-STRAIGHTNESS DATA ...............................................233
APPENDIX A.2 - OUT-OF-PLUMB DATA...............................................................237
APPENDIX A.3 - LOADING ECCENTRICITY DATA ............................................241
APPENDIX B.1 - JOINT STIFFNESS COMPARISONS ...........................................243
APPENDIX B.2 - PROBABILISTIC MODELS FOR JOINT STIFFNESS................258
APPENDIX C.1 - INITIAL IMPERFECTIONS OF SUBASSEMBLY TESTS .........260
APPENDIX D.1 - TYPICAL STRAND7 API SUBROUTINE FOR MONTE CARLO
SIMULATIONS............................................................................................................291
List of Figures vi
______________________________________________________________________

LIST OF FIGURES

Figure 1.1: Typical support scaffold systems ...................................................................2


Figure 1.2: Timber bearer supported by the U-head screw jack .......................................2
Figure 2.1: Typical scaffold systems: (a) access scaffold, and (b) support scaffold ......12
Figure 2.2: Various types of scaffold unit: (a) simple (knee-braced) door type; (b)-(e)
standard door type; (f) stick construction with Cuplok joints or wedge-type joints.......12
Figure 2.3: Configuration of typical stick-construction scaffold frame..........................13
Figure 2.4: Schematic of spigot joint ..............................................................................14
Figure 2.5: Schematic of (a) Cuplok joint; and (b) wedge-type joint .............................14
Figure 2.6: Schematic of brace connections: (a) hook connection; and (b) pin connection
.........................................................................................................................................15
Figure 2.7: Schematic of jack base .................................................................................15
Figure 2.8: Schematic of U-head jack.............................................................................16
Figure 2.9: Top view of rectangular, L, and U shapes of scaffold systems (adapted from
Peng et al. 2003)..............................................................................................................20
Figure 2.10: P- and P- effects .....................................................................................22
Figure 2.11: Typical moment-rotation curves for Cuplok and wedge-type joints by
Godley and Beale (1997) and Godley and Beale (2001) ................................................23
Figure 2.12: Spigot joint model (Enright et al. 2000) .....................................................25
Figure 2.13: Model of 2-bay shoring system (adapted from Peng 2004) .......................28
Figure 2.14: Typical failure mode of single storey door-type scaffold (adapted from Yu
et al. 2004).......................................................................................................................29
Figure 2.15: Schematic failure modes of one-to-three storey knee-braced scaffolds
(adapted from Huang et al. 2000a)..................................................................................30
Figure 2.16: Assumption of proposed analytical model (adapted from Huang et al.
2000c)..............................................................................................................................31
Figure 2.17: Computational critical loads based on two-dimensional model (adapted
from Huang et al. 2000a) ................................................................................................33
Figure 2.18: Model for approximating moment of inertia of scaffold (adapted from Peng
et al. 1998).......................................................................................................................34
Figure 2.19: Two layer shoring system with V-type and N-type bracings (adapted from
Peng 2004) ......................................................................................................................40
Figure 2.20: Basic structural reliability problem (adapted from Ellingwood and
Galambos 1982) ..............................................................................................................42
Figure 3.1: Cuplok scaffold systems...............................................................................51
Figure 3.2: Schematic of device used to measure out-of-straightness ............................52
Figure 3.3: Pictures of (a) device used to measure out-of-straightness (b) actual
measurement ...................................................................................................................53
Figure 3.4: Theodolite employed for storey out-of-plumb measurement .......................54
Figure 3.5: Histogram of normalised out-of-straightness of the standards (/Lh) ..........55
Figure 3.6: Histogram of normalised storey out-of-plumb (/H)...................................55
Figure 3.7: Histogram of loading eccentricity ................................................................56
Figure 3.8: Fitted lognormal distribution of normalised out-of-straightness of standards
without spigot joints........................................................................................................58
Figure 3.9: Fitted lognormal distribution of normalised out-of-straightness of standards
with spigot joints.............................................................................................................59
List of Figures vii
______________________________________________________________________

Figure 3.10: Fitted normal distribution of normalised storey out-of-plumb...................59


Figure 3.11: Fitted lognormal distribution of loading eccentricity .................................60
Figure 4.1: Typical Cuplok system .................................................................................61
Figure 4.2: Typical Cuplok components.........................................................................62
Figure 4.3: Locking mechanism of Cuplok ....................................................................63
Figure 4.4: Schematic of Cuplok joint test setup ............................................................64
Figure 4.5: Typical Cuplok joint test setup.....................................................................65
Figure 4.6: Test equipment setup ....................................................................................65
Figure 4.7: Typical Cuplok standard and ledger.............................................................66
Figure 4.8: Joint stiffness axes........................................................................................67
Figure 4.9: Joint configurations and loading directions in top view...............................68
Figure 4.10: Typical numberings and location of LVDTs on test specimen ..................70
Figure 4.11: Cuplok test results for KzA1 ........................................................................72
Figure 4.12: Cuplok test results for KzA2 ........................................................................72
Figure 4.13: Cuplok test results for KzB1.........................................................................73
Figure 4.14: Cuplok test results for KzB2.........................................................................73
Figure 4.15: Cuplok test results for KzC1.........................................................................74
Figure 4.16: Cuplok test results for KzC2.........................................................................74
Figure 4.17: Cuplok test results for KzD1 ........................................................................75
Figure 4.18: Cuplok test results for KzD2 ........................................................................75
Figure 4.19: Cuplok test results for KyA1 ........................................................................76
Figure 4.20: Cuplok test results for KyB1 ........................................................................76
Figure 4.21: Cuplok test results for KyC1 ........................................................................77
Figure 4.22: Cuplok test results for KyC2 ........................................................................77
Figure 4.23: Cuplok test results for KyD1 ........................................................................78
Figure 4.24: Cuplok test results for KyD2 ........................................................................78
Figure 4.25: Cuplok components after being tested........................................................79
Figure 4.26: Tri-linear moment-rotation for the Cuplok joints.......................................83
Figure 5.1: Typical test frame from top view (CASE 2006) ..........................................88
Figure 5.2: Enlarged view of (a) top eccentricity and (b) bottom eccentricity (CASE
2006) ...............................................................................................................................89
Figure 5.3: Typical test configuration in plan and elevation view (CASE 2006)...........92
Figure 5.4: Test No. 8 setup (CASE 2006) .....................................................................93
Figure 5.5: Failure in (a) spigot and (b) jack (CASE 2006)............................................95
Figure 6.1: Typical support scaffold ...............................................................................98
Figure 6.2: Schematic of spigot joint ..............................................................................98
Figure 6.3: Schematic of Cuplok joint ............................................................................99
Figure 6.4: Typical moment-rotation curves for Cuplok and wedge-type joints by
Godley and Beale (1997) and Godley and Beale (2001) ..............................................102
Figure 6.5: Spigot joint model by Enright et al. (2000) ................................................104
Figure 6.6: Schematic of spigot joint model (adapted from Enright et al. 2000) .........106
Figure 6.7: Tri-linear moment-rotation for the Cuplok joints.......................................108
Figure 6.8: Bending axes of the Cuplok joints..............................................................109
Figure 6.9: (a) base plate on uneven ground and (b) base eccentricity model..............110
Figure 6.10: P- and P- effects ...................................................................................112
Figure 6.11: Stress-strain curve for standard ................................................................114
Figure 6.12: Stress-strain curve for ledger....................................................................114
Figure 6.13: Stress-strain curve for jack .......................................................................115
Figure 6.14: Stress-strain curve for base plate ..............................................................115
List of Figures viii
______________________________________________________________________

Figure 6.15: Stress-strain curve for brace .....................................................................116


Figure 6.16: Stress-strain curve for spigot ....................................................................116
Figure 6.17: Finite element model of Test No. 3 showing axes ...................................119
Figure 6.18: Test capacity/Prediction of fifteen support scaffold systems ...................120
Figure 6.19: Calibration of load-deflection responses at mid-height of the standard in
the 3rd lift of the 2nd row of the frame for Test No. 2....................................................120
Figure 6.20: Calibration of load-deflection responses at mid-height of the standard in
the 3rd lift of the 2nd row of the frame for Test No. 3....................................................121
Figure 6.21: Calibration of load-deflection responses at mid-height of the standard in
the 3rd lift of the 2nd row of the frame for Test No. 4....................................................121
Figure 6.22: Calibration of load-deflection responses at mid-height of the standard in
the 3rd lift of the 2nd row of the frame for Test No. 5....................................................122
Figure 6.23: Calibration of load-deflection responses at mid-height of the standard in
the 3rd lift of the 2nd row of the frame for Test No. 6....................................................122
Figure 6.24: Calibration of load-deflection responses at mid-height of the standard in
the 3rd lift of the 2nd row of the frame for Test No. 8....................................................123
Figure 6.25: Calibration of load-deflection responses at mid-height of the standard in
the 3rd lift of the 2nd row of the frame for Test No. 9....................................................123
Figure 6.26: Calibration of load-deflection responses at mid-height of the standard of
the 3rd lift of the 2nd row of the frame for Test No. 10..................................................124
Figure 6.27: Calibration of load-deflection responses at mid-height of the standard in
the 3rd lift of the 2nd row of the frame for Test No. 11..................................................124
Figure 6.28: Calibration of load-deflection responses at mid-height of the standard in
the 3rd lift of the 2nd row of the frame for Test No. 12..................................................125
Figure 6.29: Calibration of load-deflection responses at the 2nd lift of the 1st row of the
frame for Test No. 13 ....................................................................................................125
Figure 6.30: Calibration of load-deflection responses at mid-height of the standard in
the 3rd lift of the 2nd row of the frame for Test No. 14..................................................126
Figure 6.31: Calibration of load-deflection responses at mid-height of the standard in
the 2nd lift of the 2nd row of the frame for Test No. 15 .................................................126
Figure 6.32: Calibration of load-deflection responses at mid-height of the standard in
the 3rd lift of the 2nd row of the frame for Test No. 16..................................................127
Figure 6.33: Calibration of load-deflection responses at mid-height of the standard in
the 3rd lift of the 2nd row of the frame for Test No. 18..................................................127
Figure 6.34: S-shape member buckling ........................................................................129
Figure 6.35: Lateral frame buckling..............................................................................129
Figure 7.1: Typical finite element model used in the parametric studies .....................133
Figure 7.2: Parametric studies of loading eccentricity of 1.0 m lift scaffold systems ..138
Figure 7.3: Parametric studies of loading eccentricity of 1.5 m lift scaffold systems ..138
Figure 7.4: Parametric studies of loading eccentricity of 2.0 m lift scaffold systems ..139
Figure 7.5: Parametric studies of number of lifts of scaffold systems..........................140
Figure 7.6: Typical failure mode of scaffold systems with 9 lifts ................................140
Figure 7.7: Bracing configurations in the parametric studies .......................................141
Figure 7.8: Parametric studies of out-of-straightness of 1.0 m lift scaffold systems....145
Figure 7.9: Parametric studies of out-of-straightness of 1.5 m lift scaffold systems....145
Figure 7.10: Parametric studies of out-of-straightness of 2.0 m lift scaffold systems..146
Figure 8.1: 1x1 bay scaffold configuration in the studies.............................................154
Figure 8.2: 3x3 bays scaffold configuration in the studies ...........................................154
Figure 8.3: 3x6 bays scaffold configuration in the studies ...........................................155
List of Figures ix
______________________________________________________________________

Figure 8.4: 9x9 bays scaffold configuration in the studies ...........................................156


Figure 8.5: Typical finite element model for Monte Carlo simulations .......................157
Figure 8.6: Graph of (a) typical Cuplok joint behaviour (b) tri-linear joint model ......160
Figure 8.7: Histogram for the ultimate strength of 1x1 bay, 1 m lift height, 3 lifts, and
100 mm top and bottom jack extensions support scaffold frame .................................165
Figure 8.8: Histogram for the ultimate strength of 1x1 bay, 1 m lift height, 3 lifts, and
300 mm top and bottom jack extensions support scaffold frame .................................165
Figure 8.9: Histogram for the ultimate strength of 1x1 bay, 1 m lift height, 3 lifts, and
600 mm top and bottom jack extensions support scaffold frame .................................166
Figure 8.10: Histogram for the ultimate strength of 3x3 bays, 1 m lift height, 3 lifts, and
100 mm top and bottom jack extensions support scaffold frame .................................166
Figure 8.11: Histogram for the ultimate strength of 3x3 bays, 1 m lift height, 3 lifts, and
300 mm top and bottom jack extensions support scaffold frame .................................167
Figure 8.12: Histogram for the ultimate strength of 3x3 bays, 1 m lift height, 3 lifts, and
600 mm top and bottom jack extensions support scaffold frame .................................167
Figure 8.13: Histogram for the ultimate strength of 3x6 bays, 1 m lift height, 3 lifts, and
100 mm top and bottom jack extensions support scaffold frame .................................168
Figure 8.14: Histogram for the ultimate strength of 3x6 bays, 1 m lift height, 3 lifts, and
300 mm top and bottom jack extensions support scaffold frame .................................168
Figure 8.15: Histogram for the ultimate strength of 3x6 bays, 1 m lift height, 3 lifts, and
600 mm top and bottom jack extensions support scaffold frame .................................169
Figure 8.16: Histogram for the ultimate strength of 9x9 bays, 1 m lift height, 3 lifts, and
100 mm top and bottom jack extensions support scaffold frame .................................169
Figure 8.17: Histogram for the ultimate strength of 9x9 bays, 1 m lift height, 3 lifts, and
300 mm top and bottom jack extensions support scaffold frame .................................170
Figure 8.18: Histogram for the ultimate strength of 9x9 bays, 1 m lift height, 3 lifts, and
600 mm top and bottom jack extensions support scaffold frame .................................170
Figure 8.19: Histogram for the ultimate strength of 1x1 bay, 1.5 m lift height, 3 lifts,
and 100 mm top and bottom jack extensions support scaffold frame...........................171
Figure 8.20: Histogram for the ultimate strength of 1x1 bay, 1.5 m lift height, 3 lifts,
and 300 mm top and bottom jack extensions support scaffold frame...........................171
Figure 8.21: Histogram for the ultimate strength of 1x1 bay, 1.5 m lift height, 3 lifts,
and 600 mm top and bottom jack extensions support scaffold frame...........................172
Figure 8.22: Histogram for the ultimate strength of 3x3 bays, 1.5 m lift height, 3 lifts,
and 100 mm top and bottom jack extensions support scaffold frame...........................172
Figure 8.23: Histogram for the ultimate strength of 3x3 bays, 1.5 m lift height, 3 lifts,
and 300 mm top and bottom jack extensions support scaffold frame...........................173
Figure 8.24: Histogram for the ultimate strength of 3x3 bays, 1.5 m lift height, 3 lifts,
and 600 mm top and bottom jack extensions support scaffold frame...........................173
Figure 8.25: Histogram for the ultimate strength of 3x6 bays, 1.5 m lift height, 3 lifts,
and 100 mm top and bottom jack extensions support scaffold frame...........................174
Figure 8.26: Histogram for the ultimate strength of 3x6 bays, 1.5 m lift height, 3 lifts,
and 300 mm top and bottom jack extensions support scaffold frame...........................174
Figure 8.27: Histogram for the ultimate strength of 3x6 bays, 1.5 m lift height, 3 lifts,
and 600 mm top and bottom jack extensions support scaffold frame...........................175
Figure 8.28: Histogram for the ultimate strength of 9x9 bays, 1.5 m lift height, 3 lifts,
and 100 mm top and bottom jack extensions support scaffold frame...........................175
Figure 8.29: Histogram for the ultimate strength of 9x9 bays, 1.5 m lift height, 3 lifts,
and 300 mm top and bottom jack extensions support scaffold frame...........................176
List of Figures x
______________________________________________________________________

Figure 8.30: Histogram for the ultimate strength of 9x9 bays, 1.5 m lift height, 3 lifts,
and 600 mm top and bottom jack extensions support scaffold frame...........................176
Figure 8.31: Histogram for the ultimate strength of 1x1 bay, 2 m lift height, 3 lifts, and
100 mm top and bottom jack extensions support scaffold frame .................................177
Figure 8.32: Histogram for the ultimate strength of 1x1 bay, 2 m lift height, 3 lifts, and
300 mm top and bottom jack extensions support scaffold frame .................................177
Figure 8.33: Histogram for the ultimate strength of 1x1 bay, 2 m lift height, 3 lifts, and
600 mm top and bottom jack extensions support scaffold frame .................................178
Figure 8.34: Histogram for the ultimate strength of 3x3 bays, 2 m lift height, 3 lifts, and
100 mm top and bottom jack extensions support scaffold frame .................................178
Figure 8.35: Histogram for the ultimate strength of 3x3 bays, 2 m lift height, 3 lifts, and
300 mm top and bottom jack extensions support scaffold frame .................................179
Figure 8.36: Histogram for the ultimate strength of 3x3 bays, 2 m lift height, 3 lifts, and
600 mm top and bottom jack extensions support scaffold frame .................................179
Figure 8.37: Histogram for the ultimate strength of 3x6 bays, 2 m lift height, 3 lifts, and
100 mm top and bottom jack extensions support scaffold frame .................................180
Figure 8.38: Histogram for the ultimate strength of 3x6 bays, 2 m lift height, 3 lifts, and
300 mm top and bottom jack extensions support scaffold frame .................................180
Figure 8.39: Histogram for the ultimate strength of 3x6 bays, 2 m lift height, 3 lifts, and
600 mm top and bottom jack extensions support scaffold frame .................................181
Figure 8.40: Histogram for the ultimate strength of 9x9 bays, 2 m lift height, 3 lifts, and
100 mm top and bottom jack extensions support scaffold frame .................................181
Figure 8.41: Histogram for the ultimate strength of 9x9 bays, 2 m lift height, 3 lifts, and
300 mm top and bottom jack extensions support scaffold frame .................................182
Figure 8.42: Histogram for the ultimate strength of 9x9 bays, 2 m lift height, 3 lifts, and
600 mm top and bottom jack extensions support scaffold frame .................................182
Figure 8.43: Typical failure mode of 1.0 m lift system with 100 mm jack extensions.185
Figure 8.44: Typical failure mode of 2.0 m lift system with 600 mm jack extensions.185
Figure 8.45: Elastic-perfectly plastic model for steel ...................................................188
Figure 8.46: Comparison between Australian and ASCE design formulae based on
Ln/Dn ..............................................................................................................................197
Figure 8.47: System resistance factor for 1.0 m lift scaffold system............................197
Figure 8.48: System resistance factor for 1.5 m lift scaffold system............................198
Figure 8.49: System resistance factor for 2.0 m lift scaffold system............................198
Figure 9.1: Idealised nominal design model (elevation view) ......................................202
Figure 9.2: Typical three-dimensional finite element (advanced) analysis model .......203
Figure 9.3: Illustration of a spigot model in Strand7 ....................................................204
Figure 9.4: Illustration of a loading eccentricity in Strand7 .........................................205
Figure 9.5: Illustration of a base plate model in Strand7 ..............................................206
Figure 9.6: Idealised tri-linear joint model ...................................................................207
Figure 9.7: Elastic-perfectly plastic material model for nominal design ......................209
Figure 9.8: Advanced analysis model for case study 1.................................................216
Figure 9.9: Load-deflection response at the mid-height of the 2nd lift for case study 1217
Figure 9.10: Cross section of the box girder for case study 2.......................................218
Figure 9.11: Advanced analysis model for case study 2...............................................220
Figure 9.12: Load-deflection response at the mid-height of the 6th lift for case study 2
.......................................................................................................................................221
Figure B.1.1: Initial Cuplok joint stiffness for KzA1 based on the number of hammer
blows applied to tighten the cup ...................................................................................243
List of Figures xi
______________________________________________________________________

Figure B.1.2: Initial Cuplok joint stiffness for KzB1 based on the number of hammer
blows applied to tighten the cup ...................................................................................244
Figure B.1.3: Initial Cuplok joint stiffness for KzC1 based on the number of hammer
blows applied to tighten the cup ...................................................................................244
Figure B.1.4: Initial Cuplok joint stiffness for KzD1 based on the number of hammer
blows applied to tighten the cup ...................................................................................245
Figure B.1.5: Initial Cuplok joint stiffness for KzA2 based on the number of hammer
blows applied to tighten the cup ...................................................................................245
Figure B.1.6: Initial Cuplok joint stiffness for KzB2 based on the number of hammer
blows applied to tighten the cup ...................................................................................246
Figure B.1.7: Initial Cuplok joint stiffness for KzC2 based on the number of hammer
blows applied to tighten the cup ...................................................................................246
Figure B.1.8: Initial Cuplok joint stiffness for KzD2 based on the number of hammer
blows applied to tighten the cup ...................................................................................247
Figure B.1.9: Initial Cuplok joint stiffness for KyA1 based on the number of hammer
blows applied to tighten the cup ...................................................................................247
Figure B.1.10: Initial Cuplok joint stiffness for KyB1 based on the number of hammer
blows applied to tighten the cup ...................................................................................248
Figure B.1.11: Initial Cuplok joint stiffness for KyC1 based on the number of hammer
blows applied to tighten the cup ...................................................................................248
Figure B.1.12: Initial Cuplok joint stiffness for KyD1 based on the number of hammer
blows applied to tighten the cup ...................................................................................249
Figure B.1.13: Initial Cuplok joint stiffness for KyC2 based on the number of hammer
blows applied to tighten the cup ...................................................................................249
Figure B.1.14: Initial Cuplok joint stiffness for KyD2 based on the number of hammer
blows applied to tighten the cup ...................................................................................250
Figure B.1.15: Initial Cuplok joint stiffness for KzA1 based on the type of finish
(galvanised or painted)..................................................................................................250
Figure B.1.16: Initial Cuplok joint stiffness for KzB1 based on the type of finish
(galvanised or painted)..................................................................................................251
Figure B.1.17: Initial Cuplok joint stiffness for KzC1 based on the type of finish
(galvanised or painted)..................................................................................................251
Figure B.1.18: Initial Cuplok joint stiffness for KzD1 based on the type of finish
(galvanised or painted)..................................................................................................252
Figure B.1.19: Initial Cuplok joint stiffness for KzA2 based on the type of finish
(galvanised or painted)..................................................................................................252
Figure B.1.20: Initial Cuplok joint stiffness for KzB2 based on the type of finish
(galvanised or painted)..................................................................................................253
Figure B.1.21: Initial Cuplok joint stiffness for KzC2 based on the type of finish
(galvanised or painted)..................................................................................................253
Figure B.1.22: Initial Cuplok joint stiffness for KzD2 based on the type of finish
(galvanised or painted)..................................................................................................254
Figure B.1.23: Initial Cuplok joint stiffness for KyA1 based on the type of finish
(galvanised or painted)..................................................................................................254
Figure B.1.24: Initial Cuplok joint stiffness for KyB1 based on the type of finish
(galvanised or painted)..................................................................................................255
Figure B.1.25: Initial Cuplok joint stiffness for KyC1 based on the type of finish
(galvanised or painted)..................................................................................................255
List of Figures xii
______________________________________________________________________

Figure B.1.26: Initial Cuplok joint stiffness for KyD1 based on the type of finish
(galvanised or painted)..................................................................................................256
Figure B.1.27: Initial Cuplok joint stiffness for KyC2 based on the type of finish
(galvanised or painted)..................................................................................................256
Figure B.1.28: Initial Cuplok joint stiffness for KyD2 based on the type of finish
(galvanised or painted)..................................................................................................257
Figure B.2.1: Fitted normal distribution of normalised Cuplok joint stiffness k1
(looseness alone) in bending about the horizontal axis.................................................258
Figure B.2.2: Fitted normal distribution of normalised Cuplok joint stiffness k2 in
bending about the horizontal axis .................................................................................259
Figure B.2.3: Fitted normal distribution of normalised Cuplok joint stiffness k3 in
bending about the horizontal axis .................................................................................259
Figure C.1.1: Gridlines used for imperfection readings................................................260
Figure C.1.2: Measuring points along standards for imperfection readings .................260
List of Tables xiii
______________________________________________________________________

LIST OF TABLES

Table 2.1: Recommended target reliability index in literature .......................................46


Table 3.1: Fitted probability distributions for the normalised out-of-straightness, the
normalised out-of-plumb and the loading eccentricity of the support scaffold systems.58
Table 4.1: Labels for test configurations.........................................................................69
Table 4.2: Summary of initial (elastic) Cuplok joint stiffness........................................80
Table 4.3: Statistical results for initial (elastic) Cuplok joint stiffness ...........................81
Table 4.4: Mean Cuplok joint stiffness (kNm/rad) .........................................................84
Table 4.5: Mean rotation for Cuplok joints (rad)............................................................84
Table 4.6: Coefficient of variation of Cuplok joint stiffness ..........................................84
Table 5.1: Summary of test configurations .....................................................................90
Table 5.2: Summary of test results..................................................................................94
Table 6.1: Average Cuplok joint stiffness (kNm/rad)...................................................108
Table 6.2: Average rotation for Cuplok joints (rad) .....................................................108
Table 6.3: Ramberg-Osgood parameters for scaffold components...............................113
Table 6.4: Parametric calibration results.......................................................................118
Table 6.5: Load calibration results................................................................................119
Table 7.1: Mean values for parameters in the parametric studies (refer to Chapters 3-5)
.......................................................................................................................................134
Table 7.2: Ultimate system strength for reference cases (capacity per upright)...........135
Table 7.3: Parametric studies of bracing configurations of scaffold systems...............142
Table 7.4: Parametric studies of Cuplok joint stiffness of scaffold systems ................143
Table 7.5: Parametric studies of member out-of-straightness of scaffold systems.......146
Table 7.6: Parametric studies of member yield stress of scaffold systems...................148
Table 7.7: Parametric studies of member cross-section of scaffold systems................149
Table 8.1: Mean Cuplok joint stiffness (kNm/rad) .......................................................160
Table 8.2: Mean rotation for Cuplok joints (rad)..........................................................161
Table 8.3: Coefficient of variation of Cuplok joint stiffness ........................................161
Table 8.4: Statistical data for initial geometric imperfections used in the simulations 162
Table 8.5: Statistical data for loading eccentricity used in the simulations ..................164
Table 8.6: Statistical summary of simulation results for system ultimate strengths.....183
Table 8.7: Input parameters for nominal design models...............................................186
Table 8.8: Nominal ultimate strength for nominal design models................................187
Table 8.9: Statistics for the system resistance of support scaffold systems..................190
Table 8.10: Statistics for construction loads .................................................................196
Table 8.11: System resistance factors for support scaffold systems ( = 3.5)..............199
Table 9.1: Cuplok joint stiffness (kNm/rad) for nominal design model .......................207
Table 9.2: Rotation for Cuplok joints (rad) for nominal design model ........................207
Table 9.3: Nominal yield stress of steel scaffold components in this research.............209
Table 9.4: Summary of system resistance factors for support scaffold systems ( = 3.5)
.......................................................................................................................................212
Table 9.5: Design comparisons between advanced analysis and member-based design to
AS 4100 (Boral design charts) ......................................................................................214
Table A.1: Data for out-of-straightness of the standards ..............................................233
Table A.2: Data for storey out-of-plumb ......................................................................237
List of Tables xiv
______________________________________________________________________

Table A.3: Data for loading eccentricity.......................................................................241


Table C.1.1: Initial geometric imperfections (mm) for subassembly Test No. 2..........261
Table C.1.2: Initial geometric imperfections (mm) for subassembly Test No. 3..........263
Table C.1.3: Initial geometric imperfections (mm) for subassembly Test No. 4..........265
Table C.1.4: Initial geometric imperfections (mm) for subassembly Test No. 5..........267
Table C.1.5: Initial geometric imperfections (mm) for subassembly Test No. 6..........269
Table C.1.6: Initial geometric imperfections (mm) for subassembly Test No. 8..........271
Table C.1.7: Initial geometric imperfections (mm) for subassembly Test No. 9..........273
Table C.1.8: Initial geometric imperfections (mm) for subassembly Test No. 10........275
Table C.1.9: Initial geometric imperfections (mm) for subassembly Test No. 11........277
Table C.1.10: Initial geometric imperfections (mm) for subassembly Test No. 12......279
Table C.1.11: Initial geometric imperfections (mm) for subassembly Test No. 13......281
Table C.1.12: Initial geometric imperfections (mm) for subassembly Test No. 14......283
Table C.1.13: Initial geometric imperfections (mm) for subassembly Test No. 15......285
Table C.1.14: Initial geometric imperfections (mm) for subassembly Test No. 16......287
Table C.1.15: Initial geometric imperfections (mm) for subassembly Test No. 18......289
Abbreviations xv
______________________________________________________________________

ABBREVIATIONS

API Application programming interface


ASD Allowable stress design
CHS Circular hollow sections
COV Coefficient of variation
EBM Scaling of eigenbuckling modes
FORM First-order reliability method
FOSM First-order second-moment method
IGI Direct modelling of initial geometric imperfections
LRFD Load and resistance factor design
LVDT Linear variable displacement transducer
NHF Application of notional horizontal forces
STD Standard deviation
Principal Notations xvi
______________________________________________________________________

PRINCIPAL NOTATIONS

a Robertson constant
A cross-sectional area
Ae effective area
Ag gross area
B number of storeys
d distance between objects
de external diameter of the tube
di internal diameter of the tube
do outside diameter
d* initial design point for dead load

D mean dead load


Dn nominal dead load
E Young's modulus
E0 initial Young's modulus
fQ(x) probability density function of Q
fy yield strength
FR(x) cumulative distribution function of R
G limit state function
Gpartial partial derivative of the limit state function
h height of the scaffold unit
he effective height of the scaffold
H height of the scaffold system
I moment of inertia
k effective length factor
ke member effective length factor
kf form factor
K rotational stiffness; effective length coefficient
Principal Notations xvii
______________________________________________________________________

l* initial design point for live load


ls total length of spigot insert
L one-storey height of the scaffold unit; member length

L mean live load


Lh lift height
Ln nominal live load
M moment
n compression member imperfection factor; Ramberg-Osgood parameter
N number of storeys
Nc nominal member capacity
pc compressive strength of the column
pE elastic buckling strength of the column
py yield strength of the steel tube
P vertical load
Pcr critical load
Pf probability of failure
Q total load effect
Qi nominal design loads
r radius of gyration
r* initial design point for system resistance
R structural resistance; structural system resistance

R mean strength; mean structural system resistance


Rn nominal strength; nominal system resistance
Rpredict system resistance from advanced analysis
Rtest system resistance from experimental test
S ratio of the height of shore extension to the height of the scaffold unit
t wall thickness
VD coefficient of variation of dead load
VL coefficient of variation of live load
VM coefficient of variation from modelling uncertainty
Principal Notations xviii
______________________________________________________________________

VR coefficient of variation of system resistance


VS coefficient of variation from specimen
Vtest/predict coefficient of variation from Rtest/Rpredict
VT coefficient of variation from testing
xi design point
Xi random variable
zi reduced variate
sensitivity factor
a compression member factor
b compression member constant
c slenderness reduction factor
reliability index
i load factors
deflection
horizontal sway; displacement
t top horizontal sway under a unit load from linear analysis
Perry factor
rotation
slenderness ratio; elastic buckling load factor
0 limiting slenderness
a load amplifier
e slenderness of a flat plate element
n modified compression member slenderness
D bias factor for dead load
L bias factor for live load
R bias factor for system resistance
statistical mean
D mean value of the dead load
G mean value of the limit state function
L mean value of the live load
Principal Notations xix
______________________________________________________________________

R mean value of the system resistance


modified compression member factor
standard deviation
G standard deviation of the limit state function
0.2 0.2% proof stress
member capacity reduction factor
system system resistance factor
Chapter 1 - Introduction 1
______________________________________________________________________

CHAPTER 1 - INTRODUCTION

1.1 Problem Statement

Support scaffold systems provide temporary support for the construction of bridges,
commercial and residential buildings, car parks and similar structures. Support scaffold
systems (Figure 1.1) are used as platforms while building the timber formwork for
reinforced concrete, and are usually loaded heavily under the weight of wet concrete
and construction equipment. They are removed and reused once the concrete sets and
gains strength. Components of support scaffold systems are usually made from steel
circular hollow sections in the form of standards (uprights), ledgers (beams), and braces,
except jacks that are made of threaded steel rods and base plates that are made of steel
sheets. From substantial use and reuse of these components, geometric imperfections
are deemed to occur and influence the strength of scaffold systems. The magnitude of
loading eccentricity that can occur between the timber bearer and the U-head (Figure
1.2) is also a crucial factor affecting the strength of scaffold systems. In order to resolve
the issue of imperfections and loading eccentricity in scaffold systems, actual data is
needed. At present, no available data is published for geometric imperfections and
loading eccentricity of scaffold systems. This thesis presents the methods and results in
the procurement of geometric imperfection data including measurements of the loading
eccentricity between the timber bearer and the U-head for Cuplok scaffold systems.

The joints between standard and ledger present another challenging task in modelling of
scaffold systems. In the past, the joints have been modelled as simple flexible joints,
unable to transfer moments, as assumed by Harung et al. (1975) and Milojkovic et al.
(1996). In actuality, the joints are semi-rigid as noticed by Godley and Beale (2001),
and good agreement between test and analysis is achievable when the actual joint
characteristics are modelled, as shown by Prabhakaran et al. (2006). Nevertheless,
information on semi-rigid scaffold joint characteristics is limited and different joint
types are available in the market such as wedge-type and Cuplok joints. This thesis
Chapter 1 - Introduction 2
______________________________________________________________________

presents the tests on Cuplok joints and proposes a tri-linear joint model in terms of
moment-rotation curves for different joint configurations.

Figure 1.1: Typical support scaffold systems

Figure 1.2: Timber bearer supported by the U-head screw jack


Chapter 1 - Introduction 3
______________________________________________________________________

The international research community is currently focused on the stability behaviour of


scaffold systems, concentrating on advanced analysis models for large frame systems,
including behavioural and design oriented research. Early structural analyses of scaffold
systems employed simple linear buckling analyses, as shown by Harung et al. (1975)
and Lightfoot and Oliveto (1977), while more recently, geometric second-order
structural analyses have been applied to steel scaffold systems, such as those performed
by Weesner and Jones (2001), Chan et al. (2002), Vaux et al. (2002) and Yu and Chung
(2004). This thesis presents advanced analysis models of support scaffold systems that
take into account of the effects of geometric and material nonlinearity as well as
nonlinear joint stiffness.

The design by advanced analysis is relatively new among professional engineers. The
Australian steel design code (Standard Australia 1998), which is in the ultimate limit
states, or load and resistance factor design (LRFD) format, permits the use of design by
advanced analysis for steel frames, comprising of members of compact section with full
lateral restraint. Nevertheless, in the absence of research to determine system resistance
factors (system capacity reduction factors) for various structural systems, the design by
advanced analysis usually relies on engineering judgement as to what resistance factors
should be applied. Some researchers have studied the system strength predictions of
steel frames by advanced analysis, involving uncertainties in loads and material yield
strength (Buonopane and Schafer 2006). By nonlinear structural simulations and
reliability analysis, system resistance factors based on certain target reliability for the
design of steel frames by advanced analysis can be determined. The same concept is
applied in this thesis for the design by advanced analysis of support scaffold systems.
The strength of support scaffold systems is affected by variations in geometric and
material parameters, particularly geometric imperfections, joint stiffness, loading
eccentricity, and yield stress. Thus, this thesis presents the effects of these uncertainties
on the ultimate strength of multi-storey steel support scaffold systems through a rational
statistical framework and a second-order inelastic finite element (advanced) analysis.
Finally, the thesis presents a reliability analysis of support scaffold systems to determine
system resistance factors, leading to the formulation of a design method for support
scaffold systems based on advanced analysis, which requires no recourse to a steel
design specification for checking member and connection strengths.
Chapter 1 - Introduction 4
______________________________________________________________________

1.2 Scope and Objectives

The scope and objectives of this research are summarised as follows:

(a) Development of an accurate finite element (advanced) analysis model to capture


the behaviour of support scaffold systems, which includes spigot joints,
nonlinear moment-rotation joint stiffness for the standard-to-ledger connections,
brace connections, base plate eccentricity, loading eccentricity, initial geometric
imperfections, and nonlinear material model.

(b) Procurement of initial geometric imperfection and loading eccentricity data of


support scaffold systems from various construction site field measurements and
probabilistic studies of such data.

(c) Experimental tests of Cuplok joint to study the behaviour of semi-rigid joints of
support scaffold systems and determine the joint stiffness for rotations about
vertical and horizontal axes in various joint configurations, as well as
probabilistic studies of the joint stiffness.

(d) Parametric studies to determine the factors affecting the strength of support
scaffold systems.

(e) Probabilistic assessment of support scaffold systems using Monte Carlo


simulations to determine the statistics of system strength, and reliability analysis
to compute system resistance factors to be used in the design by advanced
analysis according to the LRFD statistical framework.

(f) Development of guidelines for the design of support scaffold systems by


advanced analysis.
Chapter 1 - Introduction 5
______________________________________________________________________

1.3 Outline of Thesis

The thesis develops methods of advanced analysis and design of support scaffold
systems. Accurate finite element models for support scaffold systems are presented and
calibrated against the ultimate loads and deflection responses of full-scale subassembly
tests. Upright-to-beam connections, known as Cuplok joints, were experimentally tested
to determine joint stiffness and joint behaviour in various configurations so that
accurate joint modelling can be established. A survey on geometric imperfections
(storey out-of-plumb and out-of-straightness of uprights) and loading eccentricity of
support scaffold systems was carried out in construction sites around the Sydney
metropolitan area. Parametric studies were carried out to determine how various
parameters affect the strength of support scaffold systems. Probabilistic studies of joint
stiffness, geometric imperfections, and loading eccentricity were performed to
determine their distribution functions. Subsequently, the assessment of the strength of
support scaffold systems using Monte Carlo simulations of the influencing parameters
was achieved. A reliability analysis based on the LRFD framework was used to
determine resistance factors for support scaffold systems. Step-by-step advanced
analysis design of support scaffold systems was proposed and recommendations for the
design of support scaffold were given. A brief overview of the material contained in this
thesis is described in the following summary.

Chapter 2 presents an overview of scaffold research and current practice in the design of
scaffold systems. It covers a brief description of scaffold systems including the types of
joints and materials currently used. Also, types of analysis, loads, initial geometric
imperfections, and the modelling of complex joints are described. The prediction of the
ultimate loads of scaffold systems derived from simplified equations and their failure
modes are shown. In addition, the chapter explains the design of scaffold systems based
on the British and Australian standards as well as how effective lengths and bracings
commonly apply. The chapter also covers design by advanced analysis combined with a
reliability analysis as a tool to achieve a new innovative design methodology for
scaffold systems. Recommendations are provided for the modelling, analysis and design
of scaffold systems.
Chapter 1 - Introduction 6
______________________________________________________________________

Chapter 3 describes the findings from various site measurements of geometric


imperfections on support scaffold systems, also known as falsework in industry. The
measurements consist of out-of-straightness of the standards (uprights), out-of-plumb of
the frame and loading eccentricity between the timber bearer and the U-head. A special-
made tool instrumented with a dial gauge was used to measure the out-of-straightness of
standards at the mid-height of each lift. A theodolite was employed to measure the angle
difference between top and bottom of the frame in order to compute storey out-of-
plumb, and a vernier calliper was used to measure the loading eccentricity at the top.
The measurements were taken from different support scaffold construction sites before
the pouring of concrete, thus representing actual initial imperfections and loading
eccentricity encountered in practice. The statistical analysis of the data is presented for
practical application in modelling.

Chapter 4 describes the setup, procedure, and results of scaffold Cuplok joint tests. The
aims of the tests are to investigate the joint stiffness for rotations about vertical and
horizontal axes in various joint configurations, and carry out statistical analyses of the
experimental results. The tests were performed in the laboratory of the School of Civil
Engineering at the University of Sydney using second-hand Cuplok scaffold parts
provided by Boral Formwork and Scaffolding Pty. Ltd. A total of 172 tests were carried
out on various joint configurations, bending axes, loading directions, types of material
(galvanised or painted components), and degree of tightening of the joints. The results
were shown graphically in terms of moment-rotation curves. Since there was substantial
variation in Cuplok joint stiffness, a statistical analysis on the results was performed.
These experimental studies are especially useful for modelling and performing
probabilistic analysis on the Cuplok scaffold systems.

Chapter 5 summarises the details and results of 18 support scaffold subassembly tests,
conducted at the University of Sydney in 2006 by the Centre for Advanced Structural
Engineering to study the behaviour and ultimate load-carrying capacities of such
systems. The subassemblies consisted of a grid of 33 bays with a constant nominal bay
width of 1829 mm. The first 14 tests featured three lifts with equal nominal lift height of
1.5 m. In the last four tests, lift heights of 1 m and 2 m were used and the number of
Chapter 1 - Introduction 7
______________________________________________________________________

lifts varied between 2 and 4. All testing materials were sourced from stocks of material
currently in use. The testing materials therefore featured geometric imperfections
representative of those encountered in practice. The initial geometric imperfections are
tabulated for each test in Appendix C.1 for modelling purposes. The ultimate loads of
the systems and load-deflection curves are presented in Chapter 5 and Chapter 6,
respectively.

In Chapter 6, accurate three-dimensional advanced analysis models are developed to


capture the behaviour of support scaffold systems, as observed in full-scale
subassembly tests consisting of three-by-three bay scaffold systems with combinations
of various lift heights, number of lifts and jack extensions. The chapter proposes
methods for modelling spigot joints, semi-rigid upright-to-beam connections and base
plate eccentricities. Material nonlinearity is taken into account in the models based on
the Ramberg-Osgood expression fitted to available experimental data. Actual initial
geometric imperfections including member out-of-straightness and storey out-of-plumb
are also incorporated in the models. The ultimate loads from the nonlinear analyses are
calibrated against failure loads and load-deflection responses obtained from full-scale
subassembly tests. The numerical results show very good agreement with tests,
indicating that it is possible to accurately predict the behaviour and strength of highly
complex support scaffold systems using material and geometric nonlinear analysis.

Chapter 7 presents parametric studies of various factors influencing the ultimate load of
scaffold systems. These parameters include loading eccentricity, jack extensions, lift
heights, number of lifts, bracing arrangements, Cuplok joint stiffness, geometric
imperfections as well as yield stress and cross-sectional area of the main loaded
components. Investigation of the factors affecting the strength of support scaffold
systems helps to determine which factors should be included and treated as random
variables or deterministic values in the probabilistic assessment of the system strength.

Chapter 8 studies the effects of uncertainties on the ultimate strength of multi-storey


steel support scaffold systems through a rational statistical framework and a second-
order inelastic finite element (advanced) analysis. The chapter also presents a reliability
Chapter 1 - Introduction 8
______________________________________________________________________

analysis of support scaffold systems to determine system resistance factors that can be
used for the design by advanced analysis according to the LRFD framework.

Chapter 9 provides a structural design guideline for the LRFD-based advanced analysis
method for determining the strength of the support scaffold systems. The chapter gives
recommendations on how to create practical nonlinear finite element scaffold models to
use with the design by advanced analysis. The step-by-step structural analysis and
design procedures for support scaffold systems are presented. Some comparisons of the
ultimate strengths obtained by the design using advanced analysis to those obtained
from a design chart by Boral Formwork & Scaffolding Pty. Ltd. (2002) are presented.
Two design case studies are demonstrated to show the practicality of the design of
support scaffold systems by advanced analysis.

Chapter 10 draws conclusions on this research and makes recommendations for future
research studies.

1.4 List of Publications

During this research work, the following research reports, conference and journal papers,
have been published, accepted or submitted for publication.

Research reports:

Chandrangsu, T., and Rasmussen, K.J.R. (2008). Scaffold Cuplok joint tests. Research
Report No. R893. Centre for Advanced Structural Engineering, School of Civil
Engineering, University of Sydney.

Chandrangsu, T., and Rasmussen, K.J.R. (2009). Investigation of geometric


imperfections of support scaffold systems. Research Report No. R895. Centre for
Advanced Structural Engineering, School of Civil Engineering, University of Sydney.
Chapter 1 - Introduction 9
______________________________________________________________________

Chandrangsu, T., and Rasmussen, K.J.R. (2009). Structural modelling of support


scaffold systems. Research Report No. R896. Centre for Advanced Structural
Engineering, School of Civil Engineering, University of Sydney.

Chandrangsu, T., and Rasmussen, K.J.R. (2009). Review of past research on scaffold
systems. Research Report No. R905. Centre for Advanced Structural Engineering,
School of Civil Engineering, University of Sydney.

Conference papers:

Chandrangsu, T., and Rasmussen, K.J.R. (2009). Geometric imperfection


measurements and joint stiffness of support scaffold systems. 6th International
Conference on Advances in Steel Structures (ICASS09), Hong Kong, 1075-1082.

Chandrangsu, T., and Rasmussen, K.J.R. (2009). Full-scale tests and advanced
structural analysis of formwork subassemblies. 6th International Conference on
Advances in Steel Structures (ICASS09), Hong Kong, 1083-1090.

Chandrangsu, T., Rasmussen, K.J.R., and Zhang, H. (2010). Reliability-based


methodology for the design of support formwork systems by advanced structural
analysis. 4th International Conference on Structural Engineering, Mechanics and
Computation (SEMC10), Cape Town, South Africa (Accepted).

Zhang, H., Chandrangsu, T., and Rasmussen, K.J.R. (2009). Probabilistic study of the
strength of steel scaffold systems. 10th International Conference on Structural Safety
and Reliability (ICOSSAR09), Osaka, Japan, 1376-1382.
Chapter 1 - Introduction 10
______________________________________________________________________

Zhang, H., Chandrangsu, T., and Rasmussen, K.J.R. (2009). System reliability of steel
scaffold systems. 6th International Conference on Advances in Steel Structures
(ICASS09), Hong Kong, 1065-1074.

Journal papers:

Chandrangsu, T., and Rasmussen, K.J.R. (2010). Investigation of geometric


imperfections and joint stiffness of support scaffold systems. Journal of Constructional
Steel Research (Under review).

Chandrangsu, T., and Rasmussen, K.J.R. (2010). Structural modelling of support


scaffold systems. Journal of Constructional Steel Research (Under review).

Chandrangsu, T., Rasmussen, K.J.R., and Zhang, H. (2010). Design of support scaffold
systems by advanced analysis. ASCE Journal of Structural Engineering (Ready for
submission).

Zhang, H., Chandrangsu, T., and Rasmussen, K.J.R. (2010). Probabilistic study of the
strength of steel scaffold systems. Structural Safety, In Press, Corrected Proof.
Chapter 2 - Literature Review 11
______________________________________________________________________

CHAPTER 2 - LITERATURE REVIEW

2.1 Introduction

This chapter presents an overview of scaffold research and current practice in the design
of scaffold systems. It covers brief description of scaffold systems including types of
joints and materials currently used. Also, types of analysis, loads, initial geometric
imperfections, and modelling of complex joints are described. In terms of modelling, it
focuses on how complex joints and boundary conditions have been modelled and how
geometric imperfections have been taken into account. The prediction of the ultimate
load of scaffold systems derived from simplified equations and their failure modes are
shown. In addition, the chapter explains the design of scaffold systems based on British
and Australian standards as well as how effective lengths and bracings commonly apply.
The chapter also covers design by advanced analysis with reliability analysis as a tool to
achieve a new innovative design for scaffold systems. The general recommendations are
provided for modelling, analysis and design of scaffold systems.

2.2 Scaffold Systems

Scaffolds are temporary structures generally used in construction to support various


types of loads. The vertical loads on scaffold can be from labourers, construction
equipment, formworks, and construction materials. Commonly, scaffolds must also be
designed to withstand lateral loads, including wind loads, impact loads, and earthquake
loads. Depending on the use of the scaffolds, they may be categorised as the access
scaffolds or the support scaffolds. The access scaffolds are used to support light to
moderate loads from labourers, small construction material and equipment for safe
working space. They are usually attached to buildings with ties and only one bay wide.
Support scaffolds, or sometimes called falsework, are subjected to heavy loads, for
example, concrete weight in the formwork. Both types of scaffolds can be seen in
everyday construction as shown in Figure 2.1.
Chapter 2 - Literature Review 12
______________________________________________________________________

(a) (b)

Figure 2.1: Typical scaffold systems: (a) access scaffold, and (b) support scaffold

2.2.1 Scaffold Configurations

Scaffolds are generally made up of slender framework. The configurations of scaffold


units vary from one manufacturer to another, as shown in Figure 2.2; however, they
share common features. Scaffolds normally consist of standards (vertical members),
ledgers (horizontal members), and braces, as illustrated in Figure 2.3.

(a) (b) (c)

(d) (e) (f)

Figure 2.2: Various types of scaffold unit: (a) simple (knee-braced) door type; (b)-(e)
standard door type; (f) stick construction with Cuplok joints or wedge-type joints
Chapter 2 - Literature Review 13
______________________________________________________________________

U-Head Screw Jack

Ledger

Brace

Standard

Jack Base

Figure 2.3: Configuration of typical stick-construction scaffold frame

The scaffold standards (vertical members) are connected to create a lift via couplers,
also known as spigot joints (Figure 2.4), and to connect ledgers (horizontal members) to
standards, Cuplok or wedge-type joints (Figure 2.5) are usually preferred because no
bolting or welding is required; though, in some systems manually adjusted pin-jointed
couplers are still being used. The connections for brace members are usually made of
hooks for easy assembling; however, in some systems pin-jointed couplers are used
(Figure 2.6). The base of scaffolds consists of adjustable jack bases (Figure 2.7), which
can be extended up to typically 600 mm by a wing nut to accommodate irregularity of
the ground. The access scaffolds usually have ties connecting them to a permanent
structure to increase the lateral stability of the system; in contrast, the support scaffolds
have adjustable shore extensions with U-head screw jacks (Figure 2.8) to support timber
bearers at the top to ensure the levelling of the formwork. Scaffold systems can be from
one storey (lift) up to many storeys, and can have many bays, and rows depending on
the type of construction. A scaffold unit is prefabricated to specific dimensions, and
assembled on site for the ease of construction. Moreover, scaffold members are reused
from one job to another, and for that reason, quality control program is required to
ensure that geometric imperfections, notably the crookedness of standards, remains
within stipulated tolerances.
Chapter 2 - Literature Review 14
______________________________________________________________________

Top Standard

Spigot Joint

Bottom Standard

Figure 2.4: Schematic of spigot joint

3
Locking Pin

2
Top Cup
Ledger Blade
1

Bottom Cup
Ledger
Standard

(a)

Wedge pin

Ledger

Clamp

Standard

(b)

Figure 2.5: Schematic of (a) Cuplok joint; and (b) wedge-type joint
Chapter 2 - Literature Review 15
______________________________________________________________________

Ledger

(a) Brace

Hook

Standard

Bolt Brace
(b)

Figure 2.6: Schematic of brace connections: (a) hook connection; and (b) pin connection

Standard

Ground
Wing Nut

Jack Base

Figure 2.7: Schematic of jack base


Chapter 2 - Literature Review 16
______________________________________________________________________

Timber Bearer

Adjustable U-Head

Wing Nut

Standard

Figure 2.8: Schematic of U-head jack

Normally, the total height of scaffold systems varies from 1.2 m up to 25 m. The height
of individual panel (lift) is usually between 1.0 m to 2.5 m, and the bay width ranges
from 0.7 m to 2.5 m. The plan configurations of the scaffold systems can be in different
shapes, for instance, rectangular shape, L shape, and U shape. Depending on
construction requirement, scaffold systems can be easily constructed to suit the needs
because of their flexibility in dimension and size.

2.2.2 Materials

Different natural materials such as timber and bamboo have been used in the past and
are still being used in Asia to construct scaffolds. In the western world, cold-formed
circular hollow steel sections are mainly used as members of scaffold system due to
their high strength and reusability. The steel tubes used for standards and ledgers
commonly have an outside diameter between 42 mm to 48 mm with thickness of
approximately 3 mm. As for bracing, various types of steel sections are currently used
in scaffold construction. Some braces are constructed with two periscopic tubes that can
slide inside one another to adjust the brace length. Following the trend of maximising
the efficiency in construction, aluminium is becoming increasingly utilised as members
in scaffold construction because of its lighter weight and ease of handling. Many
aluminium scaffold manufacturers are now located in China, Australia, New Zealand
and United Kingdom.
Chapter 2 - Literature Review 17
______________________________________________________________________

2.3 Collapses of the Scaffolds

2.3.1 Construction Stages

In 1985, Hadipriono and Wang (1987) compiled a report on the causes of failure of
worldwide support scaffold systems from 1961 until 1982. It was found that over 74%
of the collapses occur during concrete pouring operations due to the impact forces of
concrete pouring. In addition, some failures were reported to occur during the
disassembly of the formwork.

2.3.2 Method of Determining Causes of the Collapses

Hadipriono and Wang (1987) classified the failure occurrence into three groups,
representing the triggering causes, enabling causes, and procedural causes. The
triggering causes are external incidents that start scaffold collapses, for instance, heavy
loads on the scaffolds. The enabling causes are incidents that present insufficient design
and deficient construction. The procedural causes are linked with the triggering and
enabling causes, and are typically faults in communication among parties. Hadipriono
(1986) also introduced fuzzy set and fuzzy concept in measuring scaffold safety. His
method can be applied to determine the probability of event combinations that lead to
scaffold failure; therefore, it can be extended to control and minimise risks in scaffold
construction.

2.3.3 Main Causes of the Collapses

Hadipriono and Wang (1987) concluded that most triggering causes were due to
excessive loading on scaffolds, and impact load from concrete pouring was the major
concern for support scaffold systems. For enabling causes of failure, inadequate bracing
Chapter 2 - Literature Review 18
______________________________________________________________________

in scaffold systems was the main problem that led to the collapse of scaffolds during
construction. Inadequate review of scaffold design and absence of inspection during the
scaffold construction were the most important factors in procedural causes. Additionally,
Hadipriono and Wang reported some other significant causes of support scaffold
collapse such as improper or premature formwork removal, inadequate design, and
vibration from equipment.

In a study of high clearance scaffolds by Peng et al. (1996a), the possible causes of
support scaffold collapses were identified as overloading of the scaffold systems,
instability of shoring components, partial loading of fresh concrete in the formwork,
specific concrete placement pattern on the formwork, and load concentration from
concrete placement. The cause of the collapses due to load patterns notified by Peng et
al. (1996c) was presented in detail by the method of influence surfaces in a separate
study.

Milojkovic et al. (2002) presented an inspection by the HSE in the UK on typical faults
in access scaffold systems. The most common cause of collapse was insufficient tying
to a permanent structure. Some other structural faults included in the report were a
settlement of support, out-of-plumb and out-of-straightness of standards, and inadequate
bracing.

2.4 Analysis and Modelling of Scaffold Systems

2.4.1 Nonlinear Structural Analysis

With the ready availability of powerful computers and sophisticated structural analysis
software packages, nonlinear structural analysis has become feasible and practical.
Nonlinear analysis allows researchers and practitioners to more accurately predict the
failure load and deformation of scaffold systems. Nonlinear analysis involves the
modelling of changes of the geometry of structures as a result of loading and/or inelastic
material properties. In research by Gylltoft and Mroz (1995), and Chan et al. (2005), the
Chapter 2 - Literature Review 19
______________________________________________________________________

models were analysed considering both nonlinear material and geometric modelling.
However, in many cases research on scaffold systems focuses on nonlinear geometric
modelling associated with second-order effects since scaffold members are slender and
sensitive to stability effects. For example, elastic geometric nonlinear analyses were
reported by Peng et al. (2007), Prabhakaran et al. (2006), Yu et al. (2004), Chu et al.
(2002), and Weesner and Jones (2001). Geometric nonlinear analysis is also a common
practice in design offices, whereas the use of inelastic analysis is rare.

2.4.2 Three-Dimensional Model vs. Two-Dimensional Model

Access scaffold systems usually fail in complex three-dimensional modes locally or


globally, and require the use of three-dimensional analysis models to accurately predict
their behaviour and strength. Support scaffold systems are often more regular in
geometry, and can then be analysed and designed using two-dimensional models.
Particular attention needs to be paid to local eccentricity and member imperfections in
the nonlinear analysis of scaffold systems. By means of available commercial finite
element softwares such as ANSYS, GMNAF, and NIDA, many new studies on scaffold
behaviour were carried out through three-dimensional models such as those presented
by Prabhakaran et al. (2006), Milojkovic et al. (2002), and Godley and Beale (1997) .
Some past models proposed by Huang et al. (2000a) and Peng et al. (1997) were two-
dimensional for simplicity and less demanding computability.

2.4.3 Load Combinations and Load Paths

Scaffold systems usually require consideration of different types of load patterns, load
sequences, and load combinations. As a result of concrete operation on support scaffold
systems, several load patterns and sequences normally occur in considering the load
combination of gravity loads and lateral loads. The research by Peng et al. (2003)
presented three different sequential loading patterns on 3-storey scaffold systems
described as model R (rectangle scaffold plan), model L (L-shape scaffold plan), and
model U (U-shape scaffold plan), as shown in Figure 2.9. In all models, sequential paths
Chapter 2 - Literature Review 20
______________________________________________________________________

were investigated and compared with uniform loads. It was shown that the critical loads
of the scaffold system under different sequential paths and uniform loads were about the
same. This finding was in good agreement with the analysis of concrete placement load
effects using influence surfaces from earlier research by Peng et al. (1996c). Thus,
designer can safely assume uniform loads in practical design of the scaffold system. In
terms of load combinations, self weight and imposed (working) load are usually
considered critical in predicting the behaviour and strength of scaffold systems.

Rectangular Shape

L Shape

U Shape

Figure 2.9: Top view of rectangular, L, and U shapes of scaffold systems (adapted from
Peng et al. 2003)

Some researchers such as Milojkovic et al. (2002) have considered wind loads in
perpendicular and parallel directions in their access scaffold model. Godley and Beale
(2001) considered the combinations of dead, imposed and wind loads with different
magnitudes for both in-use and out-of-service conditions of scaffold in construction
practice. For design purposes, the magnitude of imposed (working) load and wind load
Chapter 2 - Literature Review 21
______________________________________________________________________

applied are usually taken from international design codes, such as provided by British
Standards Institution (1996) and Standard Australia (1995).

2.4.4 Initial Imperfections

Scaffold structures are slender by nature; therefore, small initial imperfections


producing member P- and frame P- second order effects must be considered in the
model to accurately predict the behaviour and load carrying capacity of the system
(Figure 2.10). There are many efficient ways of taking geometric imperfection effects
into account. For instance, Chan et al. (2005) considered two types of geometrical
imperfections in portal frames, i.e. imperfections from initial sway and initial member
distortion. The same imperfections were considered in the modelling of scaffold
systems by Yu and Chung (2004). Three methods of modelling imperfections were
trialled, including the scaling of eigenbuckling modes (EBM), the application of
notional horizontal forces (NHF), and the direct modelling of initial geometric
imperfections (IGI). EBM was performed by carrying out eigenbuckling analysis on the
structural model, and then scaling and superimposing the lowest eigenmode onto the
perfect geometry to create an initial imperfect structural frame for the second-order
structural analysis. In the NHF approach, additional lateral point loads were applied at
the top of each column in one direction of the frame and initial member out-of-
straightness could be represented by lateral distributed forces along each member. The
IGI method consisted of applying an initial sway of the frame and an out-of-straightness
to each column in the frame.
Chapter 2 - Literature Review 22
______________________________________________________________________

Figure 2.10: P- and P- effects

For scaffold systems, these same approaches can be applied to model the effects of
initial imperfections in the analysis. For example, Yu et al. (2004) integrated EBM with
the magnitude of the column out-of-straightness of 0.001 of the height of the scaffold
units into the model. Moreover, Yu and Chung (2004) investigated a method called
critical load approach where initial imperfections were integrated directly into a Perry-
Robertson interaction formula to determine the failure loads of the scaffolds in the
analysis. In other research on scaffold systems by Chu et al. (2002), the notional
horizontal force approach was incorporated in the model by applying a horizontal
notional force of 1% of the vertical loads on the scaffold. Godley and Beale (2001)
adopted an initial geometric imperfection approach by imposing a sinusoidal bow to the
members and angular out-of-plumb to the frame. In all these proposed methods, careful
calibration against test results or numerical reference values is required.

2.4.5 Joint Modelling and Boundary Conditions

Scaffold joints are complex in nature due to need for rapid assembly and reassembly in
construction. The Cuplok connections behave as semi-rigid joints, and show looseness
with small rotational stiffness at the beginning of loading. Once the joints lock into
place under applied load, the joints become stiffer as shown by Godley and Beale
Chapter 2 - Literature Review 23
______________________________________________________________________

(1997). Wedge-type joints are generally more flexible and closer to pinned connections.
They also often display substantial looseness at small rotations as discussed by Godley
and Beale (2001). Figure 2.11 shows typical moment-rotation curves for Cuplok joints
and wedge-type joints, investigated by Godley and Beale (1997) and Godley and Beale
(2001). As to spigot joints, out-of-straightness of the standards can occur due to the
space between the standard and the spigot and the lack of fit in the joints can create
complexity in modelling as discussed by Enright et al. (2000). Various scaffold
researchers devised ways in modelling joints; moreover, the study of boundary
conditions of scaffold systems is crucial because the top and bottom restraints can
highly influence the stability and strength of the systems as stated by Yu (2004).

3.5

3.0
Cuplok joint
2.5 Wedge-type joint
M oment (kNm)

2.0

1.5

1.0

0.5

0.0
0.00 0.05 0.10 0.15 0.20
Rotation (radian)

Figure 2.11: Typical moment-rotation curves for Cuplok and wedge-type joints by
Godley and Beale (1997) and Godley and Beale (2001)

In recent research by Peng et al. (2007), analysis models of wedge-type jointed, 3-storey,
3-bay, and 5-row scaffold system were presented. Experimental test on scaffold joints
showed that the joint stiffness was between 4.903 kNm/rad (50 ton cm/rad) and 8.826
kNm/rad (90 ton cm/rad) with the average value of 6.865 kNm/rad (70 ton cm/rad)
being adopted for all joints into their model. Godley and Beale (2001) found that
scaffold connections were frequently made of wedge-type joints, for which the joint
Chapter 2 - Literature Review 24
______________________________________________________________________

stiffness exhibited different response under clockwise and counter-clockwise rotations,


and occasionally exhibited looseness in connections with low stiffness. Consequently,
Prabhakaran et al. (2006) modified the stiffness matrix for the end points of the beam to
include connection flexibility, using a piecewise linear curve to model the moment-
rotation response.

Yu (2004) studied the boundary conditions of the scaffold system, and categorised them
into four cases, i.e. Pinned-Fixed, Pinned-Pinned, Free-Fixed, and Free-Pinned, with the
first term being translational restraint at the top of the scaffold, and the second term
being the rotational restraint at the base of the scaffold. In all analyses, the rotation at
the top was assumed to be free. These conditions were incorporated into the models of
one bay of one-storey modular steel scaffolds (MSS1), and two-storey modular steel
scaffolds (MSS2). Yu found that for MSS1 the failure load results for Free-Fixed and
Pinned-Pinned conditions are reasonably close to test results; however, for MSS2 the
model results are considerably higher than the test results. Subsequently, Yu suggested
that since the top of the scaffolds normally has lateral restraints then joints at the top can
be modelled as translational springs, and for the bottom rotational spring can be applied.
A stiffness of 100 kN/m for the top translational spring and stiffness of 100 kNm/rad for
the bottom rotational spring gave very comparable results to the tests.

In single storey double bay scaffold research by Chu et al. (2002), in the presence of
restraints in the loading beam and the jack bases, the top and the base were modelled
with various boundary conditions, and the scaffold connections were assumed to be
rigid. The researchers found that both Pinned-Pinned and Pinned-Fixed conditions gave
higher load carrying capacities than the experimental results; on the other hand, the
Free-Fixed condition gave satisfactory results compared to the tests. Research on the
stability of single storey scaffold system by Vaux et al. (2002) found that Cuplok
connections represented by pin joints, and connections of the top and bottom jacks to
the standards assumed as rigid with the top-bottom boundary conditions taken as
Pinned-Pinned gave good agreement between numerical and experimental failure loads.

Weesner and Jones (2001) studied the load carrying capacity of three-storey scaffolds
assuming rigid joints between the stories, and pin joints for the top and the bottom
Chapter 2 - Literature Review 25
______________________________________________________________________

boundary conditions. The results of their elastic buckling analysis came out to be rather
larger than the test values with the percentage differences ranging from 6% to 17%. In
the analysis of large access scaffold systems by Godley and Beale (2001), cantilever
arm tests were done on scaffold wedge-type joints. The nonlinear moment-rotation
curve from the tests showed joint looseness and different values of rotational stiffness
under positive (counter-clockwise) rotation and negative (clockwise) rotation. The
authors suggested the use of a multi-linear or nonlinear moment-rotation curve for
scaffold joint modelling.

In the work by Enright et al. (2000), the spigot joints were studied for the stability
analysis of scaffold systems. The spigot inserts were considered to have bending
resistance, but not to transmit axial load; therefore, the model adopted two vertical
members connected by pin joints representing the standards, and on the side, the
entirely rigid spigot member was connected at the top, centre, and bottom to the
standard via short and axially stiff members capable of transferring only lateral forces,
as shown in Figure 2.12. Due to the axial load in the out-of-plumb standards, the spigot
would be in bending, and the amount of bending would depend on the amount of axial
load and the degree of out of plumb. From research of Harung et al. (1975), it was
found that if the spigot joints were modelled as fully continuous joints, the analysis
would overestimate the load carrying capacity of the system.

Axial load

Pin joint
Standards
Spigot

Figure 2.12: Spigot joint model (Enright et al. 2000)


Chapter 2 - Literature Review 26
______________________________________________________________________

Milojkovic B. et al. (1996) studied eccentricity in the modelling of scaffold connections.


Given that the neutral axes of the connections were offset by 50 mm, the authors
modelled the eccentric joint with a finite spring of length equal to the eccentricity of 50
mm. The spring had specific rotational stiffness, and was assumed to be axially stiff.
The authors concluded that for large frames, unless torsion failures can occur, then the
effects of joint eccentricity are insignificant.

In the scaffold study by Gylltoft and Mroz (1995), the braces were represented as truss
members with pinned joints connected to the standards, and the connections between
other members were modelled as short finite elements with nonlinear stiffness in all
directions. To model the shores of the scaffold system, Peng et al. (1997) applied rigid
links with pinned supports at both ends, given that actual shores were connected loosely
by nails at the top and bottom.

2.4.6 Suggestions

To accurately study the behaviour and strength of scaffold systems, geometric and
material, three-dimensional nonlinear analyses are efficient tools. Geometric
imperfections have to be incorporated into the model, in order to consider the second
order effects that exist in the structures. Further research into joint modelling of scaffold
systems is needed since scaffold joints exhibit nonlinear behaviour and present a lack of
fit at early stage. To model the systems accurately, these factors must be taken into
account. Moreover, boundary conditions must be considered carefully since top and
bottom jacks can have eccentricities, which can greatly affect the overall stability of the
system. The degree of rotational and translational fixity over the top and bottom has to
be calibrated correctly to achieve accurate results.
Chapter 2 - Literature Review 27
______________________________________________________________________

2.5 Ultimate Load of Scaffold Systems

2.5.1 Parametric Studies

Yu et al. (2004) investigated the influence of the number of storeys and boundary
conditions on the load carrying capacity. They analysed one, two, and three storey steel
scaffolds, and found that two-storey and three-storey scaffolds had only 85% and 80%
of the load carrying capacity of the single storey steel scaffold respectively because the
different numbers of storey presented considerable variation in buckling behaviour.
Moreover, through different boundary conditions applied at the top and the bottom, the
analytical load carrying capacity varied in the range from 50% to 120% of those of the
experimental tests.

The comprehensive study on wood shoring (simple type of support scaffold systems) of
double-layer systems by Peng (2004) showed the effects of the length of horizontal
stringers (horizontal timbers to connect uprights) and vertical shores, stiffness of
stringers, and positions of strong shores on the load carrying capacities of the shoring
systems. Peng found that adding strong shores (vertical shores with horizontal bracing
in a closed pattern) to the systems could increase the ultimate loads. In contrast, when
the stiffness of the horizontal stringer decreased or the length of the stringers increased
for the cases of unsymmetrical arrangement of strong shores that are at least one
combination of strong shore and leaning column (pinned-ended column) in a vertical
direction as shown in Figures 2.13(c) and 2.13(d), the system ultimate loads would be
reduced; however, for symmetrical cases as shown in Figures 2.13(a) and 2.13(b) the
ultimate loads were unaffected by the change in stringer stiffness. The varied lengths of
vertical shores had different effects depending on the strong shore arrangement in the
system. In addition, strong shores were not as effective when applied at the outmost
location in the system as to apply to the inner. Peng (2002) concluded that the system
ultimate load only increased by adding the strong shores, but not the leaning columns.
Chapter 2 - Literature Review 28
______________________________________________________________________

Strong shore Strong shore


Stringer Stringer
Strong shore Strong shore

(a) (b)

Leaning column Leaning column

Stringer Stringer
Strong shore Strong shore

(c) (d)

Figure 2.13: Model of 2-bay shoring system (adapted from Peng 2004)

Research on the correlation between the load carrying capacity and the number of
storeys of shoring scaffold system by Huang et al. (2000c) showed that the critical loads
of the system reduced rapidly from the range of two to eight storeys, followed by a
gradual decrease thereafter. Furthermore, it was found by Peng et al. (1997) that when
the initial geometric imperfection simulated by 1.50% notional horizontal force
approach was applied at mid-height, the reduction in ultimate load of simple door-type
scaffold systems was found to be around 16%, which was conservative compared to
experimental results, and the relationship between the initial imperfection and reduction
in ultimate load of the scaffold system was nearly linear. With the simulated
imperfection of 0.1% notional horizontal force applied to the model, the predicted
ultimate load showed good agreement with test results. Also, Peng et al. (1997) found
that with long shores installed, the ultimate load of the scaffold system could be as little
as 25% of that of the system without shores. From the analyses of high clearance steel
scaffolds by Peng et al. (1996a), the optimum load carrying capacity for steel scaffolds
with shoring occurred in the range of three to six storeys. In addition, scaffolds of more
Chapter 2 - Literature Review 29
______________________________________________________________________

than eight storeys were not recommended due to the high reduction in strength of such
systems.

Other factors influencing the ultimate load of scaffold systems are bracing arrangement,
loading eccentricity, and ground irregularity. Investigating of bracing arrangement can
help designers to achieve the optimum design for the scaffold system. Loading
eccentricity and ground irregularity exist in any construction; therefore, knowing the
extent of those parameters can reduce the risks in scaffold construction.

2.5.2 Failure Modes

Due to the high slenderness of members in scaffold systems, failure usually occurs by
buckling. The two common types of buckling in scaffold systems are the out-of-plane
mode perpendicular to the plane of scaffold unit and the in-plane mode. The critical
mode depends on the relative stiffness of the connecting members in each direction. The
standards can buckle in single or double curvatures, depending on the configuration of
the scaffolds and support conditions. Figure 2.14 shows common out-of-plane failure
mode of single storey door-type scaffold.

Out-of-plane
In-plane

Figure 2.14: Typical failure mode of single storey door-type scaffold (adapted from Yu
et al. 2004)
Chapter 2 - Literature Review 30
______________________________________________________________________

Yu et al. (2004) performed vertical load tests on multi-storey door-type steel scaffolds.
The researchers found that single storey and double storey scaffolds both buckled out-
of-plane, and deflected in single and double curvatures respectively. In addition, Yu and
co-workers noticed that there were large displacements of the standards in the plane of
the cross-bracings at failure, suggesting that the door-type systems were stiffer in the in-
plane direction. From the three-storey scaffold test results by Weesner and Jones (2001)
on four different door-type frame, most of the scaffolds failed by buckling out-of-plane.
Only one of the tests failed in torsion.

Huang et al. (2000a) carried out tests on one, two and three storey scaffolds test, as
shown in Figure 2.15. The tested scaffold unit was a portal frame with knee braces at
the top. The one-storey scaffolds failed by out-of-plane buckling, whereas the two-
storey and three-storey scaffolds displayed in-plane buckling at failure, and the highest
lateral displacement was found to be at the top of first story, as shown in Figure 2.15.

Figure 2.15: Schematic failure modes of one-to-three storey knee-braced scaffolds


(adapted from Huang et al. 2000a)

From three-dimensional analyses of high clearance steel scaffolds, Peng et al. (1996a)
observed that the deformation modes of the steel scaffolds were dependent on the
Chapter 2 - Literature Review 31
______________________________________________________________________

relative strength between the steel scaffold units and the cross-braces providing lateral
support. If the cross-braces offered stiff lateral support, then the plain scaffold units
would deform in-plane; conversely, the scaffold units would deform in the out-of-plane
direction, in case of flexible cross-braces.

2.5.3 Simplified Equations

Several researchers have proposed simplified equations to determine the ultimate load
of the scaffold system based on structural analysis models and experimental tests.
Huang et al. (2000c) used a two dimensional model to derive a closed-form solution for
the critical load of scaffold systems with knee-braced units. The solution was based on a
bifurcation (eigenvalue) method to the elastic-buckling condition, and the critical loads
were calculated as functions of the material properties, the number of storeys, and the
section properties of the scaffolds. The assumptions for the derivation were as follows:
all members behaved elastically, and the frame buckled in-plane at the lowest storey
(Figure 2.16).

Figure 2.16: Assumption of proposed analytical model (adapted from Huang et al.
2000c)
Chapter 2 - Literature Review 32
______________________________________________________________________

The analytical solution was given as:

2( N 1)(1 sec kL) NkL tan kL tan 2 kL 0 (2.1)

where N = number of storeys; k = effective length factor; and L = one-storey height of


the scaffold unit. From Eq. (2.1), kL could be solved and applied as the effective length
to compute the critical load.

In other research of scaffolds by Huang et al. (2000a), the critical loads were calibrated
and modified based on failure modes and critical loads from the computational critical
loads to the experimental values, then the modified values could be taken as the critical
loads of the scaffold systems for any number of storeys shown in the published graph of
critical loads versus number of storeys (Figure 2.17). In case the scaffold units were
different from the ones used by the authors (portal frames with knee braces), the critical
load should be based on computational results of different section properties that were
functions of the slenderness ( Ei I i / L2i ) of the uprights, and adjusted as:

Ei I i / L2i
( Pcr )i Pcr , graph (2.2)
3830

where ( Pcr ) i = critical load of the scaffold in concern; Pcr , graph = critical load from the

published graph of Huang et al. (2000a); Ei = Youngs modulus in N/cm2; I i = moment

of inertia in cm4; and Li = one-storey height of the scaffold in cm.


Chapter 2 - Literature Review 33
______________________________________________________________________

70

60
Computational critical loads (kN)
50

40

30

20

10

0
0 2 4 6 8 10 12 14 16 18 20
Number of stories

Figure 2.17: Computational critical loads based on two-dimensional model (adapted


from Huang et al. 2000a)

Peng et al. (1998) proposed simple formulae for finding critical buckling loads of
scaffold systems using a sway frame concept. The following equation could be used to
calculate the critical loads:

2 EI
Pcr (2.3)
( KBh) 2

where

6S
1
K 5B (2.4)
S
3
B

and

H ( Bh) 3
I approx (2.5)
3E t
Chapter 2 - Literature Review 34
______________________________________________________________________

where Pcr = critical load of the scaffold; E = Youngs modulus of the scaffold; I =
equivalent moment of inertia of the equivalent column of the scaffold; K = effective
length coefficient; S = ratio of the height of shore extension to the height of the scaffold
unit h ; B = the number of storeys of the system; and t = top horizontal-sway
displacement under a unit horizontal load H at the top from a linear analysis of in-plane,
two-dimensional frame (Figure 2.18). Eq. (2.4) allows for the effect of shoring. In the
case of scaffolds with no shore extensions, K = 0.7 as R = 0. By substituting I approx from

Eq. (2.5) into I of Eq. (2.3), Pcr was found to be close and fairly conservative compared
with the accurate three-dimensional nonlinear analysis of the model.

Figure 2.18: Model for approximating moment of inertia of scaffold (adapted from Peng
et al. 1998)

Based on the analysis results of the high clearance steel scaffold systems, Peng et al.
(1996b) also suggested that the critical loads of the scaffold systems could be quickly
estimated by using a set concept. The set concept utilised the relationship between the
number of steel scaffold sets and the critical load of the systems. For example, the one-
bay-two-row-two-storey scaffold would consist of four sets of scaffold unit, thus by
multiplying the critical load of one unit scaffold with the number of sets (four in this
case), the critical load of the scaffold in interest could be approximated. This method
Chapter 2 - Literature Review 35
______________________________________________________________________

could be applied to scaffolds with shores to estimate the critical load since the ratio of
critical loads between the scaffolds with and without shores was found to be constant
for a given number of storeys. If the critical load of scaffolds without shores is known,
the critical load of scaffolds with shores by the same number of bays, rows, and storeys
can be computed by multiplying the former value by the proposed ratio. These ratios
were presented in the research by Peng et al. (1996b).

2.6 Design of Scaffold Systems

2.6.1 British Standards

BS 5975 by British Standards Institution (1996) provides guidelines for the loads and
load combinations to be applied in the design of falsework. Recommended applied
loads given in this code of practice consist of self-weights, imposed loads, and
environmental loads. The practical design of steel scaffold systems follows the steel
column buckling design method given in BS 5950 by British Standards Institution
(2000) to assess the load carrying capacities based on modified slenderness ratios of the
column members. The summarized design procedure is as follows:

(a) First, the area and the second moment of area of the circular tube member are
calculated respectively as:


A (d e2 di2 ) (2.6)
4

and


I (d e4 di4 ) (2.7)
64

where d e = external diameter of the tube; and di = internal diameter of the tube.
Chapter 2 - Literature Review 36
______________________________________________________________________

(b) The slenderness ratio of the column member is then computed based on the ratio of
the effective height of the scaffold, he , and the radius of gyration, r , as given by:

he
(2.8)
r

where he ke h ; ke = effective length factor (discuss later on); h = height of the

column member between restraints; and r I / A .

(c) The elastic buckling strength of the column member, pE , is then computed by:

2E
pE (2.9)
2

where E is the Youngs modulus.

(d) Finally, the compressive strength of the column member, pc , can be obtained as
follows:

pE p y
pc (2.10)
( pE p y )0.5
2

in which:

p y ( 1) pE
(2.11)
2

where p y = yield strength of the steel tube; the Perry factor, , is calculated as

a ( 0 ) 1000 ; the Robertson constant, a 5.5 for cold-formed steel tubes; and the

limiting slenderness, 0 0.2( 2 E / p y )0.5 .


Chapter 2 - Literature Review 37
______________________________________________________________________

2.6.2 Australian Standards

AS 3610 by Standard Australia (1995) specifies the loads and load combinations to be
applied in the design of formwork assemblies, which can also be adopted for load
calculations in support scaffolds since this type of scaffolds is generally used to carry
loads from concrete construction. The loads are considered in three stages: before,
during, and after concrete placement. These loads consist of vertical loads such as dead
load, concrete load, live load, and material loads, as well as horizontal loads such as
wind loads, and earthquake loads. AS 4100 by Standard Australia (1998) is commonly
applied to the structural design of steel scaffold systems. The simple design procedure
for load capacities of the standard is described as follows:

(a) The area and the second moment of area, A and I , are calculated as in Eq. (2.6) and
(2.7) respectively.

(b) The radius of gyration can be obtained from r I / A , and the form factor ( k f ) is

d f
taken as 1 when the slenderness, e o y 82 for circular tubular members;
t 250
Ae
otherwise k f where d o = outside diameter of the section, t = wall thickness of the
Ag

section, f y is the yield strength of the column tube, Ae = effective area of the section

specified in Clause 6.2.4 of AS 4100 of Standard Australia (1998), and Ag = gross area

of the section. The compression member constant, b is taken as -0.5 for cold-formed
steel tubes.

(c) The effective length, le , is computed as kel where ke = member effective length

factor determined from Clause 4.6.3 in AS 4100 of Standard Australia (1998), and l =
actual length of the standard between restraints.

(d) The modified compression member slenderness, n , is then computed by:


Chapter 2 - Literature Review 38
______________________________________________________________________

le fy
n (2.12)
r 250

(e) The compression member factor, a , is defined as:

2100(n 13.5)
a (2.13)
15.3n 2050
2
n

(f) The elastic buckling load factor, , is given by:

n a b (2.14)

(g) To account for member imperfection, the compression member imperfection factor,
, is calculated as;

0.00326( 13.5) 0 (2.15)

(h) The modified compression member factor, , is defined as:

( 90) 2 1
(2.16)
2( 90) 2

(i) The slenderness reduction factor, c , is determined as:

90
2

c 1 1 (2.17)

(j) The nominal member capacity, N c , is then computed by:

Nc c A f y (2.18)
Chapter 2 - Literature Review 39
______________________________________________________________________

To obtain the design member capacity, the capacity reduction factor, = 0.9 is applied
to N c .

2.6.3 Effective Lengths

The design approaches described in Sections 2.6.1 and 2.6.2 rely on the determination
of the column effective length. Since levels of end restraints of the standards in scaffold
systems are difficult to determine, researchers have proposed values of column effective
length based on buckling analysis of their models. Yu et al. (2004) found that the
effective length coefficients of the door-type steel scaffolds up to three storeys could be
conservatively assumed to be 1.6 for any idealised boundary conditions. Also, they
pointed out that cross-bracings effectively reduced the effective lengths of scaffold
columns. In a separate investigation into the behaviour of door-type steel scaffolds, Yu
(2004) back-calculated the effective lengths from the finite element results of the load
carrying capacities based on various boundary conditions. The effective length factors
were found to be in the range of 1.06 to 1.40. In addition, Harung et al. (1975) proposed
that an effective length of the steel scaffolds should be about 1.2 times the height of
each storey based on the measurement of the largest distance between closest points of
contra-flexure or zero bending moment on the buckled columns in the analysis model.

2.6.4 Bracing Systems

Bracings are important in terms of increasing the stability and the load carrying capacity
of scaffold systems. Peng (2004) studied two different types of bracings for two-layer
shoring system, as shown in Figure 2.19. The V-type bracing was found to be stiffer
than the N-type bracing. The load carrying capacity of the system with V-type bracing
was twice as much as that of the N-type bracing. The study showed that the diagonal
braces offered a very efficient sideway restraint to the system, as confirmed by very
small lateral displacements compared to the shoring system without bracing. Moreover,
Peng et al. (1996b) investigated the effect of bamboo cross-braces on the exterior in-
plane surface of the high-clearance steel scaffold system, and noticed that the critical
load of the scaffold was improved by about 20%; on the other hand, if the braces were
Chapter 2 - Literature Review 40
______________________________________________________________________

fitted to the exterior out-of-plane surface, there was no significant improvement. Further
research on bracing configuration will be useful in determining the optimum design of
scaffold systems.

Figure 2.19: Two layer shoring system with V-type and N-type bracings (adapted from
Peng 2004)

2.6.5 Design by Advanced Structural Analysis

The current practice in designing support scaffold systems is to follow the available
structural design codes or use the load carrying capacity recommended by the
manufacturers based on numbers of full-scale load tests including a judgmental factor of
safety. In the latest load and resistance factor design (LRFD) development, together
with the ready availability of powerful computer and finite element software package,
real behaviour of structural systems can be accurately captured by performing advanced
analysis including material and geometric nonlinearities, semi-rigid joints, residual
stresses, and etc.

In literature, design by advanced analysis has been performed for steel frames as
reported by Buonopane and Schafer (2006), Li and Li (2004), Buonopane et al. (2003),
and Kim and Chen (1999). The ultimate system limit state design procedure starts from
Chapter 2 - Literature Review 41
______________________________________________________________________

determining load combinations and preliminary sizing of members, and then the
nonlinear finite element model including geometric imperfections for the structure is
created. Advanced structural analysis is performed with the incremental combined
factored loads applied to the system until the ultimate load is reached. Finally, the
adequacy of the structural system can be readily evaluated by comparing the system
ultimate strength systemRn with the applied factored loads iQi. The design is
satisfactory if iQi systemRn where i = load factors, Qi = nominal design loads, system
= system resistance factor and Rn = nominal system strength determined by advanced
analysis. If the design is unsatisfactory, the system can be redesigned and the same
procedure is reapplied.

With the design by advanced analysis, it eliminates the need for individual member
checking and provides greater accuracy in determining frame strength than elastic
LRFD, as stated by Buonopane et al. (2003). This method can certainly be applied to
the design of scaffold systems; however, the problem remains as what resistance factors
should be used with advanced analysis. To answer this problem, structural reliability
techniques (Monte Carlo simulations) allow the combination of statistical models of the
random variables affecting the system strength (geometric imperfections, joint stiffness
and material properties) with advanced structural analyses to obtain a reliability analysis
of the system, using either first-order second-moment method (FOSM) or first-order
reliability method (FORM) with suitable load models that will provide probabilistic
justification for appropriate system resistance factors which may be used in the LRFD
design.

2.6.6 Reliability Analysis

The fundamental structural reliability problem is to find the probability of failure Pf of


the structure, as defined by:

Pf P( R Q 0) P(G ( R, Q) 0) FR ( x) f Q ( x)dx (2.19)


Chapter 2 - Literature Review 42
______________________________________________________________________

in which R is the structural resistance or capacity and Q is the total load effect. R and Q
are modelled as random variables. G(R, Q) represents the limit state function and
G(R,Q) 0 defines the unsafe or failure region. FR(x) is the cumulative distribution
function of R, and fQ(x) is the probability density function of Q. Figure 2.20 graphically
summarises a basic structural reliability problem.

Figure 2.20: Basic structural reliability problem (adapted from Ellingwood and
Galambos 1982)

The integration in Eq. (2.19) can be approximated using the first-order second-moment
method (FOSM) by Hasofer and Lind (1974). In the case of statistically independent
normal R and Q, then Pf becomes:

R Q
Pf ( ) ( ) (2.20)
R2 Q2

where ( ) is the standard normal distribution function, and represent the mean and
standard deviation, respectively, and is the so-called reliability index as an alternative
measurement of Pf. A more accurate way to compute is to use the Rackwitz-Fiessler
procedure that requires the knowledge of the probabilistic distributions for all the
variables involved (Rackwitz and Fiessler 1978). This procedure is also known as the
first-order reliability method (FORM). The main idea of this procedure is to calculate
the equivalent normal values of the mean and standard deviation for each non-normal
Chapter 2 - Literature Review 43
______________________________________________________________________

random variable and use them in the analysis in iterative manner. It requires the
following steps:

(a) Formulate the limit state function G and determine the probability distributions
with appropriate parameters for all random variables Xi [i = 1, 2,, n] involved.

(b) Acquire an initial design point [ xi* ] by assuming values (usually mean values)
for n 1 of the random variables Xi and solve the limit state equation G = 0 for
the remaining random variables.

(c) Determine the equivalent normal mean xei and standard deviation xei for each

of the design point values xi* .

(d) Compute the reduced variates [ zi* ] corresponding to the design point [ xi* ] from

xi* xei
zi* (2.21)
xei

(e) Calculate the partial derivatives of the limit state function with respect to the
reduced variates and define a column vector Gpartial as the vector with elements
of partial derivatives multiplied by -1, as described by

G1

G partial G2 where Gi Gz (2.22)
i evaluated at the design po int
Gn

(f) Compute an estimate of using the following equation:


Chapter 2 - Literature Review 44
______________________________________________________________________

z1*


G partial z
T * *

z
where z 2
*
(2.23)
G partial G
T
partial
zn*

(g) Calculate a column vector of the sensitivity factors from


G partial (2.24)
G partial G
T
partial

(h) Find a new design point in reduced variates for n 1 of the variables from

zi* i (2.25)

(i) Compute the values of corresponding design point in original coordinates for the
n 1 using

xi* xei zi* xei (2.26)

(j) Solve for the value of the remaining random variable by setting the limit state
function G = 0.

(k) Repeat steps (c) to (j) until and the design point { xi* } converge to obtain the
results.

For practical use in the design, Eq. (2.19) is complemented by the LRFD formula:

Rn i Qi (2.27)

in which Rn is the nominal resistance determined according to the design procedure, Qi


are the nominal design loads, is the member resistance factor, and i are the load
factors. The partial factors and i are determined using reliability analysis (FOSM or
Chapter 2 - Literature Review 45
______________________________________________________________________

FORM) to achieve a target reliability index for various structural members (Ellingwood
and Galambos 1982). A complete description of FOSM and FORM can be found in
Melchers (1999) and Nowak and Collins (2000).

It should be noted that Eq. (2.27) was originally developed for member capacity design.
For design by advanced analysis, Eq. (2.27) must be applied at the system level, with R
representing the resistance of the system rather than of an individual structural member.
However, developing such an LRFD type formula for a system capacity check on
structures proves to be quite difficult. It involves the following steps:

(a) Identify the appropriate system limit states and failure modes;

(b) Predict accurately the behaviour of the system using advanced analyses;

(c) Find the statistics of system resistance;

(d) Find the statistics of applied loads;

(e) Choose an appropriate target system reliability index;

(f) Solve for system resistance factor and/or load factors using FOSM or FORM.

In literature, Buonopane and Schafer (2006) compared the system reliabilities of a series
of two-story two-bay steel frames designed by elastic LRFD and advanced analysis. It
was shown that for a target system reliability index of 3.0 on ultimate frame strength,
the values of system range from 0.86 to 0.91. Li and Li (2004) developed the resistance
and load factors for advanced analysis of steel portal frames. A design formula was
proposed as 0.8Rn = 1.05Dn + 2.05Ln, in which Dn and Ln represent the nominal dead
load and nominal live load, respectively. A target system reliability index = 3.7 was
achieved. Table 2.1 shows recommended target reliability index () in literature.
Chapter 2 - Literature Review 46
______________________________________________________________________

Table 2.1: Recommended target reliability index in literature

Target
Researcher(s) Structural type Justification
reliability
Steel tension member
3.0 (member)
(yield)

Steel beam 2.5 (member)

Steel column (intermediate


Ellingwood 3.5 (member) Existing loading
slenderness)
and Galambos criteria
(1982) Reinforced concrete beam 3.0 (member)

Reinforced concrete tied


3.5 (member)
column
Masonry unreinforced wall
5.0 (member)
in compression

Steel beam 2.6 (member)

2.7-3.6 Code calibration to


AISC (1986) Steel column
(member) existing design
4.0-6.0
Connection
(member)
Little assistance from
Frame instability under
4.0 (system) non-structural
factored gravity loads
components
Frame instability under
Galambos Resistance against
factored gravity loads with 3.0 (system)
(1990) side-way buckling
cladding

Design to fail as a
Rigid frame 2.5 (system)
plastic mechanism

Common range of
Zhang et al. 3.0-3.5
Pile groups reliability of shallow
(2001) (system)
foundations

Li and Li Steel portal frames with Typical value for


3.70 (system)
(2004) tapered members frame structures

Sivakumar Acceptable reliability


3.0-3.2
Babu and Cantilever retaining walls index for civil
(system)
Basha (2008) engineering structures
Chapter 2 - Literature Review 47
______________________________________________________________________

It should be noticed that earlier research is focused on member design, but in recent
years performance-based design and probabilistic-based design that look at system
behaviour are in attention among international researchers; therefore, the system target
reliability index becomes a main focus in structural system reliability analysis. Design
by advanced analysis is permitted in the Australian steel design code (Standard
Australia 1998); however, no guidance is provided in the code for specifying the value
of system resistance factor, system, suggesting that further research is needed in this area.
In this thesis, one of the aims is to develop LRFD design procedures for support
scaffold systems using the probabilistic techniques presented herein.

2.6.7 Safety in Construction of Scaffold Systems

A monitoring method for support scaffolds was proposed by Huang et al. (2000b) to
prevent collapse. In order to avoid the buckling failure of the standards, the procedure in
construction was to observe axial forces and lateral displacements. From the analyses
and the site tests, the critical locations in the scaffold systems to be monitored for axial
forces were the standards next to the outmost standards along the perimeter and any
locations where heavy loads were expected. As for lateral displacements, the top of the
lowest storey should be observed. The allowable lateral displacement suggested in the
literature by Huang et al. (2000b) was 10 mm.

Strain gauges and linear variable differential transducers (LVDTs) were recommended
for monitoring axial forces and lateral displacements respectively, as suggested by
Huang et al. (2000b). These devices could be connected to a computer to collect real-
time data and send off warning signals when the allowable limits for axial forces and
lateral displacement were approached. Moreover, as suggested by Yu and Chung (2004),
erection tolerances based on construction practice such as out-of-plumb between any
two storeys of the scaffolds and maximum out-of-straightness of each beam or standard
should be limited at 5 mm. Also, the overall out-of-plumb of scaffold structure should
be within a tolerance of 25 mm.
Chapter 2 - Literature Review 48
______________________________________________________________________

2.7 Conclusions

This literature review provides guidelines for modelling, analysis and design of scaffold
systems based on past research. In modelling, initial geometric imperfections that
include sway of the frame and out-of-straightness of the uprights need to be
incorporated so that second-order effects are considered in the nonlinear analysis. The
magnitudes of imperfections applied to the model are usually taken from the available
codes of practice. In many cases, modelling of semi-rigid joints between standard and
ledger based on initial (elastic) rotational stiffness from joint test is adequate; however,
the top and bottom boundary conditions applied for the model are significant in
determining the ultimate load; therefore, careful calibration has to be done.

The most common analyses used in practice are linear elastic buckling and geometric
nonlinear analyses. When linear elastic buckling analysis is applied, the member
buckling load is used in the determination of the moment amplification factor, and this
factor is then applied to the corresponding moment to be used in the design. If
geometric nonlinear analysis is used, the internal axial forces and moments can be
applied directly in the design. Besides, many codes of practice allow the use of
advanced analysis that takes into account of material properties and geometric
imperfections, provided that a structure has sufficient section capacity. However, before
the design by advanced analysis can be fully put into practice, a reliability analysis of
the structural system is necessary in order to establish the required system resistance
factor in LRFD framework.

To ensure safety during construction, support scaffolds should be monitored by their


axial forces and displacements of the standards especially during concrete placement,
and inspected if bracings are applied correctly and adequately. For access scaffolds,
sufficient ties to a permanent structure must be provided to prevent excessive lateral
movement.
Chapter 3 - Geometric Imperfections and Loading Eccentricity of Support Scaffold Systems 49
______________________________________________________________________

CHAPTER 3 - GEOMETRIC IMPERFECTIONS


AND LOADING ECCENTRICITY OF SUPPORT
SCAFFOLD SYSTEMS

3.1 Introduction

Support scaffold systems are used to temporarily support heavy loads such as weight of
fresh concrete, formwork, rebar, equipment and workers. In construction, the
components of scaffold systems are used repeatedly to save cost; therefore, they usually
show signs of wear and out-of-straightness of members. In addition, scaffold systems
generally show some degree of out-of plumb of the frame as a result of uneven ground
surface or erection inaccuracy. Due to heavy loads and geometric imperfections of the
systems, the strength and behaviour of support scaffolds are susceptible to adverse 2nd
order member P- and frame P- effects.

Significant research has been undertaken on methods for incorporating geometric


imperfections in the structural analysis models of steel frames such as the research work
by Chan et al. (2005), Yu and Chung (2004), and Yu et al. (2004). Two main types of
geometric imperfections are particularly important in modelling, consisting of initial
out-of-plumb (sway) and initial member out-of-straightness (crookedness). A common
approach of incorporating geometric imperfections is to scale one or more critical
elastic buckling modes and add the scaled displacements to the perfect geometry.
However, many issues arise with this method, for instance, how many buckling modes
should be included and what scaling factors should be applied? One option is to use the
maximum allowable imperfection values stipulated in available structural codes; for
example, the Australian steel structures design code by Standard Australia (1998)
specifies a maximum out-of-straightness of L/1000, where L is the member length.
Nonetheless, this method often produces conservative predictions of the strength of
systems. An alternative approach is to apply notional horizontal forces, but some doubts
Chapter 3 - Geometric Imperfections and Loading Eccentricity of Support Scaffold Systems 50
______________________________________________________________________

remain about the magnitude of the forces and at which positions they should be applied
on the frame.

In order to resolve the issue of modelling imperfections in scaffold systems, actual data
of geometric imperfections is needed. At present, no available data is published for
geometric imperfections of scaffold systems. This thesis presents the methods and
results in the procurement of geometric imperfection data including measurements of
the loading eccentricity between the timber bearer and the U-head for Cuplok scaffold
systems.

The Cuplok scaffold systems (Figure 3.1) in this research consist of standards, attached
with Cuplok joints, made from cold-formed circular steel tube grade 450 MPa with
nominal outside diameter of 48.3 mm and nominal thickness of 4 mm. The grade 350
MPa ledgers made of steel tube with nominal outside diameter of 48.3 mm and
thickness of 3.2 mm attached with end blades are connected to the standards via Cuplok
joints. The telescopic braces with hook ends were made of 48.3 mm x 4.0 mm outer
tube and 38.2 mm x 3.2 mm inner tube with nominal yield stress of 400 MPa. The
adjustable jacks with 600 mm maximum extension capability were made of 36 mm
diameter grade 430 MPa threaded steel rod and the base plates were 180 mm x 180 mm
x 10 mm in dimension with a nominal yield stress of 250 MPa. The U-heads attached to
the jacks have a clear internal width of 210 mm. More information on the scaffold
components can be found in Boral Formwork & Scaffolding Pty. Ltd. (2002).

The data of geometric imperfections and loading eccentricity was collected from four
different construction sites around Sydney metropolitan area by the use of custom-made
tools, measuring equipment and a theodolite. In order to obtain a wide representative
range of imperfection and loading eccentricity values for support scaffold systems, each
construction site was for different functions including a train station platform,
residential building, parking structure and office building. Moreover, different
supporting ground conditions, including a concrete slab and compacted ground, were
observed from one location to another, likely to lead to a broad range of out-of-plumb
values. At the sites, Cuplok scaffold systems were comprised of 1 to 3 lifts with the lift
heights of 1.0 m and 1.5 m; also, typical jack extensions were varied between 200 mm
Chapter 3 - Geometric Imperfections and Loading Eccentricity of Support Scaffold Systems 51
______________________________________________________________________

and 400 mm. Positions of the spigot joints were observed to be random and usually in
2nd or 3rd lift.

Figure 3.1: Cuplok scaffold systems

3.2 Methods of Procurement

A special-made device instrumented with a dial gauge was used to measure the out-of-
straightness (crookedness) of the standards (uprights) at the mid-height of each lift in
two perpendicular directions aligned with the ledgers and referred to as the N-S and E-
W directions. Figure 3.2 shows schematics of the device in two different lengths that fit
into 1.0 m and 1.5 m lift heights of the scaffold systems for out-of-straightness
measurement. The dial gauge attached to the device was calibrated with a perfectly flat
surface so that it read directly the imperfection value once aligned with the standard.
Figure 3.3(a) shows the actual devices used to measure the out-of straightness at mid-
height of the standard in the lift and Figure 3.3(b) exemplifies the measurement of the
out-of-straightness of a 1.0 m lift standard on one of the sites.
Chapter 3 - Geometric Imperfections and Loading Eccentricity of Support Scaffold Systems 52
______________________________________________________________________

A theodolite was set up to measure the angle difference between top and bottom of the
frame in order to compute storey out-of-plumb (Figure 3.4). The horizontal distance
between the standard (upright) and the theodolite was also measured together with the
angle difference so that a simple trigonometric calculation could be carried out for the
storey out-of-plumb. Measurements were made for both the N-S and E-W axes of the
construction plans.

A vernier calliper was used to measure the loading eccentricity at the top between the
centrelines of the timber bearer and the U-head. All measurements were taken before
the pouring of concrete on to the formwork, representing actual initial geometric
imperfections and loading eccentricity encountered in practice.

48.3 mm

700 mm V-notch plate


V-notch plate
50 mm
Top panel Standard
Mercer dial gauge type 52
35 mm
60x25x10 mm plate
L50x50x3
Bottom panel

V-notch plate

1200 mm Top panel

Mercer dial gauge type 52

60x25x10 mm plate

L50x50x3

Bottom panel

Figure 3.2: Schematic of device used to measure out-of-straightness


Chapter 3 - Geometric Imperfections and Loading Eccentricity of Support Scaffold Systems 53
______________________________________________________________________

(a)

(b)

Figure 3.3: Pictures of (a) device used to measure out-of-straightness (b) actual
measurement
Chapter 3 - Geometric Imperfections and Loading Eccentricity of Support Scaffold Systems 54
______________________________________________________________________

Figure 3.4: Theodolite employed for storey out-of-plumb measurement

3.3 Survey Results of Geometric Imperfections and Loading


Eccentricity

A total of 302 on-site measurements of the out-of-straightness of standards were taken


and 80 measurements of storey out-of-plumb were acquired. In addition, 74
measurements of loading eccentricity were obtained from various construction sites.
From data observation, it was found that there was no correlation between the out-of-
straightness of the standards and the storey out-of-plumb; moreover, the directions
(axes) of these geometric imperfections were random. The directions of the loading
eccentricity were shown to be random and occurred on either side perpendicular to the
bearer. The results of out-of-straightness of the standards were normalised with the lift
height and the results of storey out-of-plumb were normalised with the storey height of
the scaffold system measured from the base plate up to the U-head. Figures 3.5-3.7
show the histograms of normalised out-of-straightness of the standards (/Lh where is
the deflection at mid-height of the lift and Lh is the lift height), normalised storey out-
of-plumb (/H where is the sway and H is the height of the scaffold system) and
loading eccentricity, respectively. A complete data can be found in Appendices A.1-A.3.
Chapter 3 - Geometric Imperfections and Loading Eccentricity of Support Scaffold Systems 55
______________________________________________________________________

120
100
Frequency 80
60
40
20
0

24

28

32
04

08

12

16
0

00

00

00

00
00

00

00
00

0.

0.

0.

0.
0.

0.

0.

0. Normalised out-of-straightness of standard (mm/mm)

Figure 3.5: Histogram of normalised out-of-straightness of the standards (/Lh)

18
16
14
12
Frequency

10
8
6
4
2
0
04

08

12

16

28
24
2
0

00
00

00

00

00

00

00
0.
0.

0.

0.

0.

0.

0.

Normalised storey out-of-plumb (mm/mm)

Figure 3.6: Histogram of normalised storey out-of-plumb (/H)


Chapter 3 - Geometric Imperfections and Loading Eccentricity of Support Scaffold Systems 56
______________________________________________________________________

25

20

Frequency
15

10

0
0 5 10 15 20 25 30 35 40 45 50 55 60
Loading eccentricity (mm)

Figure 3.7: Histogram of loading eccentricity

The mean normalised out-of-straightness of the standards, including standards with and
without spigot joints, is 0.00048 (Lh/2080) with a standard deviation of 0.00042;
however, it is computed that the mean normalised out-of-straightness of the standards
with spigot joints is 0.0013 (Lh/770) with a standard deviation of 0.0008 and the mean
normalised out-of-straightness of the standards without spigot joints is 0.0004 (Lh/2500)
with a standard deviation of 0.0003. The mean normalised storey out-of-plumb is
0.0016 (H/625) with a standard deviation of 0.0005 whereas the mean loading
eccentricity is 18.09 mm with a standard deviation of 10.67 mm.

3.4 Discussion

The geometric imperfection and loading eccentricity data is obtained from Cuplok
scaffold systems; therefore, other support scaffold systems may present a different range
of values of geometric imperfections and loading eccentricity when compared to this
data. Nevertheless, this data gives a general expectation of the level of geometric
imperfections and loading eccentricity in other support scaffold systems. It is possible
that weaker systems, such as wedge-type scaffolds, might present higher values of
geometric imperfections since members of such systems are usually more flexible as
they tend to have smaller cross-sections, and upright-to-beam connections are less stiff.
Chapter 3 - Geometric Imperfections and Loading Eccentricity of Support Scaffold Systems 57
______________________________________________________________________

The level of geometric imperfections in scaffold systems is expected to depend


primarily on section properties, material properties, method of construction and number
of repeated uses. For the loading eccentricity, it is obvious that the magnitude depends
on the clear width of the U-head that the bearer sits on; besides, in good construction
practice the U-head is twisted about a vertical axis until it bears against the timber
bearer to minimise the amount of loading eccentricity.

The mean of the normalised out-of-straightness of the standards without spigot joints
shows a somewhat smaller value than the maximum permissible out-of-straightness of
L/1000 where L is the member length in the Australian design standard for steel
structures by Standard Australia (1998); however, the mean of the normalised out-of-
straightness of the standards with spigot joints is larger than the maximum permissible
out-of-straightness value from the code, suggesting that it is not conservative to adopt
the code value for modelling scaffold lifts with spigot joints. The mean of the
normalised storey out-of-plumb of H/625 is within the maximum allowable out-of-
plumb of H/400 where H is the height of the scaffold system in the Australian standard
for formwork for concrete by Standard Australia (1995).

For the loading eccentricity, the Australian code of formwork for concrete by Standard
Australia (1995) limits the eccentricity at 40 mm or 25% of the bearer width, whichever
is less (18.75 mm for the systems investigated herein). According to the data obtained,
the mean of the loading eccentricity of 18.09 mm is still within this limit; nonetheless,
adopting the code recommended loading eccentricity for modelling and analysis may
not produce conservative results since higher magnitude of loading eccentricity might
occur.

Ultimately, these uncertainties must be studied by statistical methods to find appropriate


distributions so that probabilistic assessment of the strength of the support scaffold
systems can be carried out. The Kolmogorov-Smirnov method is applied for fitting of
suitable probability distributions (D'Agostino and Stephens 1986). Table 3.1 shows
fitted statistical functions for the normalised out-of-straightness of the standards (/Lh),
the normalised storey out-of-plumb (/H) and the loading eccentricity of the support
Chapter 3 - Geometric Imperfections and Loading Eccentricity of Support Scaffold Systems 58
______________________________________________________________________

scaffold systems. The fitted probability distributions are shown graphically in Figures
3.8-3.11.

Table 3.1: Fitted probability distributions for the normalised out-of-straightness, the
normalised out-of-plumb and the loading eccentricity of the support scaffold systems

Probability Standard
Random variable Mean
distribution deviation

Normalised out-of-straightness of
the standards without spigot joints Lognormal 0.0004 0.0003
(/Lh)
Normalised out-of-straightness of
the standards with spigot joints Lognormal 0.0013 0.0008
(/Lh)

Normalised storey out-of-plumb


Normal 0.0016 0.0005
(/H)

Loading eccentricity Lognormal 18.09 mm 10.67 mm

0.4
0.35
Relative frequency

0.3 Lognormal
0.25
0.2
0.15
0.1
0.05
0
0.0000 0.0002 0.0004 0.0006 0.0008 0.0010 0.0012 More
Normalised out-of-straightness of standard without spigot joint
(mm/mm)

Figure 3.8: Fitted lognormal distribution of normalised out-of-straightness of standards


without spigot joints
Chapter 3 - Geometric Imperfections and Loading Eccentricity of Support Scaffold Systems 59
______________________________________________________________________

0.3

0.25
Relative frequency

0.2 Lognormal
0.15

0.1

0.05

0
0.0000 0.0005 0.0010 0.0015 0.0020 0.0025 0.0030 More
Normalised out-of-straightness of standard with spigot joint
(mm/mm)

Figure 3.9: Fitted lognormal distribution of normalised out-of-straightness of standards


with spigot joints

0.4
0.35
Relative frequency

0.3
0.25
0.2
Normal
0.15
0.1
0.05
0
0.0000 0.0004 0.0008 0.0012 0.0016 0.0020 0.0024 0.0028 More
Normalised storey out-of-plumb (mm/mm)

Figure 3.10: Fitted normal distribution of normalised storey out-of-plumb


Chapter 3 - Geometric Imperfections and Loading Eccentricity of Support Scaffold Systems 60
______________________________________________________________________

0.35
0.3
Relative frequency
0.25
0.2
0.15 Lognormal

0.1
0.05
0
0 5 10 15 20 25 30 35 40 45 50 55 More

Loading eccentricity (mm)

Figure 3.11: Fitted lognormal distribution of loading eccentricity

3.5 Conclusions

In this research, measurements of geometric imperfections and loading eccentricity of


Cuplok support scaffold systems have been obtained from various construction sites in
the Sydney metropolitan area, and the results are presented as histograms and fitted
statistical functions. The statistics of the data are presented and compared to the values
from the Australian standards. By code comparison to the mean values, it is shown that
the actual out-of-straightness of standards with spigot joints is over the limit imposed by
the code of Standard Australia (1998); in addition, the magnitude of the loading
eccentricity may be higher than the expected loading eccentricity from the code by
Standard Australia (1995). Consequently, the structural modelling based on existing
codes might lead to less conservative or inadequate design of support scaffold systems.
Nevertheless, the mean of the storey out-of-plumb is shown to be well within the limit
imposed by the code from Standard Australia (1995). The probability distributions for
geometric imperfections and loading eccentricity are presented for further use in
probabilistic assessment of the strength of support scaffold systems in Chapter 8.
Chapter 4 - Cuplok Joint Tests 61
______________________________________________________________________

CHAPTER 4 - CUPLOK JOINT TESTS

4.1 Introduction

Cuplok scaffold systems, renowned type of support scaffold systems, are widely used in
the construction industry for providing general access and supporting vertical loads. For
example, a typical Cuplok system is used to support formwork as shown in Figure 4.1.
The system is fast to erect and easy to disassemble with the Cuplok joint providing
flexibility and ease of construction. Cuplok components consist of a lower cup, which is
welded on to the standard (vertical) at 500 mm intervals, and a sliding upper cup, as
shown in Figure 4.2.

Figure 4.1: Typical Cuplok system

The locking mechanism of Cuplok joints uses no nuts, bolts or wedges; instead the
method of connecting the ledger (horizontal member) to the standard (vertical member)
is by simply positioning the end blades of up to 4 ledgers at desired angles into the
Chapter 4 - Cuplok Joint Tests 62
______________________________________________________________________

lower cup, moving the upper cup down, and rotating it clockwise by hammer blows
until it locks up against a locking bar to achieve a tight connection. Figure 4.3 shows the
components of a Cuplok joint. Conversely, disassembling of the Cuplok joint is done by
applying hammer blows in the counter clockwise direction until the upper cup can be
lifted up to allow the ledgers to be taken out.

Figure 4.2: Typical Cuplok components

Cuplok joints can be assembled into 2-, 3- and 4-way connections at any angles
providing versatility of application, where 2-, 3- and 4-way connections are used at
corner, side and internal joints, respectively. Cuplok parts are made of steel, and,
depending on the manufacturer, they are usually either galvanised or painted for
corrosion resistance, durability, and better handling. These components are commonly
reused in construction practice.

Little information on Cuplok joint tests is available in the literature; only Godley and
Beale (1997) provide some details and results of Cuplok joint tests performed in 1990.
Due to the lack of available test data, this chapter presents tests on Cuplok joints and
discusses the semi-rigid joint behaviour observed from the tests as well as the joint
Chapter 4 - Cuplok Joint Tests 63
______________________________________________________________________

stiffness derived from the tests. Particular attention is paid to the nonlinear moment-
rotation curve. The slope of this curve is a direct measure of the stiffness of the joint,
which can be incorporated in the numerical modelling of Cuplok scaffold system.
Furthermore, since the Cuplok joint tests exhibit considerable variability in joint
stiffness, statistical analyses of the joint test data have been carried out in preparation
for numerical modelling in Chapter 6.

Figure 4.3: Locking mechanism of Cuplok

4.2 Test Setup

The setup was adapted from a multi-purpose test rig described by Lightfoot and Bhula
(1977). Figure 4.4 shows schematically the Cuplok joint test setup. The test rig was
specifically designed and built for the Cuplok joint tests, consisting of a 310UC158
column mounted on to a strong floor and a pivoted U-shaped clamp made from three
lengths of 150x50x5 RHS section, attached to the column by M20 bolts and a rotatable
pin. The clamp was designed to be able to rotate via the pin so that specimens can be
tested in any loading direction. Two 50 mm thick plates were attached to the inside of
the top and bottom tips of the U-shaped clamp to grip the test specimen. Each plate
featured a hole slightly smaller than the nominal diameter of the standard and two M12
bolts for tightening the grip, as shown in Figure 4.4. A hydraulic jack, capable of
producing up to 32-kN of load and attached with a load cell and a half-circular loading
plate, was mounted on to the rail beams of the strong floor to apply loading. Five
Chapter 4 - Cuplok Joint Tests 64
______________________________________________________________________

LVDTs, clamped on external posts, were used to read the displacements at different
locations along the specimen, as shown in Figure 4.4. Flat and smooth plates were
attached with pipe clamps to the specimen at those locations to ensure accurate readings
of the LVDTs. Figure 4.5 shows a typical test setup.

The jack was connected to a hydraulic pump and an MTS controller which operated the
actuator in stroke control mode. A data logger was used for taking measurements of the
LVDTs, load, and stroke. The data logger was connected to a computer with
StrainSmart software and configured to take readings every second. Figure 4.6 shows
detail of the data logging equipment.

Note: LVDTs are clamped


on to movable posts.
Plate attached to a pipe
clamp to ensure complete
contact to the specimen
during loading.

620
150x50x5 RHS 500 Plan view
M12

2000 Specimen
M20
600 32 kN Hydraulic jack with
Pivot
load cell in stroke control
mode
M12

150x50x5 RHS
857
504 784
583

M20
Strong floor Unit: mm Elevation view
Rail beam

Figure 4.4: Schematic of Cuplok joint test setup


Chapter 4 - Cuplok Joint Tests 65
______________________________________________________________________

Figure 4.5: Typical Cuplok joint test setup

Figure 4.6: Test equipment setup


Chapter 4 - Cuplok Joint Tests 66
______________________________________________________________________

4.3 Test Materials

The materials for testings were provided by Boral Formwork and Scaffolding Pty. Ltd.
from available stocks of used components. It was decided that all test specimens should
be sourced from second-hand materials in order to present the real stiffness and strength
of the Cuplok joints being employed in construction practice. The materials supplied by
Boral consisted of 12 Cuplok open ended 2.80 m standards and 27 Cuplok 1.83 m
ledgers. Figure 4.7 shows details of those components. The nominal tubular cross-
section dimensions and yield stress of the Cuplok standards were 48 mm x 4 mm and
450 MPa respectively. Also, the Cuplok ledgers were of nominal tubular dimensions of
48 mm x 3.2 mm with a nominal yield stress of 350 MPa. Some of the standards and
ledgers showed visible out-of-straightness and some Cuplok joints showed signs of
wear. These materials were cut into specific lengths to fit in the test rig. Each standard
was cut into four usable Cuplok joint specimens with an approximate length of 500 mm
and each ledger was cut into 2 usable ledger specimens with an end blade at one end.
The lengths of the ledger specimens were 300 mm for connection elements and 600 mm
for loading elements, as shown in Figure 4.8.

Figure 4.7: Typical Cuplok standard and ledger


Chapter 4 - Cuplok Joint Tests 67
______________________________________________________________________

4.4 Test Series

The objectives of these experiments were to investigate the stiffness and strength of
Cuplok joints in the vertical direction (rotation about the z-axis) and horizontal direction
(rotation about the y-axis), see Figure 4.8. To clearly distinguish the factors affecting
the stiffness, the tests were categorised based on the axis of bending, joint configuration,
and loading direction (up or down and left or right). Moreover, the type of finish
(galvanised or painted), and the number of hammer blows exerted on the cup during
assembly were recorded for each test to study the influence of these factors.

Figure 4.8 shows the bending axes and Figure 4.9 shows different joint configurations
and loading directions being considered. Types A, B and C, and D represent 4-way, 3-
way, and 2-way connections respectively. The loading directions consist of up or down
for bending about the z-axis and left or right for bending about the y-axis. To be able to
differentiate the test results, different labels for each test configuration are used, as
shown in Table 4.1.

Figure 4.8: Joint stiffness axes


Chapter 4 - Cuplok Joint Tests 68
______________________________________________________________________

Figure 4.9: Joint configurations and loading directions in top view


Chapter 4 - Cuplok Joint Tests 69
______________________________________________________________________

Table 4.1: Labels for test configurations

Label Bending axis Joint type Loading direction

KzA1 z A (4-way) 1(down)


KzB1 z B (3-way ) 1(down)
KzC1 z C (3-way ) 1(down)
KzD1 z D (2-way) 1(down)
KzA2 z A (4-way) 2(up)
KzB2 z B (3-way ) 2(up)
KzC2 z C (3-way ) 2(up)
KzD2 z D (2-way) 2(up)
KyA1 y A (4-way) 1(right)
KyB1 y B (3-way ) 1(right)
KyC1 y C (3-way ) 1(right)
KyD1 y D (2-way) 1(right)
KyC2 y C (3-way ) 2(left)
KyD2 y D (2-way) 2(left)

4.5 Test Procedure

Prior to testing, the five LVDTs were calibrated based on their displacement recordings
from StrainSmart software against displacement readings obtained using a digital
vernier caliper with an accuracy of 0.01 mm. The load cell was also calibrated prior to
testing. The Cuplok joint was assembled based on the required testing joint
configuration and then clamped onto the test rig and rotated via the pivot to the desired
loading position. Afterwards, the clamp was fastened to the column by M20 bolts, and
the LVDTs were fitted at five locations along the specimen, as shown in Figure 4.10.
Chapter 4 - Cuplok Joint Tests 70
______________________________________________________________________

620
315

LVDT

LVDT

LVDT
LVDT 3 2 1

85 100 Specimen

270

5
LVDT Loading
48

Figure 4.10: Typical numberings and location of LVDTs on test specimen

The hydraulic jack was programmed to run at the rate of 3 mm/min in stroke control
mode and the data logger was set to record data at 1 scan/second. The half-circular
loading plate on the hydraulic jack was then moved toward the specimen until barely
touching before zeroing all recording values. The recording of data and the movement
of the jack were started simultaneously at the commencement of the test.

Some of the tests of each configuration were kept in the elastic range and the Cuplok
joints were retested by varying the number of hammer blows (3-to-7) applied to tighten
the cup. These tests were carried out to compare the initial stiffness of the Cuplok joints.
Also, for each configuration two types of finish (galvanised or painted components)
were tested to see if the finish resulted in significant differences in joint stiffness.
Finally, some of the tests of each configuration were continued into plastic range to
obtain the joint strength.

4.6 Test Results

From the data recorded, the joint behaviour was investigated by determining the
relationship between moment and rotation. The moment, M, is calculated as
Chapter 4 - Cuplok Joint Tests 71
______________________________________________________________________

M FL (4.1)

where F is the applied load and L is the perpendicular distance between load application
point and the centre of the joint (the centre line of the standard). The rotation, , is given
as

2 3 4
arctan arctan 5 (4.2)
d d d
23 5 4

where 2, 3, 4, and 5 are the displacement of LVDT2, LVDT3, LVDT4 and LVDT5
respectively, d2-3 is the distance between LVDT2 and LVDT3, d4 is the perpendicular
distance between LVDT4 and the centre of the joint, and d5 is the perpendicular distance
between LVDT5 and centre of the joint. The LVDT numbering system is shown in
Figure 4.10. LVDT1 and stroke readings are used only for calibration.

After calculation of moment and rotation, moment-rotation curves were plotted for each
joint configuration, as categorised in Table 4.1. Figures 4.11, 4.13, 4.15, and 4.17 show
the results for joints bending downward about the z-axis for different joint
configurations. Figures 4.12, 4.14, 4.16, and 4.18 present the results for joints bending
upward about the z-axis with different joint configurations. Figures 4.19, 4.20, 4.21 and
4.23 illustrate the results for joints bending to the right about the y-axis for four
different joint configurations. Figures 4.22 and 4.24 present the results for joints
bending to the left about the y-axis for two different joint configurations. For each
configuration, a number of tests are shown correspondingly to different combinations of
the number of hammer blows and type of finish (galvanised or painted components).
Test 1 to test 5 and test 6 to test 10 corresponded to 7 hammer blows and 3 hammer
blows respectively, and test 11 to test 14 were all tightened using 5 blows. Galvanised
components were used in test 1 to test 5, test 11 and test 13. Painted components were
used in test 6 to test 10, test 12 and test 14.
Chapter 4 - Cuplok Joint Tests 72
______________________________________________________________________

4.5

4 Test 1
Test 2
3.5
Test 3
3 Test 4
Moment (kNm)

Test 5
2.5
Test 6
2 Test 7
Test 8
1.5
Test 9
1 Test 10
Test 11
0.5
Test 12
0 Test 13
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 Test 14
Rotation (rad)

Figure 4.11: Cuplok test results for KzA1

4.5

4 Test 1
Test 2
3.5
Test 3
3 Test 4
Moment (kNm)

Test 5
2.5
Test 6
2 Test 7
Test 8
1.5
Test 9
1 Test 10
Test 11
0.5
Test 12
0 Test 13
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 Test 14
Rotation (rad)

Figure 4.12: Cuplok test results for KzA2


Chapter 4 - Cuplok Joint Tests 73
______________________________________________________________________

4.5

4 Test 1
Test 2
3.5
Test 3
3 Test 4
Moment (kNm)

Test 5
2.5
Test 6
2 Test 7
Test 8
1.5
Test 9
1 Test 10
Test 11
0.5
Test 12
0 Test 13
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 Test 14
Rotation (rad)

Figure 4.13: Cuplok test results for KzB1

4.5

4 Test 1
Test 2
3.5
Test 3
3 Test 4
Moment (kNm)

Test 5
2.5
Test 6
2 Test 7
Test 8
1.5
Test 9
1 Test 10
Test 11
0.5
Test 12
0 Test 13
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 Test 14
Rotation (rad)

Figure 4.14: Cuplok test results for KzB2


Chapter 4 - Cuplok Joint Tests 74
______________________________________________________________________

4.5

4 Test 1
Test 2
3.5
Test 3
3 Test 4
Moment (kNm)

Test 5
2.5
Test 6
2 Test 7
Test 8
1.5
Test 9
1 Test 10
Test 11
0.5
Test 12
0 Test 13
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 Test 14
Rotation (rad)

Figure 4.15: Cuplok test results for KzC1

4.5

4 Test 1
Test 2
3.5
Test 3
3 Test 4
Moment (kNm)

Test 5
2.5
Test 6
2 Test 7
Test 8
1.5
Test 9
1 Test 10
Test 11
0.5
Test 12
0 Test 13
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 Test 14
Rotation (rad)

Figure 4.16: Cuplok test results for KzC2


Chapter 4 - Cuplok Joint Tests 75
______________________________________________________________________

4.5

4 Test 1
Test 2
3.5
Test 3
3 Test 4
Moment (kNm)

Test 5
2.5
Test 6
2 Test 7
Test 8
1.5
Test 9
1 Test 10
Test 11
0.5
Test 12
0 Test 13
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 Test 14
Rotation (rad)

Figure 4.17: Cuplok test results for KzD1

4.5

4 Test 1
Test 2
3.5
Test 3
3 Test 4
Moment (kNm)

Test 5
2.5
Test 6
2 Test 7
Test 8
1.5
Test 9
1 Test 10
Test 11
0.5
Test 12
0 Test 13
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 Test 14
Rotation (rad)

Figure 4.18: Cuplok test results for KzD2


Chapter 4 - Cuplok Joint Tests 76
______________________________________________________________________

0.7

0.6
Test 1
Test 2
0.5
Test 3
Moment (kNm)

Test 4
0.4
Test 5
Test 6
0.3
Test 7
Test 8
0.2
Test 9
0.1 Test 10

0
0 0.02 0.04 0.06 0.08 0.1 0.12
Rotation (rad)

Figure 4.19: Cuplok test results for KyA1

0.7

0.6
Test 1
Test 2
0.5
Test 3
Moment (kNm)

Test 4
0.4
Test 5
Test 6
0.3
Test 7
Test 8
0.2
Test 9
0.1 Test 10

0
0 0.02 0.04 0.06 0.08 0.1 0.12
Rotation (rad)

Figure 4.20: Cuplok test results for KyB1


Chapter 4 - Cuplok Joint Tests 77
______________________________________________________________________

0.7

0.6
Test 1
Test 2
0.5
Test 3
Moment (kNm)

Test 4
0.4
Test 5
Test 6
0.3
Test 7
Test 8
0.2
Test 9
0.1 Test 10

0
0 0.02 0.04 0.06 0.08 0.1 0.12
Rotation (rad)

Figure 4.21: Cuplok test results for KyC1

0.3

0.25 Test 1
Test 2
0.2 Test 3
Moment (kNm)

Test 4
Test 5
0.15
Test 6
Test 7
0.1 Test 8
Test 9
0.05 Test 10

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04
Rotation (rad)

Figure 4.22: Cuplok test results for KyC2


Chapter 4 - Cuplok Joint Tests 78
______________________________________________________________________

0.7

0.6
Test 1
Test 2
0.5
Test 3
Moment (kNm)

Test 4
0.4
Test 5
Test 6
0.3
Test 7
Test 8
0.2
Test 9
0.1 Test 10

0
0 0.02 0.04 0.06 0.08 0.1 0.12
Rotation (rad)

Figure 4.23: Cuplok test results for KyD1

0.3

0.25 Test 1
Test 2
0.2 Test 3
Moment (kNm)

Test 4
Test 5
0.15
Test 6
Test 7
0.1 Test 8
Test 9
0.05 Test 10

0
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08
Rotation (rad)

Figure 4.24: Cuplok test results for KyD2


Chapter 4 - Cuplok Joint Tests 79
______________________________________________________________________

The elastic Cuplok joint stiffness was determined from the initial slope of the moment-
rotation curves. Table 4.2 summarises the elastic Cuplok joint stiffness for each test for
the various configurations. The tests show that approximately 30% of the joints
exhibited looseness at the beginning of loading, notably among the joints loaded in
vertical bending (rotation about the z-axis). A statistical analysis of the coordinates of
the moment-rotation curves corresponding to the end of the initial looseness range for
those joints exhibiting looseness shows that looseness is overcome when the moment
reaches 0.4 kNm with a corresponding average rotation of 0.01 rad.

The ultimate moment capacity of the joints bent about the z-axis is about 3.5 kNm on
average. In contrast, joints bent about the y-axis have a very small moment capacity of
0.4 kNm on average and much smaller initial stiffness than the joints bent about the z-
axis. Most of the Cuplok joints have a large plastic range associated with large rotation,
except a few of the joints that failed abruptly as a result of fracture at the welded
connection of the end blade of ledger. Figure 4.25 shows a Cuplok component with
plastic deformations and minor cracks after being tested.

Ledger
Cuplok standard

Figure 4.25: Cuplok components after being tested


Chapter 4 - Cuplok Joint Tests 80
______________________________________________________________________

Table 4.2: Summary of initial (elastic) Cuplok joint stiffness

Initial (elastic) Cuplok joint stiffness (kNm/rad)


Test KzA1 KzB1 KzC1 KzD1 KzA2 KzB2 KzC2 KzD2
1 110 100 96 82 113 105 113 88
2 109 92 90 90 130 63 135 94
3 112 95 91 99 120 78 128 55
4 124 96 95 100 120 78 115 75
5 150 101 104 97 113 80 117 99
6 100 60 75 65 113 80 96 94
7 160 70 100 60 135 115 117 90
8 105 80 120 53 118 118 120 83
9 102 75 118 54 133 117 120 85
10 90 73 100 59 138 125 125 79
11 85 79 83 60 90 79 90 85
12 75 68 85 55 88 80 65 100
13 80 75 92 88 113 84 60 68
14 105 100 70 75 115 100 72 80

Initial (elastic) Cuplok joint stiffness (kNm/rad)


Test KyA1 KyB1 KyC1 KyD1 KyC2 KyD2
1 22 12 10 2 13 2
2 20 12 9 3 10 3
3 21 15 12 3 11 4
4 16 17 12 2 20 4
5 18 13 9 3 12 5
6 10 12 10 3 15 1
7 13 10 15 8 22 3
8 30 12 17 9 20 4
9 17 11 8 10 21 5
10 29 13 12 5 16 8
Chapter 4 - Cuplok Joint Tests 81
______________________________________________________________________

4.7 Discussion

By examining the test results, the number of hammer blows applied to tighten the cup
and the type of finish (galvanised or painted) have no substantial influence on the joint
stiffness. According to construction practice, the hammer blows on the cup are usually a
minimum of three strong blows suggesting that the Cuplok joint is completely fastened.
More blows, up to seven as investigated herein, do not increase the joint stiffness.
Moreover, galvanised or painted components have insignificant effects on the joint
stiffness. The comparisons of these results based on initial (elastic) joint stiffness are
shown in Appendix B.1. Nonetheless, the joint stiffness depends greatly on the axis of
bending. A statistical analysis is presented in Table 4.3 for each test configuration based
on different tested joints.

Table 4.3: Statistical results for initial (elastic) Cuplok joint stiffness

Initial (elastic) Cuplok joint stiffness (kNm/rad)


KzA1 KzB1 KzC1 KzD1 KzA2 KzB2 KzC2 KzD2
Mean 96.23 81.73 87.97 71.62 108.54 88.99 87.32 83.07
Standard deviation 23.28 13.29 13.62 17.46 14.25 18.99 23.48 11.86
Coefficient of
0.24 0.16 0.15 0.24 0.13 0.21 0.27 0.14
variation

Initial (elastic) Cuplok joint stiffness (kNm/rad)


KyA1 KyB1 KyC1 KyD1 KyC2 KyD2
Mean 19.50 12.65 11.38 4.77 15.95 3.87
Standard deviation 6.09 2.06 2.71 2.91 4.28 1.82
Coefficient of variation 0.31 0.16 0.24 0.61 0.27 0.47

The statistical results show that Cuplok joints bent in the vertical plane (rotation about
the z-axis) are much stiffer than when bent in the horizontal plane (rotation about the y-
axis), and that the 4-way joint connection is the stiffest among all other configurations,
particularly the 2-way joint connection which has the lowest stiffness. Also, both types
of 3-way connections provide similar stiffness for the same bending axis. The difference
Chapter 4 - Cuplok Joint Tests 82
______________________________________________________________________

in vertical stiffness between 4-way and 2-way joints is about 30%. Besides, the upward
bending stiffness tends to be slightly higher than the downward bending stiffness which
is rational since the degree of fixity in the welded lower cup is higher than that of the
locking upper cup and hence less joint movement is possible under upward bending.
The coefficient of variation values shown in Table 4.3 suggest that the variation in
Cuplok joint stiffness is substantial and hence may significantly influence the ultimate
strength of scaffold systems.

The relation between the moment and rotation of the Cuplok connections can be
modelled by a tri-linear curve, as illustrated in Figure 4.26. The parameters that describe
the tri-linear curve (k1, k2, k3, 1, 2, 3) were obtained from the joint test data presented
herein. The mean values for k1, k2, and k3 for different joint configurations and bending
axes are shown in Table 4.4. The mean values of k1 are presented for joint test data
considering looseness alone (moment-rotation curves of the Cuplok joints that have
looseness), as well as with and without looseness. The slopes k2 and k3 represent elastic
and plastic joint stiffness, respectively. The mean values for 1, 2 and 3 are presented
for different joint configurations and bending axes in Table 4.5.

Since the probabilistic assessment of the strength of scaffold systems depends mainly
on the response of joints to bending about the horizontal axis, but not about the vertical
axis, only variation in the vertical bending stiffness may need to be considered. The
modelling of the horizontal bending stiffness can be assumed to be deterministic using
the mean values for the tri-linear curve, as presented in Tables 4.4-4.5 for bending about
the vertical axis.

For practicality in probabilistic modelling, the values of 1, 2 and 3 are assumed to be


deterministic, taken as their mean values; however, the variations of k1, k2, and k3 in
bending about the horizontal axis are taken into account. The coefficient of variation
values of the joint stiffness for k1, k2, and k3 for bending about the horizontal axis for
different joint configurations are presented in Table 4.6. The coefficient of variation
values of k1 are presented for joint test data considering looseness alone, and with and
without looseness. It is expected that the probabilistic modelling of joint stiffness with
Chapter 4 - Cuplok Joint Tests 83
______________________________________________________________________

k1 considering looseness alone will produce conservative results of the strength of


scaffold systems.

The joint stiffness values (k1, k2, k3) for bending about the horizontal axis are
normalised with their mean values for each configuration and fitted to normal
distributions for probabilistic modelling purposes, as shown in Appendix B.2 with
corresponding statistical functions. In the probabilistic model of joint stiffness, normal
distributions can be applied for k1, k2, and k3 to generate random values of normalised
stiffness with corresponding mean and coefficient of variation, as shown in Appendix
B.2. Random values of joint stiffness can be obtained by multiplying random
normalised stiffness values by their corresponding mean joint stiffness values as
obtained from Table 4.4, for each joint configuration.

Moment

k3

k2

k1

Rotation
1 2 3

Figure 4.26: Tri-linear moment-rotation for the Cuplok joints


Chapter 4 - Cuplok Joint Tests 84
______________________________________________________________________

Table 4.4: Mean Cuplok joint stiffness (kNm/rad)

Bending about horizontal axis Bending about vertical axis


k1
Joint With and k2 k3 k1 k2 k3
configuration Looseness without
alone
looseness

4-way 39 80 102 5.3 15 7.5 0.8

3-way 36 75 87 5.1 14 7 1

2-way 41 70 77 4.6 7.5 5 1.5

Table 4.5: Mean rotation for Cuplok joints (rad)

Bending about horizontal axis Bending about vertical axis


Joint
1 2 3 1 2 3
configuration

4-way 0.014 0.036 0.16 0.02 0.04 0.1

3-way 0.012 0.036 0.16 0.02 0.04 0.1

2-way 0.007 0.036 0.16 0.02 0.04 0.1

Table 4.6: Coefficient of variation of Cuplok joint stiffness

Bending about horizontal axis


k1

Joint configuration With and k2 k3


Looseness
without
alone
looseness

4-way 0.22 0.35 0.18 0.30

3-way 0.38 0.37 0.21 0.37

2-way 0.35 0.27 0.20 0.46


Chapter 4 - Cuplok Joint Tests 85
______________________________________________________________________

4.8 Conclusions

This chapter investigates the joint stiffness of Cuplok scaffold systems. Using second-
hand Cuplok components provided by Boral Formwork and Scaffolding Pty. Ltd.,
Cuplok joints were assembled and tested in a specially designed rig in the structures
laboratory of the School of Civil Engineering at the University of Sydney. The joint
tests featured bending about two distinct axes, different loading directions, four types of
joint configuration, two finishes (galvanised or painted components), and different
numbers of hammer blows to tighten the cup. The results were presented in the form of
moment-rotation curves, the initial slopes of which were used to determine the joint
stiffness. The joint stiffness for each test configuration was tabulated and a statistical
analysis was performed. The stiffness for bending in the vertical plane was found to be
much higher than the stiffness for bending in the horizontal plane, and 4-way
connections provided greater stiffness than other joint configurations. However, the
surface finish (galvanised or painted components), and the number of hammer blows
used to tighten the connection showed insignificant effect on the joint stiffness. The
joint stiffness data was fitted to normal distributions for probabilistic modelling. The
results from the investigation of Cuplok joint behaviour are useful for modelling and
performing probabilistic analysis of the strength of Cuplok scaffold systems, which are
discussed in Chapters 6 and 8.
Chapter 5 - Formwork Subassembly Tests 86
______________________________________________________________________

CHAPTER 5 - FORMWORK SUBASSEMBLY


TESTS

5.1 Introduction

A total of 18 support scaffold subassembly tests were conducted at the University of


Sydney in 2006 by the Centre for Advanced Structural Engineering to study the
behaviour and ultimate load-carrying capacities of such systems. Supplementary tests
on components of Cuplok scaffold systems, including tensile coupon, stub column and
bending tests, were carried out. The complete details and results were published in
CASE (2006) as part of a consultancy report for Boral Formwork and Scaffolding Pty.
Ltd. This chapter summarises the data and findings that were used for modelling and
calibrations of support scaffold systems in Chapter 6.

5.2 Setup and Procedures

In all tests, the formwork subassembly, also known as Cuplok scaffold system, was
constructed as a grid frame of three-by-three bays with a constant nominal bay width of
1829 mm in both directions (common Boral construction practice). The first fourteen
tests were systems of three lifts with equal nominal lift height of 1.5 m. The systems
featured different bracing arrangements (full, none, perimeter, core, and North-South
direction only) and different jack extension heights (300 mm or 600 mm) at both top
and bottom. In the last four subassembly tests, systems with 1 m and 2 m lift heights
with full bracing were tested with the variation in the number of lifts from 2 to 4 and
jack extension heights of 300 mm or 600 mm.

All testing components were taken from stocks of used material. As a result, the testing
materials showed some geometric imperfections representing those encountered in
practice, particularly in regard to the out-of-straightness of the standards. Besides, the
Chapter 5 - Formwork Subassembly Tests 87
______________________________________________________________________

Cuplok joints showed signs of wear from frequent use and were therefore representative
of joints used in practice in terms of joint stiffness and strength.

The standards, attached with Cuplok joints, were made from cold-formed circular steel
tube (CHS) grade 450 MPa with nominal outside diameter of 48.3 mm and thickness of
4 mm. The grade 350 MPa (CHS) ledgers with end blades were of nominal outside
diameter of 48.3 mm and thickness of 3.2 mm. The telescopic (CHS) braces with hook
ends were made of 48.3 mm x 4.0 mm outer tube and 38.2 mm x 3.2 mm inner tube
with nominal yield stress of 400 MPa. The adjustable jacks were made of 36 mm
diameter threaded steel rod of grade 430 MPa. The base plates were 180 mm x 180 mm
x 10 mm in dimension with a nominal yield stress of 250 MPa. Some of the standards
and ledgers were reused in several tests given that their permanent deformations stayed
within tolerance limits. In addition, it was checked that over a length of 2000 mm, the
out-of-straightness of the uprights was less than 5 mm (CASE 2006).

A test frame was constructed specially for the subassembly tests consisting of four
loading beams at both the top and bottom running in the North-South direction. A total
of sixteen hydraulic jacks attached to the top loading beams (four hydraulic jacks per
loading beam) were used to load the (150 mm x 77 mm) timber bearers running in the
East-West direction, which transferred loading to the top of each standard via (210 mm
clear width) U-heads. Four sets of cross-bracing were installed to prevent sway of the
top loading beams. Ball bearings were inserted between the jacks and timber bearers in
the first test. However, this condition produced excessive lateral displacements and
rotations of the adjustable jacks and was deemed non-representative of construction
practice, where the bearers support a series of closely spaced secondary bearers running
orthogonally to the main bearers. These secondary bearers elastically restrain
displacements and rotations of the primary bearers. Consequently, the results of Test No.
1 were discarded and in subsequent tests, secondary bearers spaced approximately at
600 mm were attached at the top of primary bearers. Figure 5.1 illustrates a typical test
frame showing primary and secondary bearers and bracings. The results of Test No. 7
were also discarded since the hydraulic jacks accidentally moved out of positions during
the test.
Chapter 5 - Formwork Subassembly Tests 88
______________________________________________________________________

Hydraulic jack

Secondary bearer
Primary bearer

Bracing

Figure 5.1: Typical test frame from top view (CASE 2006)

In all tests except Test No. 6, the loads were applied at an eccentricity of 25 mm in the
North-South direction to the top adjustable jacks (Figure 5.2(a)) along the second row in
the East-West direction while for the rest of the standards the loads were applied
concentrically. In addition, the base plates in the row of eccentrically loaded standards
were placed on 3 mm diameter circular steel rods at a nominal eccentricity of 15 mm, as
shown in Figure 5.2(b), as per AS 3610 by Standard Australia (1995).

The load eccentricities applied at the top and bottom of the system were arranged such
that the standards were bent in single curvature. The loads were applied equally by
Chapter 5 - Formwork Subassembly Tests 89
______________________________________________________________________

hydraulic jacks on each standard through primary bearers, except in Test No. 14 where
the loads applied to the corner, perimeter, and centre jacks were in the ratio of 1:2:4
respectively. The tests were conducted in load control and hence, it was not possible to
capture the post-ultimate response. The applied loads were recorded at each increment
of loading until failure occurred when large displacements were observed, and
theodolites were employed to measure initial geometric imperfections of all the
standards before the test began and the displacements of six selected standards during
loading.

Hydraulic jack S N

Primary bearer

U-head

Base plate

3 mm steel rod
Unit: mm (b)

(a)

Figure 5.2: Enlarged view of (a) top eccentricity and (b) bottom eccentricity (CASE
2006)

5.3 Test Configurations

A summary of the test configurations which includes test number, test date, lift height,
number of lifts, top and bottom jack extension length, position of spigot, bracing
arrangement, type of loading, and loading eccentricity is presented in Table 5.1.
Chapter 5 - Formwork Subassembly Tests 90
______________________________________________________________________

Table 5.1: Summary of test configurations

Loading
No.
Lift Jack Spigot eccentricity
Test Date of Bracing Loading
height extension (lifts) in 2nd row
lifts
standards
25 mm top &
2nd &
1 17/1/06 1.5 m 3 600 mm full uniform 15 mm
3rd
bottom
25 mm top &
2nd &
2 31/1/06 1.5 m 3 600 mm full uniform 15 mm
3rd
bottom
25 mm top &
2nd &
3 8/2/06 1.5 m 3 600 mm full uniform 15 mm
3rd
bottom
25 mm top &
2nd &
4 13/2/06 1.5 m 3 600 mm none uniform 15 mm
3rd
bottom
25 mm top &
2nd &
5 16/2/06 1.5 m 3 600 mm perimeter uniform 15 mm
3rd
bottom

2nd & 15 mm
6 22/2/06 1.5 m 3 600 mm perimeter uniform
3rd bottom

25 mm top &
2nd &
7 3/3/06 1.5 m 3 300 mm full uniform 15 mm
3rd
bottom
25 mm top &
2nd &
8 10/3/06 1.5 m 3 300 mm full uniform 15 mm
3rd
bottom
25 mm top &
2nd &
9 17/3/06 1.5 m 3 300 mm none uniform 15 mm
3rd
bottom
25 mm top &
2nd &
10 24/3/06 1.5 m 3 300 mm core uniform 15 mm
3rd
bottom
25 mm top &
2nd &
11 30/3/06 1.5 m 3 300 mm full uniform 15 mm
3rd
bottom
Chapter 5 - Formwork Subassembly Tests 91
______________________________________________________________________

Loading
No.
Lift Jack Spigot eccentricity
Test Date of Bracing Loading
height extension (lifts) in 2nd row
lifts
standards
25 mm top &
12 7/4/06 1.5 m 3 300 mm 2nd full uniform 15 mm
bottom
25 mm top &
2nd &
13 13/4/06 1.5 m 3 300 mm N-S only uniform 15 mm
3rd
bottom
25 mm top &
2nd &
14 3/5/06 1.5 m 3 300 mm core 1:2:4 15 mm
3rd
bottom
25 mm top &
nd
15 15/5/06 2m 2 300 mm 2 full uniform 15 mm
bottom
25 mm top &
16 19/5/06 1m 3 600 mm 3rd full uniform 15 mm
bottom
25 mm top &
17 7/6/06 1m 4 300 mm 3rd full uniform 15 mm
bottom
25 mm top &
18 30/6/06 1m 4 300 mm 3rd full uniform 15 mm
bottom

A schematic of a typical test configuration is shown in Figure 5.3. The figure shows a
typical 3-lift support scaffold system with a lift height of 1.5 m in a full bracing
arrangement being loaded by hydraulic jacks attached to the test frame. The figure also
shows a temporary support deck for access to the test frame. The bracing arrangement is
according to the labelling in the figure showing core and perimeter braces. Full bracing
arrangement includes both core and perimeter braces. N-S bracing arrangement consists
of braces running in the North-South direction. As an example, Figure 5.4 shows the
actual Test No. 8 setup consisting of a 3-lift, 1.5 m lift height and 300 mm jack
extension subassembly test with full bracing arrangement including top and bottom
eccentricities in 2nd row.
Chapter 5 - Formwork Subassembly Tests 92
______________________________________________________________________

Figure 5.3: Typical test configuration in plan and elevation view (CASE 2006)
Chapter 5 - Formwork Subassembly Tests 93
______________________________________________________________________

Ledger

Standard Brace

West East

Figure 5.4: Test No. 8 setup (CASE 2006)

5.4 Test Results

A summary of the test results consisting of test number, ultimate load from hydraulic
jacks at 3 different locations (corner, perimeter, and centre), and observed failure mode
is shown in Table 5.2. As noted earlier, the results of Tests No. 1 and 7 are
unrepresentative, and thus not shown in the summary.
Chapter 5 - Formwork Subassembly Tests 94
______________________________________________________________________

Table 5.2: Summary of test results

Ultimate Ultimate Ultimate


load at load at load at
Test Observed failure mode
corner perimeter centre
jacks (kN) jacks (kN) jacks (kN)
N-S sway mode, final failure of top
2 87 86 89
and bottom jacks
N-S sway mode, failure of top and
3 91 90 91
bottom jacks, and spigot
N-S sway mode, final failure of top
4 50 50 50
and bottom jacks
N-S sway of centre bay, final failure
5 60 60 60
of top jacks
N-S sway of centre bay, final failure
6 60 60 60
of top jacks and spigot
Some N-S sway, failure of standards
8 130 130 130
and spigot at top lift
N-S sway mode, final failure of top
9 65 65 65
jacks and top standards
N-S sway mode, failure of corner
10 70 70 70
standards and spigots
Some N-S sway, final failure of top
11 120 120 120
spigots and standards
Some N-S sway, failure of top
12 119 120 120
spigots and corner standards
Some E-W sway, final failure of
13 70 70 70
perimeter spigots in 2nd lift
Some N-S sway, failure of top
14 40 80 160
spigots and centre standards
Failure of corner spigot and top
15 105 105 105
standard

16 100 100 100 N-S sway mode, failure of jacks

Bearer broke off before final failure


17 140 140 140
of top jacks
Some N-S sway, final failure of
18 150 150 150
corner standard at top lift
Chapter 5 - Formwork Subassembly Tests 95
______________________________________________________________________

The test results suggest that the failure modes are controlled by the jack extension
length since when 600 mm top and bottom extensions are used the failure mode is
North-South sway with final failure at the jacks. On the contrary, when 300 mm
extensions are used, failure occurs mainly in the standards and spigots with only small
sway displacements. Noticeably, the ultimate load decreases as the jack extension
increases. The test results show that the bracing arrangement significantly influences the
ultimate load of the system. Also, the higher lift height reduces the ultimate load.
Moreover, the standards tend to fail at the top lift and around the perimeter region,
especially at the corner where there is no bracing and only two ledgers are connected.
Final failure occurred in spigots and jacks in most cases, as shown in Figure 5.5. Initial
geometric imperfection data used for model calibrations (Chapter 6) is presented in
Appendix C.1. Complete data on displacements as well as supplementary tests on
components of Cuplok scaffold systems, including tensile coupon, stub column and
bending tests are available in CASE (2006).

(a) (b)

Figure 5.5: Failure in (a) spigot and (b) jack (CASE 2006)
Chapter 5 - Formwork Subassembly Tests 96
______________________________________________________________________

5.5 Conclusions
This chapter presents the summary of data, procedure and results pertaining to the
formwork subassembly tests performed at the University of Sydney as part of a
consultancy project for Boral Formwork and Scaffolding Pty. Ltd. The tests consist of
18 full-scale, three-by-three bay Cuplok scaffold systems with combinations of various
lift heights, number of lifts and jack extensions. The configuration, initial geometric
imperfection and material data from these tests are used in Chapter 6 for finite element
modelling of support scaffold systems. Also, the ultimate load and displacement results
are used in the calibrations of the advanced analysis models in Chapter 6.
Chapter 6 - Advanced Structural Analysis Models of Support Scaffold Systems 97
______________________________________________________________________

CHAPTER 6 - ADVANCED STRUCTURAL


ANALYSIS MODELS OF SUPPORT SCAFFOLD
SYSTEMS

6.1 Introduction

Scaffolds are temporary structures commonly used in construction to support various


types of loads. The vertical loads on scaffold can be from labourers, construction
equipment, formworks, and construction materials. Commonly, scaffolds must also be
designed to withstand lateral loads, including wind loads, impact loads, and earthquake
loads. Depending on their use, scaffolds may be categorised as access scaffolds or
support scaffolds. Access scaffolds are used to support light to moderate loads from
labourers, small construction material and equipment for safe working space. They are
usually attached to buildings with ties and only one bay wide. Support scaffolds, also
sometimes called falsework, are subjected to heavy loads, for example, the weight of
wet concrete poured onto the formwork. Support scaffolds are the main focus in this
research. An example of a support scaffold system is shown in Figure 6.1. Support
scaffolds normally consist of standards (vertical members), ledgers (horizontal
members), and braces (diagonal members). The scaffold standards are connected to
each other to create a lift via couplers, also known as spigot joints (Figure 6.2). In order
to connect ledgers to standards, wedge-type or Cuplok joints (Figure 6.3) are usually
preferred because no bolting or welding is required; though, in some systems, manually
adjusted pin-jointed couplers are still being used, which require bolting. The
connections for diagonal brace members are usually made of hooks for easy assembling.
The base of scaffolds consists of threaded adjustable jacks, which can be extended up to
typically 600 mm by a wing nut to accommodate irregularity of the ground. The top of
scaffolds consists of threaded adjustable jacks with U-heads which support timber
bearers and enable the levelling of the formwork.
Chapter 6 - Advanced Structural Analysis Models of Support Scaffold Systems 98
______________________________________________________________________

Figure 6.1: Typical support scaffold

Top Standard

do
t
Spigot Insert
ls

Bottom Standard

Figure 6.2: Schematic of spigot joint


Chapter 6 - Advanced Structural Analysis Models of Support Scaffold Systems 99
______________________________________________________________________

Locking pin

2
Top cup
Ledger blade

Bottom cup
Ledger
Standard

Figure 6.3: Schematic of Cuplok joint

6.1.1 Advanced Analysis

With the ready availability of powerful computers and sophisticated structural analysis
software packages, geometric and material nonlinear structural analysis has become
feasible and practical. Nonlinear analysis allows researchers and practitioners to more
accurately predict the failure load and deformation of scaffold systems. Advanced
analysis involves the modelling of changes of the geometry of structures as a result of
loading and inelastic material behaviour. In the research by Gylltoft and Mroz (1995), a
three-dimensional geometric and material nonlinear finite element model was verified
against the results of a full scale test scaffold. The model was further applied to
determine the ultimate load of a typical access scaffold considering various
configurations and load combinations.

However, in many cases research on scaffold systems has focused on elastic nonlinear
geometric modelling associated with second-order effects. For example, elastic
geometric nonlinear analyses were reported by Peng et al. (2007), Prabhakaran et al.
(2006), Yu et al. (2004), Chu et al. (2002), and Weesner and Jones (2001). Geometric
Chapter 6 - Advanced Structural Analysis Models of Support Scaffold Systems 100
______________________________________________________________________

nonlinear analysis is also a common practice in design offices, whereas the use of
inelastic analysis is still rare. Nevertheless, with accurate finite element model,
advanced analysis method can usually fulfil the design requirement with no tedious
separate member capacity checks.

AS 4100 of the Standard Australia (1998) allows the application of advanced analysis
for the design of steel frames in which the members are of compact cross-section with
full lateral restraint, thus preventing local buckling and flexural-torsional buckling.
Advanced analysis models should include related material properties, residual stresses,
instability effects, initial geometric imperfections, actual connection behaviour,
construction methods, and interaction with the foundations. In the research described
herein, three-dimensional advanced structural analysis models are proposed in order to
develop a new design methodology for support scaffold systems.

6.1.2 Previous Scaffold Models

By means of available commercial finite element softwares such as ANSYS (1998),


LUSAS (1998), and NAF-NIDA (2001), recent studies on scaffold behaviour have been
carried out using three-dimensional models such as those presented by Prabhakaran et al.
(2006), Milojkovic et al. (2002), and Godley and Beale (1997). Three-dimensional
structural analysis is beneficial in describing complex failure modes such as those
observed when the combined effects of in-plane and out-of-plane bending are present.
Some past models proposed by Huang et al. (2000a), and Peng et al. (1997) were
created in two dimensions for simplicity and computational efficiency.

Two types of geometrical imperfections are typically required to be considered in an


advanced analysis of steel framed systems to capture the second-order effects: the initial
member out-of-straightness of the standard and the initial story out-of-plumb of the
frame. There are many ways of taking geometric imperfection effects into account.
Three methods of modelling imperfections were trialled in Chan et al. (2005), including
the scaling of eigenbuckling modes (EBM), the application of notional horizontal forces
(NHF), and the direct modelling of initial geometric imperfections (IGI). EBM was
Chapter 6 - Advanced Structural Analysis Models of Support Scaffold Systems 101
______________________________________________________________________

performed by carrying out an eigenvalue (buckling) analysis of the structural model,


and then scaling and superimposing the lowest eigenmode onto the perfect geometry to
create an initially imperfect structural frame for the second-order structural analysis. In
the NHF approach, additional lateral point loads were applied at the top of each column
in one direction of the frame and initial member out-of-straightness could be
represented by distributed lateral forces along each member. The IGI method consisted
of applying an initial sway of the frame and an out-of-straightness to each column in the
frame. All three methods were found to give similar results in the advanced analysis
when applied in the modelling.

For scaffold systems, these same approaches can be applied to model the effects of
initial imperfections in the analysis. For example, Yu et al. (2004) and Chu et al. (2002)
integrated EBM with the magnitude of the column out-of-straightness of 0.001 of the
height of the scaffold units into the model. Moreover, Yu and Chung (2004)
investigated a method called critical load approach where initial imperfections were
integrated directly into a Perry-Robertson interaction formula to determine the failure
loads of the scaffolds in the analysis. In other research on scaffold systems by Peng et al.
(2007), the NHF approach was incorporated in the model by applying a horizontal
notional force of 0.1% to 0.5% of the vertical loads at mid-height of the scaffold lift.
Godley and Beale (2001) adopted an IGI approach by imposing a sinusoidal bow to the
members and angular out-of-plumb to the frame. In each of these approaches, careful
calibration against test results or numerical reference values is required.

Scaffold joints are complex in nature due to need for rapid assembly and reassembly in
construction. The Cuplok connections behave as semi-rigid joints, and show looseness
with small rotational stiffness at the beginning of loading. Once the joints lock into
place under applied load, the joints become stiffer. Wedge-type joints are generally
more flexible and closer to pinned connections. They also often display substantial
looseness at small rotations. Figure 6.4 shows typical moment-rotation curves for
Cuplok and wedge-type joints. As to spigot joints, the spigot can create out-of-
straightness of the standards, and the possibility of the joint to open up due to the gap
between the standard and the spigot can produce complexity in modelling, as mentioned
by Enright et al. (2000). Moreover, the modelling of boundary conditions of scaffold
Chapter 6 - Advanced Structural Analysis Models of Support Scaffold Systems 102
______________________________________________________________________

systems is crucial because the top and bottom restraints can significantly influence the
stability and strength of the system, as found by Yu (2004).

In recent research by Peng et al. (2007), analysis models of wedge-type jointed, 3-storey,
3-bay, and 5-row scaffold system were presented. Experimental tests on scaffold joints
showed that the joint stiffness varied between 4.903 kNm/rad (50 ton cm/rad) and 8.826
kNm/rad (90 ton cm/rad) with the average value of 6.865 kNm/rad (70 ton cm/rad)
being adopted for all joints into their model.

Godley and Beale (2001) found that scaffold connections are frequently made of wedge-
type joints, for which the joint stiffness exhibits different responses under clockwise and
counter-clockwise rotations, and occasionally exhibits looseness in connections with
low stiffness. Consequently, Prabhakaran et al. (2006) modified the stiffness matrix for
the end points of the beam to include connection flexibility, using a piecewise linear
curve to model the moment-rotation response.

3.5

3.0
Cuplok joint
2.5 Wedge-type joint
M oment (kNm)

2.0

1.5

1.0

0.5

0.0
0.00 0.05 0.10 0.15 0.20
Rotation (radian)

Figure 6.4: Typical moment-rotation curves for Cuplok and wedge-type joints by
Godley and Beale (1997) and Godley and Beale (2001)
Chapter 6 - Advanced Structural Analysis Models of Support Scaffold Systems 103
______________________________________________________________________

Yu (2004) studied the effects of boundary conditions of scaffold systems, and


categorised them into four cases, i.e. Pinned-Fixed, Pinned-Pinned, Free-Fixed, and
Free-Pinned, with the first term being the translational restraint at the top of the scaffold,
and the second term being the rotational restraint at the base of the scaffold. In all
analyses, the rotation at the top was assumed to be free. These conditions were
incorporated into the models of one bay of one-storey modular steel scaffolds (MSS1),
and two-storey modular steel scaffolds (MSS2). Yu found that for MSS1 the failure
loads for Free-Fixed and Pinned-Pinned conditions are reasonably close to test results;
however, for MSS2 the model results are considerably higher than the test results.
Subsequently, Yu suggested that since the top of the scaffolds normally has lateral
restraints then joints at the top can be modelled as translational springs, and for the
bottom rotational spring can be applied. A stiffness of 100 kN/m for the top
translational spring and stiffness of 100 kNm/rad for the bottom rotational spring gave
comparable results to the tests.

Chu et al. (2002) studied single storey double bay scaffolds. In the presence of restraints
on the loading beam and the jack bases, the top and base were modelled with various
boundary conditions, and the scaffold connections were assumed to be rigid. The
researchers found that both Pinned-Pinned and Pinned-Fixed conditions gave higher
load carrying capacities than the experimental results; on the other hand, the Free-Fixed
condition gave satisfactory results compared to the tests. Research on the stability of
single storey scaffold systems by Vaux et al. (2002) found that when Cuplok
connections are represented by pin joints, and the connections of the top and bottom
jacks to the standards are assumed as rigid with the top-bottom boundary conditions
taken as Pinned-Pinned, good agreement can be achieved between numerical and
experimental failure loads.

Weesner and Jones (2001) studied the load carrying capacity of three-storey scaffolds
assuming rigid joints between the stories, and pin joints for the top and the bottom
boundary conditions. The results of their elastic buckling analysis were higher than the
test values with the percentage differences ranging from 6% to 17%. In the analysis of
large access scaffold systems by Godley and Beale (2001), cantilever tests were carried
out on scaffold wedge-type joints. The nonlinear moment-rotation curve obtained from
Chapter 6 - Advanced Structural Analysis Models of Support Scaffold Systems 104
______________________________________________________________________

the tests showed joint looseness and different values of rotational stiffness under
positive (counter-clockwise) rotation and negative (clockwise) rotation. The authors
suggested the use of a multi-linear or nonlinear moment-rotation curve for scaffold joint
modelling.

In the work by Enright et al. (2000), the modelling of spigot joints was studied for the
stability analysis of scaffold systems. The spigot insert (Figure 6.2) was considered to
have bending resistance, but not to transmit axial load; therefore, the model adopted two
vertical members connected by pin joints representing the standards, and on the side, the
entirely rigid spigot member was connected at the top, centre, and bottom to the
standard via short and axially stiff members capable of transferring only lateral forces,
as shown in Figure 6.5. Due to the axial load in the standards, the spigot would be in
bending, and the amount of bending would depend on the amount of axial load and the
degree of out of straightness. From research of Harung et al. (1975), it was found that if
the spigot joints are modelled as fully continuous joints, the analysis would
overestimate the load carrying capacity of the system.

Axial load

Pin joint

Standards
Spigot

Figure 6.5: Spigot joint model by Enright et al. (2000)


Chapter 6 - Advanced Structural Analysis Models of Support Scaffold Systems 105
______________________________________________________________________

Milojkovic B. et al. (1996) studied eccentricity in the modelling of scaffold connections.


Given that the neutral axes of the connections were offset by 50 mm, the authors
modelled the eccentric joint with a finite spring of length equal to the eccentricity of 50
mm. The spring had specific rotational stiffness, and was assumed to be axially stiff.
The authors concluded that for large frames, unless torsion failure occurs, then the
effects of joint eccentricity are insignificant. In the scaffold study by Gylltoft and Mroz
(1995), the braces were represented as truss members with pinned joints connected to
the standards, and the connections between other members were modelled as short finite
elements with nonlinear stiffness in all directions.

6.2 Finite Element Models

In this research, three-dimensional finite element models developed for analysing


support scaffold systems include geometric and material nonlinearities and are
performed using the commercial finite element software package Strand7 (2009). The
models present efficient and accurate methods for representing spigot joints, standard-
to-ledger connections, base plate eccentricities, and load eccentricities. Also, initial
geometric imperfections and material nonlinearity of all components of the system are
incorporated in the models. In modelling the structural elements of the scaffold system,
nonlinear beam elements, including contact, link, and connection elements, are used as
described in Sections 6.2.1 to 6.2.7. The models are compared with the subassembly
tests (CASE 2006) and calibrated against the ultimate loads and displacement responses
in Section 6.2.8.

6.2.1 Spigot Joints

In the studied systems, the spigot joint consists of an insert made from a circular hollow
steel tube with 38.2 mm outside diameter (do), 3.2 mm thickness (t), and 300 mm in
total length (ls), (see Figure 6.2). The insert feeds into the abutting top and bottom
standards to create a required lift connection, as shown in Figure 6.2. The top standard
Chapter 6 - Advanced Structural Analysis Models of Support Scaffold Systems 106
______________________________________________________________________

can slide over the insert, which is fastened to the bottom standard by a fixed pin. The
spigot modelling suggested by Enright et al. (2000) is adopted, as shown in Figure 6.6.

Load

Pinned link

Top Standard

150 mm Pinned link

Spigot
Bottom Standard
150 mm Pinned link

Figure 6.6: Schematic of spigot joint model (adapted from Enright et al. 2000)

The top and bottom standards are modelled as nonlinear beam element connected to the
spigot via pinned connections. The spigot beam element is connected to the standards
by three pinned stiff links capable of only transferring lateral forces from the standards
to the spigot. As a result, the vertical force travels through the standards and only
horizontal forces transfer to the spigot via the pinned links. When the standard bends
under vertical load, the spigot is forced to bend because of the lateral forces acting
oppositely at the top/bottom and centre of the spigot. The degree of bending of the
spigot depends on the amount of initial geometric imperfection of the standard and
vertical force. In three-dimensional analyses, the spigot model is arranged in the
direction perpendicular to the primary bearers which is in the same direction as that of
the loading eccentricity. This arrangement is reasonable since it was observed from the
subassembly tests by CASE (2006) that the spigot joint tends to fail in the same
direction as the loading eccentricity. For simplicity, the spigot model is applied at mid
Chapter 6 - Advanced Structural Analysis Models of Support Scaffold Systems 107
______________________________________________________________________

height of the lift, even though the spigot is often located at little below or above mid
height in actuality.

6.2.2 Semi-Rigid Standard-to-Ledger Connections

The connections between standard and ledger in this research consist of a semi-rigid
Cuplok-type joint that can join up to four ledgers to the standard. The relation between
the moment and rotation of the Cuplok connections is modelled by a tri-linear curve, as
illustrated in Figure 6.7. The parameters that describe the tri-linear curve (k1, k2, k3, 1,
2, 3) were obtained from laboratory tests, as presented in Chapter 4. The slope k1
represents the looseness of the joint, while the slopes k2 and k3 represent elastic and
plastic joint stiffness, respectively. Three different joint configurations were tested in
bending about vertical and horizontal axes, i.e. the 4-way, 3-way and 2-way
configurations, reflecting the number of ledgers connected at the joint. It was observed
that the more ledgers connected, the less movement in the joint itself, and hence the
greater stiffness.

The average joint stiffness values for k1, k2, and k3 for different joint configurations and
bending axes (Figure 6.8) are presented in Table 6.1. The average joint rotation values
for 1, 2 and 3 are presented for different joint configurations and bending axes in
Table 6.2. The connection element in Strand7 (2009) is used to model the relation
between moment and rotation. It requires that a multi-linear moment-rotation table is
specified for bending about vertical and horizontal axes. The connection element is used
to supply stiffness for any of the six degrees of freedom (axial, shear in 2 directions, and
bending about 3 axes). In the proposed model, only bending stiffness in the vertical and
horizontal planes is incorporated and the remaining degrees of freedom are assumed to
provide rigid connections between standard and ledger.
Chapter 6 - Advanced Structural Analysis Models of Support Scaffold Systems 108
______________________________________________________________________

Table 6.1: Average Cuplok joint stiffness (kNm/rad)

Bending about horizontal axis Bending about vertical axis

Joint configuration k1 k2 k3 k1 k2 k3

4-way 80 102 5.3 15 7.5 0.8

3-way 75 87 5.1 14 7 1

2-way 70 77 4.6 7.5 5 1.5

Table 6.2: Average rotation for Cuplok joints (rad)

Bending about horizontal axis Bending about vertical axis

Joint configuration 1 2 3 1 2 3

4-way 0.014 0.036 0.16 0.02 0.04 0.1

3-way 0.012 0.036 0.16 0.02 0.04 0.1

2-way 0.007 0.036 0.16 0.02 0.04 0.1

Moment

k3

k2

k1
Rotation
1 2 3

Figure 6.7: Tri-linear moment-rotation for the Cuplok joints


Chapter 6 - Advanced Structural Analysis Models of Support Scaffold Systems 109
______________________________________________________________________

Figure 6.8: Bending axes of the Cuplok joints

6.2.3 Brace Connections

The braces are made of telescopic members with hooks at the ends. They are modelled
using two rigidly connected elements with different cross-sections. Connection
elements with only axial stiffness form the connection between the brace member
elements and the ledgers. The axial spring stiffness is taken as 1.8 kN/mm as obtained
from test calibrations on braced scaffold systems. The braces are offset 60 mm along the
ledger from the nodal points between the standards and ledgers because in actual
construction the braces are connected to ledgers at about this distance away from the
joints.

6.2.4 Base Plate Eccentricity

The placement of the base plate of scaffold systems on an uneven or sloped ground can
create eccentricity. The amount of base eccentricity depends largely on ground surface
irregularities as shown in Figure 6.9(a). AS 3610 by Standard Australia (1995) specifies
an expected base eccentricity of no more than 40 mm or bp/4, whichever is less, where
bp is the stiff portion of bearing of an end plate, as shown in Figure 6.9(a). For example,
Chapter 6 - Advanced Structural Analysis Models of Support Scaffold Systems 110
______________________________________________________________________

bp/4 is 17 mm for the scaffold system in the study which is less than 40 mm; therefore,
the expected base eccentricity of the system is no more than 17 mm. The full-scale
subassembly tests described in Chapter 5 featured a base eccentricity of 15 mm along
the second row in the East-West direction of the frame, and so this value was used in
the calibrations of numerical model. The base eccentricity model proposed is illustrated
in Figure 6.9(b), in which the base eccentricity is labelled as e. The standard and the
base plate are modelled using nonlinear beam elements with their corresponding cross-
sectional and material properties. A contact element is used to model a gap between the
base plate and the ground. The contact element is set to provide stiffness only in
compression, and only when the nodes to which it is connected come into contact, that
is when the gap closes. The stiffness is specified as infinity, implying that when the
load transfers from the standard to the base plate causing the far end of the base plate to
rotate and touch the ground, the point of contact becomes infinitely stiff representing
solid ground or other hard surface.

Load
Standard
Standard
Base plate
Base plate e
1 Contact element
1 Uneven ground

bp

(a) (b)

Figure 6.9: (a) base plate on uneven ground and (b) base eccentricity model

6.2.5 Loading Eccentricity

A loading eccentricity can occur between the timber bearer and the U-Head since the
bearer is not always positioned such that its centre line coincides with the centre line of
the jack, even though in good construction practice the U-head is twisted against the
bearer so as to reduce the amount of eccentricity. In order to model the eccentricity, a
Chapter 6 - Advanced Structural Analysis Models of Support Scaffold Systems 111
______________________________________________________________________

rigid link with length equal to the loading eccentricity is connected to the top of the jack
in the direction perpendicular to the bearer, and a vertical point load is applied at the far
end of the link. The rigid link behaves as a short, stiff cantilever that introduces vertical
force and additional moment into the jack. The full-scale subassembly tests described in
Chapter 5 featured a top loading eccentricity of 25 mm along the second row in the
East-West direction of the frame, and so this value was used in the calibrations of
numerical model.

6.2.6 Geometric Imperfections

Scaffold systems are generally slender and sensitive to stability effects; therefore, initial
geometric imperfections producing member P- and frame P- effects (Figure 6.10)
must be considered in the analysis model. A common approach to incorporate geometric
imperfections is to scale one or more critical elastic buckling modes and apply the
scaled displacements to the perfect geometry. Nevertheless, several issues remain
unanswered in this method. For instance, how many buckling modes should be included
and what scaling factors should be applied? One alternative is to use the maximum
allowable imperfection values from available structural codes. For example, the
Australian steel structures design standard by Standard Australia (1998) specifies a
maximum out-of-straightness of L/1000, where L is the member length. Nonetheless,
this method often produces conservative predictions of the strength of systems. Another
approach is to apply notional horizontal forces, but some doubt remains as to the
magnitude of force to use.

Alternatively, for the purpose of calibrating the analysis model, in this research the
magnitudes of the member out-of-straightness and the frame out-of-plumb are
implemented directly at the nodes of the finite element models from initial imperfection
measurements taken as part of the subassembly tests. The member out-of-straightness
() is applied at mid height of the standard in each scaffold lift and the frame out-of-
plumb () is applied at each ledger-standard connection point and at the U-head at the
top of the scaffold. From the measurements by CASE (2006), the average out-of-
straightness of the standards is Lh/820 mm for the lifts with spigot joint and Lh/1700
Chapter 6 - Advanced Structural Analysis Models of Support Scaffold Systems 112
______________________________________________________________________

mm for the lifts without spigot joint, where Lh is the lift height, and the average out-of-
plumb of the frames is H/470 mm, where H is the total height. Full details of the initial
geometric imperfections of the subassembly tests are available in Appendix C.1. For
other studies on the scaffold systems, real data on initial geometric imperfections is
procured from construction sites around the Sydney area and presented in Chapter 3.

Figure 6.10: P- and P- effects

6.2.7 Geometric and Material Nonlinearities

In geometric nonlinear analysis, an accurate determination of the displacements can be


achieved which is particularly important in slender structures such as scaffold systems.
In the present beam element based analysis, the deformed geometry is used to establish
the equilibrium equations and the elements local reference system is updated at each
load increment to capture the load deflection characteristics.

In material inelastic analysis, the nonlinear relationship between stress and strain is
applied. The stress-strain relations for the scaffold components used in the models are
based on the Ramberg-Osgood expression by Ramberg and Osgood (1941) and
Rasmussen (2003) fitted to experimental data obtained from supplementary material
tests on components of the subassembly tests by CASE (2006). Figures 6.11 to 6.16
show the material test results and the Ramberg-Osgood stress-strain curves used in the
Chapter 6 - Advanced Structural Analysis Models of Support Scaffold Systems 113
______________________________________________________________________

material modelling of the experimental standard, ledger, jack, base plate, brace, and
spigot respectively. Since there is no experimental stress-strain data for the ledger, brace
and spigot, the stress-strain relations for these components are obtained by scaling the
Ramberg-Osgood stress-strain relation used for the standards to their nominal yield
stress. However, the ledger, brace and spigot insert are expected to be loaded only in
elastic range, and therefore, the material nonlinearity effects of these components are
negligible. The effects of residual stresses are only implicitly considered through the
Ramberg-Osgood curves. The Ramberg-Osgood parameters (E0, 0.2, n) for each
scaffold component are summarised in Table 6.3. In the table, E0 is the initial Youngs
modulus, 0.2 is the 0.2% proof stress (also referred to as the equivalent yield stress),
and n is a parameter which determines the sharpness of the knee of the stress-strain
curve.

The Ramberg-Osgood stress-strain relations are applied to the beam elements of each
scaffold component in finite element software package, Strand7 (2009). As the beam
cross-section and length are subdivided, sampling points in the cross-section and
integration points along the length of the beam are utilised to numerically integrate the
stiffness characteristics of the beam. As a result of this section and length-wise
integration, the propagation of yielding through the cross-section and along the beam
element can be included. The axial and bending stiffness are coupled as the neutral axis
on the yielded beam shifts. This method of analysis is referred to as plastic-zone
analysis by Clarke et al. (1992).

Table 6.3: Ramberg-Osgood parameters for scaffold components

Component E0 (GPa) 0.2 (MPa) n


Standard 200 530 38.2
Ledger 200 380 38.2
Jack 200 495 16.0
Base plate 200 260 25.0
Brace 200 430 38.2
Spigot 200 430 38.2
Chapter 6 - Advanced Structural Analysis Models of Support Scaffold Systems 114
______________________________________________________________________

6.00E+08

5.00E+08

4.00E+08
Stress (Pa)

3.00E+08

2.00E+08

1.00E+08

0.00E+00
0.0000 0.0050 0.0100 0.0150 0.0200 0.0250

Strain

Ramberg-Osgood Test result

Figure 6.11: Stress-strain curve for standard

5.00E+08

4.00E+08
Stress (Pa)

3.00E+08

2.00E+08

1.00E+08

0.00E+00
0.0000 0.0050 0.0100 0.0150 0.0200 0.0250

Strain

Ramberg-Osgood

Figure 6.12: Stress-strain curve for ledger


Chapter 6 - Advanced Structural Analysis Models of Support Scaffold Systems 115
______________________________________________________________________

7.00E+08

6.00E+08

5.00E+08
Stress (Pa)

4.00E+08

3.00E+08

2.00E+08

1.00E+08

0.00E+00
0.0000 0.0050 0.0100 0.0150 0.0200 0.0250

Strain

Ramberg-Osgood Test result

Figure 6.13: Stress-strain curve for jack

4.00E+08

3.00E+08
Stress (Pa)

2.00E+08

1.00E+08

0.00E+00
0.0000 0.0050 0.0100 0.0150 0.0200 0.0250

Strain

Ramberg-Osgood Test result

Figure 6.14: Stress-strain curve for base plate


Chapter 6 - Advanced Structural Analysis Models of Support Scaffold Systems 116
______________________________________________________________________

5.00E+08

Stress (Pa) 4.00E+08

3.00E+08

2.00E+08

1.00E+08

0.00E+00
0.0000 0.0050 0.0100 0.0150 0.0200 0.0250

Strain

Ramberg-Osgood

Figure 6.15: Stress-strain curve for brace

5.00E+08

4.00E+08
Stress (Pa)

3.00E+08

2.00E+08

1.00E+08

0.00E+00
0.0000 0.0050 0.0100 0.0150 0.0200 0.0250

Strain

Ramberg-Osgood

Figure 6.16: Stress-strain curve for spigot


Chapter 6 - Advanced Structural Analysis Models of Support Scaffold Systems 117
______________________________________________________________________

6.2.8 Calibrations

The commercial software package Strand7 (2009) was used to create a finite element
model for each of the full-scale subassembly tests using the actual frame dimensions
and measured values of imperfections. The mean of the measured dimensions of
components were used for cross-sectional properties in the finite element models. As an
example, Figure 6.17 shows the finite element model for Test No. 3 of the subassembly
tests, referred to in Chapter 5. The ultimate loads and displacements obtained from the
nonlinear analyses accounting for both material and geometric nonlinearities were
calibrated against failure loads and load-deflection responses obtained from the full-
scale subassembly tests, referred to in Chapter 5. The calibrations were achieved by
changing the stiffness of the elastic restraints applied at the U-head and base plate, as
well as the axial spring stiffness of the brace connections; the latter was changed after
the calibrations were performed on unbraced systems for the top and bottom rotational
stiffness. These stiffnesses were inherent in the loading rig and could not be measured
directly as part of the test program. However, by suitable choice of the stiffness values,
a consistently good agreement could be achieved between the numerical predictions and
experimental measurements for a wide range of bracing, loading and system
combinations. Table 6.4 shows the results of the stiffness parameters obtained from the
calibrations. In the table, K represents translational stiffness with subscript showing its
direction corresponding to Figure 6.17, and R represents rotational stiffness with
subscript showing the axis of bending according to Figure 6.17.

It should be noted that the top rotational stiffness about the x-axis is assumed to be rigid,
corresponding to the negligible strong axis bending of the supported timber bearer.
However, the y-axis bending stiffness is taken as 40 kNm/rad since bending about this
axis occurs during failure as observed in the tests CASE (2006). The bottom rotational
stiffness about the x and y axes is calibrated as 100 kNm/rad. The bottom rotational
stiffness is applied to all uprights except the uprights with bottom eccentricity for which
base plate modelling is applied. The translational stiffness at the base is taken as rigid in
all directions. At the top, the translational stiffness is assumed to be rigid in the x and y
Chapter 6 - Advanced Structural Analysis Models of Support Scaffold Systems 118
______________________________________________________________________

directions (since the formwork is usually tied to a permanent structure), but 0 in the
vertical z direction, based on the axes shown in Figure 6.17. The brace hook end
connections have an axial stiffness of 1.8 kN/mm capable of transferring only axial
forces to the ledgers. Table 6.5 shows the calibration results for the failure loads and
their statistics. The average of the ratio of test failure load to ultimate load from
advanced analysis is 1.014 with a standard deviation (STD) and a coefficient of
variation (COV) of 0.0980 and 0.0966, respectively. The ratios of test capacity and
analytical prediction for the fifteen frames are shown graphically in Figure 6.18. Figures
6.19 to 6.33 compare the finite element analysis results with the experimental load-
deflection responses at certain points of the frame, as indicated for each test in the titles
of the figures.

Table 6.4: Parametric calibration results

Bottom boundary conditions


Kx (kN/mm) Ky (kN/mm) Kz (kN/mm) Rx (kNmm/rad) Ry (kNmm/rad) Rz (kNmm/rad)
Rigid Rigid Rigid 100,000 100,000 0

Top boundary conditions


Kx (kN/mm) Ky (kN/mm) Kz (kN/mm) Rx (kNmm/rad) Ry (kNmm/rad) Rz (kNmm/rad)
Rigid Rigid 0 Rigid 40,000 0

Brace end connections


Axial stiffness (kN/mm) 1.8
Chapter 6 - Advanced Structural Analysis Models of Support Scaffold Systems 119
______________________________________________________________________

Figure 6.17: Finite element model of Test No. 3 showing axes

Table 6.5: Load calibration results

Ultimate load
from Test failure
Test Test load / Advanced analysis result
advanced load (kN)
analysis (kN)
2 96 89 0.927
3 91 91 1.000
4 45 50 1.111
5 60 60 1.000
6 66 60 0.909
8 138 130 0.942
9 50 65 1.300
10 64 70 1.094
11 127 120 0.945
12 129 120 0.930
13 68 70 1.029
14 160 160 1.000
15 105 105 1.000
16 100 100 1.000
18 147 150 1.020
Average 1.014
STD 0.0980
COV 0.0966
Chapter 6 - Advanced Structural Analysis Models of Support Scaffold Systems 120
______________________________________________________________________

1.3

Test Capacity/Prediction
1.2

1.1

1.0

0.9

0.8

0.7
0 3 6 9 12 15

Test

Figure 6.18: Test capacity/Prediction of fifteen support scaffold systems

120

100

80
Load (kN)

60

40

20

0
0 5 10 15 20 25 30 35 40
Deflection (mm)

Test result FE result

Figure 6.19: Calibration of load-deflection responses at mid-height of the standard in


the 3rd lift of the 2nd row of the frame for Test No. 2
Chapter 6 - Advanced Structural Analysis Models of Support Scaffold Systems 121
______________________________________________________________________

100
90
80
70
60
Load (kN)

50
40
30
20
10
0
0 5 10 15 20 25 30 35 40
Deflection (mm)

Test result FE result

Figure 6.20: Calibration of load-deflection responses at mid-height of the standard in


the 3rd lift of the 2nd row of the frame for Test No. 3

50
45
40
35
30
Load (kN)

25
20
15
10
5
0
0 5 10 15 20 25 30
Deflection (mm)

Test result FE result

Figure 6.21: Calibration of load-deflection responses at mid-height of the standard in


the 3rd lift of the 2nd row of the frame for Test No. 4
Chapter 6 - Advanced Structural Analysis Models of Support Scaffold Systems 122
______________________________________________________________________

70

60

50
Load (kN)

40

30

20

10

0
0 5 10 15 20 25 30 35
Deflection (mm)

Test result FE result

Figure 6.22: Calibration of load-deflection responses at mid-height of the standard in


the 3rd lift of the 2nd row of the frame for Test No. 5

70

60

50
Load (kN)

40

30

20

10

0
0 2 4 6 8 10 12
Deflection (mm)

Test result FE result

Figure 6.23: Calibration of load-deflection responses at mid-height of the standard in


the 3rd lift of the 2nd row of the frame for Test No. 6
Chapter 6 - Advanced Structural Analysis Models of Support Scaffold Systems 123
______________________________________________________________________

160

140

120

100
Load (kN)

80

60

40

20

0
0 5 10 15 20 25 30
Deflection (mm)

Test result FE result

Figure 6.24: Calibration of load-deflection responses at mid-height of the standard in


the 3rd lift of the 2nd row of the frame for Test No. 8

60

50

40
Load (kN)

30

20

10

0
0 1 2 3 4 5 6 7 8 9
Deflection (mm)

Test result FE result

Figure 6.25: Calibration of load-deflection responses at mid-height of the standard in


the 3rd lift of the 2nd row of the frame for Test No. 9
Chapter 6 - Advanced Structural Analysis Models of Support Scaffold Systems 124
______________________________________________________________________

70

60

50
Load (kN)

40

30

20

10

0
0 0.5 1 1.5 2 2.5 3
Deflection (mm)

Test result FE result

Figure 6.26: Calibration of load-deflection responses at mid-height of the standard of


the 3rd lift of the 2nd row of the frame for Test No. 10

140

120

100
Load (kN)

80

60

40

20

0
0 2 4 6 8 10 12 14 16
Deflection (mm)

Test result FE result

Figure 6.27: Calibration of load-deflection responses at mid-height of the standard in


the 3rd lift of the 2nd row of the frame for Test No. 11
Chapter 6 - Advanced Structural Analysis Models of Support Scaffold Systems 125
______________________________________________________________________

140

120

100
Load (kN)

80

60

40

20

0
0 2 4 6 8 10 12 14 16 18
Deflection (mm)

Test result FE result

Figure 6.28: Calibration of load-deflection responses at mid-height of the standard in


the 3rd lift of the 2nd row of the frame for Test No. 12

80

70

60

50
Load (kN)

40

30

20

10

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Deflection (mm)

Test result FE result

Figure 6.29: Calibration of load-deflection responses at the 2nd lift of the 1st row of the
frame for Test No. 13
Chapter 6 - Advanced Structural Analysis Models of Support Scaffold Systems 126
______________________________________________________________________

180
160
140
120
Load (kN)

100
80
60
40
20
0
0 2 4 6 8 10 12 14 16 18
Deflection (mm)

Test result FE result

Figure 6.30: Calibration of load-deflection responses at mid-height of the standard in


the 3rd lift of the 2nd row of the frame for Test No. 14

120

100

80
Load (kN)

60

40

20

0
0 2 4 6 8 10 12 14 16
Deflection (mm)

Test result FE result

Figure 6.31: Calibration of load-deflection responses at mid-height of the standard in


the 2nd lift of the 2nd row of the frame for Test No. 15
Chapter 6 - Advanced Structural Analysis Models of Support Scaffold Systems 127
______________________________________________________________________

120

100

80
Load (kN)

60

40

20

0
0 5 10 15 20 25 30
Deflection (mm)

Test result FE result

Figure 6.32: Calibration of load-deflection responses at mid-height of the standard in


the 3rd lift of the 2nd row of the frame for Test No. 16

160

140

120

100
Load (kN)

80

60

40

20

0
0 5 10 15 20 25 30 35
Deflection (mm)

Test result FE result

Figure 6.33: Calibration of load-deflection responses at mid-height of the standard in


the 3rd lift of the 2nd row of the frame for Test No. 18
Chapter 6 - Advanced Structural Analysis Models of Support Scaffold Systems 128
______________________________________________________________________

6.3 Discussion

The calibrations show that geometric and material (advanced) analysis using
geometric and material nonlinear finite element models gives very good predictions of
the ultimate loads of the systems. Most of the predictions are within 10% of the actual
failure loads. The average of the ratios between failure test load and predicted ultimate
load is very close to 1 (1.014) with a relatively small COV of 0.0966. While the average
value close to unity is a result of the calibration, the relatively small COV demonstrates
that the stiffness values calibrated in Section 6.2.8 are likely to be accurate in that they
provide consistently accurate strength predictions for a wide range of system
configurations.

In addition, advanced analysis gives good results in predicting deformation responses of


support scaffold systems. The finite element analysis results of the load-deflection
responses fit the test results from CASE (2006) reasonably closely with most of the
values within 20% of one another. It can be noticed that in some tests there are no
deflection values available for the failure load. Also, Test No. 4 only provides three
measured deflections during loading, and has been ignored in the comparison.

Two distinct failure modes are observed from the advanced analysis, one exhibiting
mainly an S-shape member buckle (Figure 6.34) and the other mainly a lateral frame
buckle with large lateral displacements at the top story (Figure 6.35). The failure modes
are noticed to be sensitive to the jack extension length, where 600 mm jack extension
produces lateral frame buckling with predominant failure deformations in the jacks and
300 mm jack extension produces S-shape buckling of the standards with predominant
failure deformations in the spigots. These failure modes were also observed in the tests,
as discussed in Section 5.4, suggesting that advanced analysis is capable of accurately
predicting the behaviour and failure mode of support scaffold systems.
Chapter 6 - Advanced Structural Analysis Models of Support Scaffold Systems 129
______________________________________________________________________

Figure 6.34: S-shape member buckling

Figure 6.35: Lateral frame buckling


Chapter 6 - Advanced Structural Analysis Models of Support Scaffold Systems 130
______________________________________________________________________

6.4 Conclusions

In this chapter, nonlinear finite element analysis models for support scaffold systems
have been developed. Models for various components of the systems including spigot
joints, semi-rigid upright-to-beam connections and base plate eccentricities are proposed.
Calibrations of these models to the full-scale subassembly tests (as referred to in
Chapter 5) consisting of three-by-three bay formwork systems with the combinations of
different numbers of lifts, jack extension, and lift height are achieved by adjusting the
top and bottom boundary conditions as well as the brace connection stiffness. The
ultimate loads obtained from advanced analysis are in close agreement with the failure
loads of the tests; moreover, comparisons of load-deflection responses also show close
agreement, demonstrating that advanced analysis is able to accurately predict the
behaviour and strength of highly complex support scaffold systems. Particular attention
is paid to the modelling of the load transfer at base plate, U-head and spigot joints of
support scaffold systems. The development of a design methodology for support
scaffold systems based on advanced analysis is described in Chapter 9.
Chapter 7 - Parametric Studies of Support Scaffold Systems 131
______________________________________________________________________

CHAPTER 7 - PARAMETRIC STUDIES OF


SUPPORT SCAFFOLD SYSTEMS

7.1 Introduction

Many studies have been carried out to determine the parameters influencing the strength
of support scaffold systems. Yu et al. (2004) studied the influence of the number of
storeys on the load carrying capacity of door-type modular scaffolds. They analysed one,
two, and three storey steel scaffolds, and found that two-storey and three-storey
scaffolds had only 85% and 80% of the load carrying capacity of the single storey steel
scaffold, respectively, because the different numbers of storeys presented considerable
variation in buckling behaviour. Also, research on the relationship between the load
carrying capacity and the number of storeys of shoring door-type scaffold system by
Huang et al. (2000c) showed that the ultimate loads of the system from incremental
finite element analysis reduced rapidly from the range of two to eight storeys, followed
by a gradual decrease thereafter.

Furthermore, it was found by Peng et al. (1997) that when the initial imperfection
simulated by 1.50% notional horizontal force approach was applied at mid-height, the
reduction in ultimate load of simple door-type scaffold systems was found to be about
16%, which was conservative compared to experimental results, and the relationship
between initial imperfection and reduction in ultimate load of the scaffold system was
nearly linear. With the notional horizontal force of 0.1% of the vertical load applied to
the model, the predicted ultimate load showed good agreement with test results. In
addition, Peng et al. (1997) found that with long shores installed, the ultimate load of
the scaffold system could be as little as 25% of that of the system without shores. From
the analyses of high clearance steel scaffolds by Peng et al. (1996a), the optimum load
carrying capacity for steel scaffolds with shoring occurred in the range of three to six
storeys. In addition, scaffolds of more than eight storeys were not recommended due to
the high reduction in strength of such systems.
Chapter 7 - Parametric Studies of Support Scaffold Systems 132
______________________________________________________________________

Most of the research in the literature considers modular scaffold systems; therefore, a
stick-type scaffold system, the Cuplok scaffold system, is examined in this thesis. The
main factors that influence the ultimate strength of scaffold systems are loading
eccentricity, jack extensions, lift heights, number of lifts, bracing arrangements, Cuplok
joint stiffness, geometric imperfections, and yield stress and cross-sectional area of the
main loaded components. Investigation of the factors affecting the strength of support
scaffold systems helps to determine which factors should be included and treated as
random variables or deterministic values in the probabilistic assessment of the system
strength, as presented in Chapter 8. In addition, it can help designers to achieve the
optimum design as to which configurations and bracing arrangements to be used.

7.2 Analysis Models for Parametric Studies

Support scaffold systems, particularly Cuplok scaffold systems, are considered in these
parametric studies. The three-dimensional nonlinear finite element models have been
created in Strand7 (2009), using the same modelling concepts as presented in Chapter 6.
In summary of the key features, the models used in the parametric studies include spigot
joints, nonlinear joint stiffness for the standard-to-ledger connections, brace connections
with 1.8 kN/mm axial spring stiffness and 60 mm offset from the nodal points, base
plate eccentricity, loading eccentricity, geometric imperfections implemented directly at
the nodes, and Ramberg-Osgood material model for all of the scaffold components.

The standards, ledgers and bracings are made of steel circular hollow sections (CHS).
The dimensions and yield stress of the standards are 48.3 mm x 4 mm and 450 MPa,
respectively. The ledgers are of dimension 48.3 mm x 3.2 mm with a yield stress of 350
MPa. The bracings are made of telescopic members of CHS 48.3 mm x 4 mm and CHS
38.2 mm x 3.2 mm, grade 400 MPa. The steel base plates are 180 mm x 180 mm x 10
mm in dimensions with nominal yield stress of 250 MPa. The spigot joints are assumed
to be located at the mid-height of the second lift and higher. There are no spigots in the
bottom lift based on general construction practice. The jacks are made of 36 mm
diameter threaded steel rod, grade 430 MPa. These scaffold components are identical to
Chapter 7 - Parametric Studies of Support Scaffold Systems 133
______________________________________________________________________

the ones being considered throughout the thesis. The jack extensions at the top and
bottom of the frame are set to equal.

For the top boundary conditions, the rotational stiffness about the x-axis is assumed to
be rigid, corresponding to the negligible strong axis bending of the supported timber
bearer; however, the y-axis bending stiffness is applied as 40 kNm/rad as obtained from
the model calibrations in Chapter 6. For the bottom boundary conditions, the base plate
model is applied with the eccentricity of 15 mm (maximum allowable base eccentricity
based on Standard Australia 1995) to each upright in the same direction as the top
loading eccentricity. The vertical point loads are applied with equal magnitude on top of
each upright. Figure 7.1 shows a typical finite element model (3x3 bays, 1.0 m lift
height, 3 lifts with 300 mm top and bottom jack extensions) used in the parametric
studies. While varying one parameter, other parameters are assumed constant and
applied to the model at their mean values. These mean values, which are presented in
Chapters 3-5, are summarised in Table 7.1. The ultimate system strength is obtained by
performing advanced structural analyses, taking into account of both geometric and
material nonlinearities. The results of the ultimate system strength for reference cases
evaluated at the mean values of each parameter are presented in Table 7.2 for further
comparisons.

Figure 7.1: Typical finite element model used in the parametric studies
Chapter 7 - Parametric Studies of Support Scaffold Systems 134
______________________________________________________________________

Table 7.1: Mean values for parameters in the parametric studies (refer to Chapters 3-5)

Geometric parameter Mean

Normalised out-of-straightness of
the standards without spigot joints 0.0004
(/Lh)
Normalised out-of-straightness of
the standards with spigot joints 0.0013
(/Lh)

Normalised storey out-of-plumb


0.0016
(/H)

Mean bending stiffness about Mean bending stiffness about


horizontal axis (kNm/rad) vertical axis (kNm/rad)

Joint
k1 k2 k3 k1 k2 k3
configuration

4-way 39 102 5.3 15 7.5 0.8

3-way 36 87 5.1 14 7 1

2-way 41 77 4.6 7.5 5 1.5

Mean rotation about horizontal Mean rotation about vertical axis


axis (rad) (rad)

Joint
1 2 3 1 2 3
configuration

4-way 0.014 0.036 0.16 0.02 0.04 0.1

3-way 0.012 0.036 0.16 0.02 0.04 0.1

2-way 0.007 0.036 0.16 0.02 0.04 0.1


Chapter 7 - Parametric Studies of Support Scaffold Systems 135
______________________________________________________________________

Component E0 (GPa) 0.2 (MPa) n

Standard 200 530 38.2

Ledger 200 380 38.2

Jack 200 495 16.0

Base plate 200 260 25.0

Brace 200 430 38.2

Spigot 200 430 38.2

Table 7.2: Ultimate system strength for reference cases (capacity per upright)

Jack Loading System


Lift height Number of Number of
extensions eccentricity strength
(m) lifts bays
(mm) (mm) (kN)

1.0 100 3 3x3 0 193.7

1.0 100 3 3x3 18 156.25

1.0 100 3 3x3 25 147

1.0 100 3 3x3 40 131

1.0 100 3 3x3 55 119.9

1.0 300 3 3x3 0 137.9

1.0 300 3 3x3 18 124.45

1.0 300 3 3x3 25 121

1.0 300 3 3x3 40 107.45

1.0 300 3 3x3 55 100.05

1.0 600 3 3x3 0 86

1.0 600 3 3x3 18 78.45

1.0 600 3 3x3 25 75.9

1.0 600 3 3x3 40 71.5


Chapter 7 - Parametric Studies of Support Scaffold Systems 136
______________________________________________________________________

Jack Loading System


Lift height Number of Number of
extensions eccentricity strength
(m) lifts bays
(mm) (mm) (kN)

1.0 600 3 3x3 55 67.5

1.5 100 3 3x3 0 146.25

1.5 100 3 3x3 18 125.5

1.5 100 3 3x3 25 119.25

1.5 100 3 3x3 40 107.9

1.5 100 3 3x3 55 98.2

1.5 300 3 3x3 0 115

1.5 300 3 3x3 18 109

1.5 300 3 3x3 25 106.7

1.5 300 3 3x3 40 98.6

1.5 300 3 3x3 55 91.25

1.5 600 3 3x3 0 71

1.5 600 3 3x3 18 69.3

1.5 600 3 3x3 25 69

1.5 600 3 3x3 40 65.9

1.5 600 3 3x3 55 63

2.0 100 3 3x3 0 96.9

2.0 100 3 3x3 18 89.25

2.0 100 3 3x3 25 85

2.0 100 3 3x3 40 81.25

2.0 100 3 3x3 55 76.75

2.0 300 3 3x3 0 85

2.0 300 3 3x3 18 80.2


Chapter 7 - Parametric Studies of Support Scaffold Systems 137
______________________________________________________________________

Jack Loading System


Lift height Number of Number of
extensions eccentricity strength
(m) lifts bays
(mm) (mm) (kN)

2.0 300 3 3x3 25 78.45

2.0 300 3 3x3 40 73.25

2.0 300 3 3x3 55 69.45

2.0 600 3 3x3 0 57

2.0 600 3 3x3 18 55.8

2.0 600 3 3x3 25 55

2.0 600 3 3x3 40 54.3

2.0 600 3 3x3 55 53.5

7.3 Loading Eccentricity

The parametric studies of loading eccentricity for different lift heights and jack
extensions (particularly used in Cuplok scaffold systems) are presented graphically in
Figures 7.2-7.4. The results reveal that with shorter jack extensions, i.e. 100 mm jack
extensions, loading eccentricity leads to higher variation in the system strength. The
differences in the ultimate strength based on maximum (55 mm) and zero loading
eccentricity of the 100 mm jack extensions for the 1.0 m lift, 1.5 m lift and 2.0 m lift are
62%, 49% and 26%, respectively. In addition, the differences in the ultimate strength of
the 300 mm jack extensions for the 1.0 m lift, 1.5 m lift and 2.0 m lift are 38%, 26% and
22%, respectively, and the differences in the ultimate strength of the 600 mm jack
extensions for the 1.0 m lift, 1.5 m lift and 2.0 m lift reduce to 27%, 13% and 7%,
correspondingly.

In terms of the influence of lift height, the 1.0 m lift scaffold systems show higher
ultimate strength than the 2.0 m lift systems of up to 100% with zero loading
eccentricity and 56% with maximum loading eccentricity, and higher ultimate strength
than the 1.5 m lift systems of up to 32% with zero loading eccentricity, but only 7%
Chapter 7 - Parametric Studies of Support Scaffold Systems 138
______________________________________________________________________

with maximum loading eccentricity of 55 mm being considered. From the figures,


higher slopes of the curves in the parametric studies of the 1.0 m lift for a corresponding
jack extension suggest that there can be higher variation in the ultimate system strength
of 1.0 m lift scaffold systems than those of 1.5 m and 2.0 m lift systems.

200

175
Ultimate system strength (kN)

150

125
100 mm jack extensions
100 300 mm jack extensions
600 mm jack extensions
75

50

25

0
0 10 20 30 40 50
Loading eccentricity (mm)

Figure 7.2: Parametric studies of loading eccentricity of 1.0 m lift scaffold systems

200

175
Ultimate system strength (kN)

150

125
100 mm jack extensions
100 300 mm jack extensions
600 mm jack extensions
75

50

25

0
0 10 20 30 40 50
Loading eccentricity (mm)

Figure 7.3: Parametric studies of loading eccentricity of 1.5 m lift scaffold systems
Chapter 7 - Parametric Studies of Support Scaffold Systems 139
______________________________________________________________________

200

175
Ultimate system strength (kN)

150

125
100 mm jack extensions
100 300 mm jack extensions
600 mm jack extensions
75

50

25

0
0 10 20 30 40 50
Loading eccentricity (mm)

Figure 7.4: Parametric studies of loading eccentricity of 2.0 m lift scaffold systems

7.4 Number of Lifts

The parametric studies of the number of lifts (storeys) for different lift heights of
scaffold systems in 3x3 bays are presented in Figure 7.5. The figure shows that the
ultimate system strength reduces significantly as the number of lifts increases from one
to five, regardless of the lift height of the systems, and when there are more than five
lifts the ultimate system strength appears to decrease only slightly. The studies also
show that 2-lift and 3-lift scaffold systems have only 93% and 85% of the ultimate
system strength of the single lift scaffold systems on average, respectively. From the 5-
lift to 9-lift scaffold systems, the ultimate system strength reduces only by
approximately 4%. As a result, the ultimate system strength is mostly affected by a
change in the number of lifts in systems with a relatively small number of lifts. Figure
7.6 shows a typical failure mode of scaffold systems with 9 lifts. For large number of
lifts, the failure of the system is governed by a lateral frame buckling with large
displacements at the top and bottom stories.
Chapter 7 - Parametric Studies of Support Scaffold Systems 140
______________________________________________________________________

180

160

140
Ultimate system strength (kN)

120 1.0 m lift height with 300 mm


jack extensions
100
1.5 m lift height with 300 mm
jack extensions
80
2.0 m lift height with 300 mm
60 jack extensions

40

20

0
1 2 3 4 5 6 7 8 9
Number of lifts

Figure 7.5: Parametric studies of number of lifts of scaffold systems

Figure 7.6: Typical failure mode of scaffold systems with 9 lifts


Chapter 7 - Parametric Studies of Support Scaffold Systems 141
______________________________________________________________________

7.5 Bracing Arrangements

The scaffold systems of 3x3 bays, 3 lifts with different lift heights and 300 mm top and
bottom jack extensions, are chosen for the studies. The bracing configurations
considered in the parametric studies include full bracing, perimeter bracing, core
bracing, and no bracing. Full bracing consists of perimeter and core bracings (see Figure
7.7), while no bracing does not have any braces in the systems. In these parametric
studies, full bracing consists of 24 braces, and both perimeter and core bracings have 12
braces. Table 7.3 shows the ultimate system strengths of systems with different bracing
arrangements for three different lift heights. The last column of the table gives
comparisons to reference cases (from Section 7.2) in percentage difference. The results
show that both perimeter and core bracings provide comparable ultimate system
strengths of around 65%, 53%, and 44% of those of the full bracing for 1.0 m, 1.5 m,
and 2.0 m lift heights, respectively. In addition, systems with no bracing have only
43.5%, 30.4%, and 29.4% of the ultimate system strength of the full bracing for 1.0 m,
1.5 m, and 2.0 m lift heights, respectively. Considering the percentage differences to the
reference cases between the lift heights, it appears that the 2.0 m lift is the most
sensitive to the number of braces.

Figure 7.7: Bracing configurations in the parametric studies


Chapter 7 - Parametric Studies of Support Scaffold Systems 142
______________________________________________________________________

Table 7.3: Parametric studies of bracing configurations of scaffold systems

Jack System
Lift height Number Number Difference
extensions Bracing strength
(m) of lifts of bays (% )
(mm) (kN)

1.0 300 3 3x3 Full 137.9 0.0%

1.0 300 3 3x3 Perimeter 90 -34.7%

1.0 300 3 3x3 Core 90 -34.7%

1.0 300 3 3x3 None 60 -56.5%

1.5 300 3 3x3 Full 115 0.0%

1.5 300 3 3x3 Perimeter 61.5 -46.5%

1.5 300 3 3x3 Core 60 -47.8%

1.5 300 3 3x3 None 35 -69.6%

2.0 300 3 3x3 Full 85 0.0%

2.0 300 3 3x3 Perimeter 38 -55.3%

2.0 300 3 3x3 Core 37.2 -56.2%

2.0 300 3 3x3 None 25 -70.6%

7.6 Cuplok Joint Stiffness

The Cuplok joint is used as the connection between the standard and ledger of support
scaffold systems. In the parametric studies, the variation of joint stiffness considers the
90% confidence interval (1.645where is the mean value and is the standard
deviation) based on normal distributions of joint statistical data. Variations in the
stiffness of 2-way, 3-way, and 4-way joint connections are included in the studies. The
3 slopes of the tri-linear moment-rotation curves for joint modelling presented in
Chapter 4 are considered at 1.645.The results are presented in Table 7.4 with
comparisons to reference cases (from Section 7.2) in percentage difference for different
Chapter 7 - Parametric Studies of Support Scaffold Systems 143
______________________________________________________________________

lift heights and jack extensions. It is shown that the lower bound of joint stiffness (-
1.645 generally gives higher percentage differences in ultimate system strength than
that of the upper bound (1.645, which means that the system strength is more
susceptible to a reduction in joint stiffness. This is caused by yielding effects in the
joints at the ultimate load. The lower bound joint stiffness shows 9% reduction in the
ultimate system strength on average, regardless of the lift height and jack extensions. In
contrast, the upper bound joint stiffness shows 5% increase in the ultimate system
strength on average. It can be seen that the joint stiffness has a significant effect on the
strength of support scaffold systems.

Table 7.4: Parametric studies of Cuplok joint stiffness of scaffold systems

Jack System
Lift height Number Number Joint Difference
extensions strength
(m) of lifts of bays stiffness (% )
(mm) (kN)

1.0 100 3 3x3 -1.645 173.75 -10.3%

1.0 300 3 3x3 -1.645 131 -5.0%

1.0 600 3 3x3 -1.645 79 -8.1%

1.0 100 3 3x3 1.645 197 1.7%

1.0 300 3 3x3 1.645 143 3.7%

1.0 600 3 3x3 1.645 91 5.8%

1.5 100 3 3x3 -1.645 135 -7.7%

1.5 300 3 3x3 -1.645 105 -8.7%

1.5 600 3 3x3 -1.645 63 -11.3%

1.5 100 3 3x3 1.645 151.5 3.6%

1.5 300 3 3x3 1.645 121 5.2%

1.5 600 3 3x3 1.645 74 4.2%

2.0 100 3 3x3 -1.645 89.4 -7.7%

2.0 300 3 3x3 -1.645 75 -11.8%


Chapter 7 - Parametric Studies of Support Scaffold Systems 144
______________________________________________________________________

Jack System
Lift height Number Number Joint Difference
extensions strength
(m) of lifts of bays stiffness (% )
(mm) (kN)

2.0 600 3 3x3 -1.645 51 -10.5%

2.0 100 3 3x3 1.645 101 4.2%

2.0 300 3 3x3 1.645 91.45 7.6%

2.0 600 3 3x3 1.645 61 7.0%

7.7 Geometric Imperfections

Geometric imperfections, which include member out-of-straightness and storey out-of-


plumb, are studied based on the 90% confidence intervals of the imperfection data
acquired on-site and presented in Chapter 3. Since the support scaffold systems studied
herein are fully braced horizontally at the top, the ultimate system strength is found to
be unaffected by changes in storey out-of-plumb. However, the member out-of-
straightness of standards, normalised with the lift height, is found to significantly affect
the system strength, as shown in Figures 7.8-7.10. The figures show that higher lift
heights lead to greater variations in the system strength. For the 1.0 m lift height, the
system strength is shown to be less affected by the member out-of-straightness when the
normalised out-of-straightness is less than 0.0013. For the 1.5 m and 2.0 m lift heights,
the system strength decreases nearly linearly as the member out-of-straightness
increases. Table 7.5 shows the results of parametric studies of member out-of-
straightness and comparisons to reference cases (from Section 7.2) in percentage
difference for different lift heights and jack extensions. In the table, it should be noticed
that Lh/3040 and Lh/256 are the lower and upper bounds of the 90% confidence interval
based on a lognormal distribution, respectively, where Lh is the lift height. Considering
the upper bound of the 90% confidence interval of member out-of-straightness, the
ultimate system strength is reduced by approximately 9%, 15%, and 19% compared to
the reference cases for the 1.0 m, 1.5 m, and 2.0 m lift heights, respectively.
Nevertheless, the lower bound of the 90% confidence interval of member out-of-
straightness does not show substantial increase in system strength compared to the mean
Chapter 7 - Parametric Studies of Support Scaffold Systems 145
______________________________________________________________________

imperfection. It can be considered that the member out-of-straightness of standards has


significant effect on the strength of support scaffold systems.

200

150
Ultimate system strength (kN)

100 mm jack extensions


100 300 mm jack extensions
600 mm jack extensions

50

0
0.0003 0.0008 0.0013 0.0018 0.0023 0.0028 0.0033 0.0038
Member out-of-straightness (mm/mm)

Figure 7.8: Parametric studies of out-of-straightness of 1.0 m lift scaffold systems

200

150
Ultimate system strength (kN)

100 mm jack extensions


100 300 mm jack extensions
600 mm jack extensions

50

0
0.0003 0.0008 0.0013 0.0018 0.0023 0.0028 0.0033 0.0038
Member out-of-straightness (mm/mm)

Figure 7.9: Parametric studies of out-of-straightness of 1.5 m lift scaffold systems


Chapter 7 - Parametric Studies of Support Scaffold Systems 146
______________________________________________________________________

200

150
Ultimate system strength (kN)

100 mm jack extensions


100 300 mm jack extensions
600 mm jack extensions

50

0
0.0003 0.0008 0.0013 0.0018 0.0023 0.0028 0.0033 0.0038
Member out-of-straightness (mm/mm)

Figure 7.10: Parametric studies of out-of-straightness of 2.0 m lift scaffold systems

Table 7.5: Parametric studies of member out-of-straightness of scaffold systems

Lift Jack System


Number Number Out-of- Difference
height extensions strength
of lifts of bays straightness (% )
(m) (mm) (kN)

1.0 100 3 3x3 Lh/3040 194 0.2%

1.0 300 3 3x3 Lh/3040 138 0.1%

1.0 600 3 3x3 Lh/3040 87 1.2%

1.0 100 3 3x3 Lh/256 176.6 -8.8%

1.0 300 3 3x3 Lh/256 129.6 -6.0%

1.0 600 3 3x3 Lh/256 75 -12.8%

1.5 100 3 3x3 Lh/3040 152.5 4.3%

1.5 300 3 3x3 Lh/3040 117.9 2.5%

1.5 600 3 3x3 Lh/3040 73 2.8%


Chapter 7 - Parametric Studies of Support Scaffold Systems 147
______________________________________________________________________

Lift Jack System


Number Number Out-of- Difference
height extensions strength
of lifts of bays straightness (% )
(m) (mm) (kN)

1.5 100 3 3x3 Lh/256 123.7 -15.4%

1.5 300 3 3x3 Lh/256 98.8 -14.1%

1.5 600 3 3x3 Lh/256 61 -14.1%

2.0 100 3 3x3 Lh/3040 104.25 7.6%

2.0 300 3 3x3 Lh/3040 91 7.1%

2.0 600 3 3x3 Lh/3040 61 7.0%

2.0 100 3 3x3 Lh/256 76.25 -21.3%

2.0 300 3 3x3 Lh/256 66.9 -21.3%

2.0 600 3 3x3 Lh/256 48.25 -15.4%

7.8 Yield Stresses of the Main Components

The parametric studies of yield stress are considered over the 90% confidence interval
(1.645where is the mean value and is the standard deviation)based on normal
distributions with a COV of 0.1 of the 0.2% proof stress of both the standards and jacks
(see Table 7.1). Consequently, the yield stresses at -1.645of the Ramberg-Osgood
material model of the standards and jacks are taken as 443 MPa and 414 MPa,
respectively, and the yield stresses at +1.645 of the standards and jacks are taken as
617 MPa and 576 MPa, respectively.

The results of ultimate system strength are shown in Table 7.6 with comparisons to
reference cases (from Section 7.2) in percentage difference for different lift heights and
jack extensions. The percentage differences show that the 2.0 m lift systems are
unaffected by the variations in the yield stresses of standards and jacks, indicating that
the 2.0 m lift systems fail by elastic buckling of the standards. In addition, the support
scaffold systems with 600 mm jack extensions show insignificant changes in the system
strength with variations of the yield stresses. This is because the main system failure
Chapter 7 - Parametric Studies of Support Scaffold Systems 148
______________________________________________________________________

occurred mainly by elastic buckling of the jacks. The overall differences in system
strength from the reference cases for 1.0 and 1.5 m lifts due to the considered variations
in yield stresses are around 2.2%. Consequently, the yield stresses of main components
are considered to have a somewhat small effect on the system strength of support
scaffold systems with shorter jack extensions and lower lift heights.

Table 7.6: Parametric studies of member yield stress of scaffold systems

Lift Jack System


Number Number Difference
height extensions Yield stress strength
of lifts of bays (% )
(m) (mm) (kN)

1.0 100 3 3x3 -1.645 186.4 -3.8%

1.0 300 3 3x3 -1.645 135 -2.1%

1.0 600 3 3x3 -1.645 86 0.0%

1.0 100 3 3x3 1.645 198.6 2.5%

1.0 300 3 3x3 1.645 141 2.2%

1.0 600 3 3x3 1.645 86 0.0%

1.5 100 3 3x3 -1.645 142.25 -2.7%

1.5 300 3 3x3 -1.645 114 -0.9%

1.5 600 3 3x3 -1.645 70 -1.4%

1.5 100 3 3x3 1.645 149 1.9%

1.5 300 3 3x3 1.645 115 0.0%

1.5 600 3 3x3 1.645 71 0.0%

2.0 100 3 3x3 -1.645 96.9 0.0%

2.0 300 3 3x3 -1.645 85 0.0%

2.0 600 3 3x3 -1.645 57 0.0%

2.0 100 3 3x3 1.645 96.9 0.0%

2.0 300 3 3x3 1.645 85 0.0%


Chapter 7 - Parametric Studies of Support Scaffold Systems 149
______________________________________________________________________

Lift Jack System


Number Number Difference
height extensions Yield stress strength
of lifts of bays (% )
(m) (mm) (kN)

2.0 600 3 3x3 1.645 57 0.0%

7.9 Cross Sections of the Main Components

The parametric studies of cross-sectional properties of the standards and jacks, which
include the variations of outside diameter of the standards and diameter of the jacks, are
considered based on manufacturing tolerance limits imposed by Australian standards
AS 1163 (Standard Australia 2009) and AS 3679.1 (Standard Australia 1996),
correspondingly. According to AS 1163, the minimum and maximum tolerances of
outside diameter of the standards are 47.8 mm and 48.8 mm, respectively. According to
AS 3679.1, the minimum and maximum tolerances of diameter of the jacks are 35.6 mm
and 36.4 mm, respectively. The results of the studies are presented in Table 7.7 with
comparisons to reference cases (from Section 7.2). From the results, the variations of
cross-sectional properties considered herein have insignificant effects on the system
strength as the percentage differences from the reference cases are about 1% on average.

Table 7.7: Parametric studies of member cross-section of scaffold systems

Lift Jack Cross- System


Number Number Difference
height extensions section strength
of lifts of bays (% )
(m) (mm) (Tolerance) (kN)

1.0 100 3 3x3 Minimum 192.2 -0.8%

1.0 300 3 3x3 Minimum 135.9 -1.5%

1.0 600 3 3x3 Minimum 84.5 -1.7%

1.0 100 3 3x3 Maximum 195 0.7%

1.0 300 3 3x3 Maximum 140 1.5%

1.0 600 3 3x3 Maximum 87.5 1.7%


Chapter 7 - Parametric Studies of Support Scaffold Systems 150
______________________________________________________________________

Lift Jack Cross- System


Number Number Difference
height extensions section strength
of lifts of bays (% )
(m) (mm) (Tolerance) (kN)

1.5 100 3 3x3 Minimum 145.45 -0.5%

1.5 300 3 3x3 Minimum 113.75 -1.1%

1.5 600 3 3x3 Minimum 70 -1.4%

1.5 100 3 3x3 Maximum 147.25 0.7%

1.5 300 3 3x3 Maximum 116.5 1.3%

1.5 600 3 3x3 Maximum 72 1.4%

2.0 100 3 3x3 Minimum 96.5 -0.4%

2.0 300 3 3x3 Minimum 83.9 -1.3%

2.0 600 3 3x3 Minimum 56.3 -1.2%

2.0 100 3 3x3 Maximum 97.5 0.6%

2.0 300 3 3x3 Maximum 86.2 1.4%

2.0 600 3 3x3 Maximum 57.8 1.4%

7.10 Discussions

The parameters considered in the studies can be ranked from highest to lowest
influences on the ultimate system strength of support scaffold systems as follows:

(a) Bracing arrangements (up to 71% difference in system strength between full and
none bracings);

(b) Loading eccentricity (up to 62% difference in system strength between 0 mm


and 55 mm loading eccentricities);

(c) Number of lifts (up to 48% difference in system strength between 1 and 9 lifts);
Chapter 7 - Parametric Studies of Support Scaffold Systems 151
______________________________________________________________________

(d) Member out-of-straightness (up to 21% difference in system strength from


reference cases);

(e) Cuplok joint stiffness (up to 12% difference in system strength from reference
cases);

(f) Yield stresses (up to 4% difference in system strength from reference cases);

(g) Cross sections (up to 1.7% difference in system strength from reference cases).

These parametric studies are useful for determining which random variables should be
considered in the probabilistic assessment of system strength of support scaffold
systems. The main parameters, considered as geometric and material uncertainties in
support scaffold systems from highest to lowest influences in ultimate system strength,
are loading eccentricity, member out-of-straightness, joint stiffness, yield stress, and
cross section. Other parameters, i.e. jack extensions, lift heights, number of lifts, and
bracing configurations, are considered to be deterministic and usually determined by the
design. Also, it should be noted that support scaffold systems in general are fully braced.

7.11 Conclusions

This chapter presents the investigation of the factors affecting the strength of support
scaffold systems with Cuplok joints to determine which factors should be included and
treated as random variables in the probabilistic assessment of the support scaffold
systems in Chapter 8. The studied parameters are loading eccentricity, number of lifts,
bracing arrangements, member out-of-straightness, joint stiffness, yield stresses, and
cross sections. The system strengths for the reference cases evaluated at the mean values
of each parameter are presented so that comparisons can be made. Each parameter is
then studied based on 90% confidence interval or specific code tolerances. The loading
eccentricity, member out-of-straightness, joint stiffness, and material yield stress are
found to be the main variables affecting the strength of support scaffold systems.
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 152
______________________________________________________________________

CHAPTER 8 - PROBABILISTIC ASSESSMENT OF


SUPPORT SCAFFOLD SYSTEMS

8.1 Introduction

Recently, researchers have studied the system strength predictions of steel frames by
advanced analysis, involving uncertainties in loads and material yield strength
(Buonopane and Schafer 2006). By nonlinear structural simulations and reliability
analysis, system resistance factors based on certain target reliability for the design of
steel frames by advanced analysis can be determined. The same concept is applied in
this research. The strength of support scaffold systems is affected by variations in
geometric and material parameters, particularly geometric imperfections, joint stiffness,
loading eccentricity, and material yield stress, as examined in Chapter 7. Thus, this
chapter studies the effects of these uncertainties on the ultimate strength of multi-storey
steel support scaffold systems through a rational statistical framework and a second-
order inelastic finite element (advanced) analysis. The chapter also presents a reliability
analysis of support scaffold systems to determine system resistance factors that can be
used in the design by advanced analysis according to the LRFD framework.

A total of 36 stick-type steel scaffold systems with Cuplok joints in various


configurations (different combinations of jack extensions, number of bays, and lift
heights) are considered in the following studies. The statistical data for the main
inherent uncertainties has been acquired through actual field measurements (Chapter 3)
and experimental tests (Chapter 4). The nonlinear finite element models are created in
Strand7 (2009) based on the modelling concepts developed in Chapter 6. A Strand7 API
subroutine is written in C++ to generate random inputs (geometric imperfections, joint
stiffness, loading eccentricity, yield stress and section properties) for the finite element
models, and then around 3,000 executions of advanced analyses for each of the 36
models (more than 100,000 advanced analyses in total) are used to predict variations in
the ultimate system strength based on those parametric uncertainties. The mentioned
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 153
______________________________________________________________________

procedures are also known as Monte Carlo simulations. Appendix D.1 shows a typical
API subroutine used with Strand7 for Monte Carlo simulations. The inherent
uncertainties and the model uncertainty are propagated through Monte Carlo
simulations to obtain the statistics of system strength. By comparison with full-scale
load tests (Chapter 5), the bias and variability of advanced analysis are estimated. The
first-order second-moment method (FOSM) and first-order reliability method (FORM)
are then employed to estimate the reliability of support scaffold systems. Finally,
system resistance factors are proposed for various system configurations to be used with
the design of support scaffold systems by advanced analysis.

8.2 Analysis Models for Probabilistic Assessment

Thirty six configurations of steel support scaffold systems with Cuplok joints are
chosen to be representative of those encountered in scaffold construction. The systems
are categorised by number of bays (1x1 bay, 3x3 bays, 3x6 bays, and 9x9 bays), lift
heights (1.0 m, 1.5 m, and 2.0 m), and jack extensions (100 mm, 300 mm, and 600 mm).
The same bay width of 1829 mm (frequently used in construction) and 3 lifts (same
number of lifts) are selected for all systems. Figures 8.1 to 8.4 show the system
configurations used in the studies. The standards, attached with Cuplok joints, are made
from cold-formed circular steel tube (CHS) grade 450 MPa with nominal outside
diameter of 48.3 mm and thickness of 4 mm. The grade 350 MPa (CHS) ledgers with
end blades are of nominal outside diameter of 48.3 mm and thickness of 3.2 mm. The
telescopic (CHS) braces with hook ends are made of 48.3 mm x 4.0 mm outer tube and
38.2 mm x 3.2 mm inner tube with nominal yield stress of 400 MPa. The adjustable
jacks are made of 36 mm diameter threaded steel rod of grade 430 MPa. The jack
extensions at the top and bottom of the frame are set to equal. The base plates are 180
mm x 180 mm x 10 mm in dimension with a nominal yield stress of 250 MPa. The
spigot joints are assumed to be located at the mid-height of the second and third lifts.
There are no spigots in the bottom lift based on general construction practice. All the
systems are considered fully braced. For the case of 9x9 bays, the frame is braced every
fourth bay according to general practice and no corner braces are used, as illustrated in
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 154
______________________________________________________________________

Figure 8.4. Also, there are no corner braces for the cases of 3x3 bays and 3x6 bays, as
shown in Figures 8.2 and 8.3, respectively.

Figure 8.1: 1x1 bay scaffold configuration in the studies

Figure 8.2: 3x3 bays scaffold configuration in the studies


Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 155
______________________________________________________________________

Figure 8.3: 3x6 bays scaffold configuration in the studies


Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 156
______________________________________________________________________

Figure 8.4: 9x9 bays scaffold configuration in the studies


Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 157
______________________________________________________________________

The three-dimensional finite element models have been created with the same
modelling concept presented in Chapter 6 for Monte Carlo simulations. The models
consist of spigot joints, nonlinear moment-rotation joint stiffness for the standard-to-
ledger connections, brace connections with 1.8 kN/mm axial spring stiffness and 60 mm
offset from the nodal points, base plate eccentricity, loading eccentricity, geometric
imperfections implemented directly at the nodes, and Ramberg-Osgood material model
for all of the scaffold components. For the top boundary conditions, the rotational
stiffness about the x-axis is assumed to be rigid, corresponding to the negligible strong
axis bending of the supported timber bearer; however, the y-axis bending stiffness is
taken as 40 kNm/rad as obtained from the model calibrations in Chapter 6. For the
bottom boundary conditions, the base plate model is applied with the eccentricity of 15
mm (close to maximum allowable bottom eccentricity specified in AS 3610 by Standard
Australia 1995) to each upright. Equal vertical point loads are applied with random
variations in eccentricity on top of each U-head screw jack. All frames are assumed to
be braced laterally at the top of the U-head screw jacks. Figure 8.5 shows a typical finite
element model of the 3x6 bays, 3 lifts, and 1 m lift height with 100 mm jack extensions.

Figure 8.5: Typical finite element model for Monte Carlo simulations
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 158
______________________________________________________________________

8.3 Uncertainties in Support Scaffold Systems

In scaffold construction, components such as standards, ledgers, jacks, joints and base
plates are reused from one job to another and new members are mixed with old ones.
Most of the components come with imperfections, whether from repeated use and/or the
manufacturing process. As a result, scaffold systems are prone to many uncertainties
that can be characterised into variations in geometric and material properties, as well as
joint stiffness. In this research, the uncertainties being considered include cross-
sectional area and moment of inertia of the standards and jacks, yield stress of the
standards and jacks, joint stiffness for standard-to-ledger connections, initial geometric
imperfections, and loading eccentricity. These uncertainties are treated as random
variables in the Monte Carlo simulations to obtain the statistics of the system strength.

8.3.1 Cross-Sectional Area and Moment of Inertia

Due to uncertainty in manufacturing environments, scaffold components can possess


different dimensions that lead to variations in the cross-sectional area and moment of
inertia of the sections. This research considers cross-sectional area and moment of
inertia of the main loaded components, which are the standards and jacks, as random
variables, while those of other components are assumed deterministic at their mean
values (taken conservatively to be the same as their nominal values in the absence of
available data). For the purpose of Monte Carlo simulations, the statistical data for the
cross-sectional area and moment of inertia of the standards and jacks is obtained from
dimensional measurements of around 80 second-hand standards and jacks. The mean
cross-sectional areas of the standards and jacks are computed to be 556.69 mm2 and
1017.876 mm2, respectively, with a coefficient of variation (COV) of 0.05 for both. The
mean moments of inertia of the standards and jacks are 137675.8 mm4 and 82447.96
mm4, respectively, with a coefficient of variation (COV) of 0.05 for both. Both cross-
sectional area and moment of inertia of the standards and jacks are statistically modelled
as lognormal distributions. For each simulation, one element in the model possesses its
own random values of cross-sectional area and moment of inertia.
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 159
______________________________________________________________________

8.3.2 Yield Stress

Because of uncertainty in manufacturing process, each type of scaffold components can


show a significant variation in yield stress which affects the strength of scaffold systems.
For the purpose of Monte Carlo simulations, the statistical data for the (static) yield
stress is obtained from the literature (Galambos and Ravindra 1978). The mean yield
stress is assumed to be 1.10Fy with a coefficient of variation (COV) of 0.1, where Fy
represents the minimum specified (nominal) yield stress for the grade of steel. Only the
yield stresses of the main components in loading (the standards and jacks) are important
and considered as random variables in the simulations, whereas those of other
components are assumed deterministic at their mean values (see Table 6.3). Therefore,
the mean yield stresses of the standards and jacks are 495 MPa and 473 MPa,
respectively with a coefficient of variation (COV) of 0.1 for both. They are statistically
modelled as normal distributions in the simulations. In one simulation, each element in
the model has its own random value of yield stress.

8.3.3 Joint Stiffness

The joints between standard and ledger present uncertain behaviours due to
imperfections from repeated use and consequent varying degrees of fixity. In order to
predict the overall stability of the frame and strength of the system, the standard-to-
ledger connections must be modelled correctly. Cuplok joints are considered in this
research as standard-to-ledger connections for support scaffold systems. Experimental
joint tests (Chapter 4) have shown that Cuplok joints behave as semi-rigid connections
and usually show looseness with small stiffness at the beginning of loading due to a lack
of fit (Figure 8.6(a)). For Monte Carlo simulations, a tri-linear joint model (Chapter 4)
as shown in Figure 8.6(b) is used to approximate the relation between the moment and
rotation of Cuplok joints.

Since probabilistic assessment of the strength of scaffold systems depends mainly on


the stiffness of joint bent in the vertical plane, but not in the horizontal plane, only
variation in the vertical bending stiffness is considered, and the modelling of the
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 160
______________________________________________________________________

horizontal bending stiffness can be assumed to be deterministic using the mean values
for the tri-linear curve, as presented in Tables 8.1-8.2 for bending about the vertical axis.
For practicality in probabilistic modelling, the values of 1, 2 and 3 are assumed to be
deterministic, taken as their mean values; however, the variations of k1, k2, and k3 for
bending about the horizontal axis are taken into account. The coefficients of variation of
the joint stiffness for k1, k2, and k3 for bending about the horizontal axis (vertical
bending stiffness) for different joint configurations are presented in Table 8.3. Each
joint stiffness is statistically modelled as a normal distribution in the simulations as
explained in Chapter 4.

Figure 8.6: Graph of (a) typical Cuplok joint behaviour (b) tri-linear joint model

Table 8.1: Mean Cuplok joint stiffness (kNm/rad)

Bending stiffness about horizontal Bending stiffness about vertical


axis (kNm/rad) axis (kNm/rad)
Joint
k1 k2 k3 k1 k2 k3
configuration

4-way 39 102 5.3 15 7.5 0.8

3-way 36 87 5.1 14 7 1

2-way 41 77 4.6 7.5 5 1.5


Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 161
______________________________________________________________________

Table 8.2: Mean rotation for Cuplok joints (rad)

Rotation about horizontal axis


Rotation about vertical axis (rad)
(rad)
Joint
1 2 3 1 2 3
configuration

4-way 0.014 0.036 0.16 0.02 0.04 0.1

3-way 0.012 0.036 0.16 0.02 0.04 0.1

2-way 0.007 0.036 0.16 0.02 0.04 0.1

Table 8.3: Coefficient of variation of Cuplok joint stiffness

Bending about horizontal axis


Joint
k1 k2 k3
configuration

4-way 0.22 0.18 0.30

3-way 0.38 0.21 0.37

2-way 0.35 0.20 0.46

8.3.4 Initial Geometric Imperfections

Scaffold systems are slender in nature, thus the strength and behaviour of support
scaffolds are susceptible to adverse 2nd order member P- and frame P- effects.
Therefore, initial geometric imperfections producing 2nd order effects must be
considered in the analysis models. Two types of initial geometric imperfections,
specifically imperfections from storey out-of-plumb (initial frame sway) and out-of-
straightness of the standards (initial member crookedness) are considered in the Monte
Carlo simulations. It is obvious that there is a lack of data available for the out-of-
straightness and out-of-plumb of scaffold systems in the literature. The Australian
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 162
______________________________________________________________________

standard (Standard Australia 1998) allows a maximum out-of-straightness of Lh/1000


and a maximum out-of plumb of H/400, where Lh is the column length and H is the
storey height. Nevertheless, these allowable values are proposed for steel buildings
rather than scaffold systems. In addition, the out-of-straightness of the standards often
exceeds the tolerance values imposed by the codes, especially where the standards are
connected with spigot joints, as shown in the survey data in Chapter 3.

The observed imperfection data shows that the initial out-of-straightness of standards
without spigots has a mean of Lh/2500 and a COV of 0.75. However, the initial out-of-
straightness of standards with spigot joints increases considerably at the location of the
spigot joint as presented in Chapter 3 to have a mean of Lh/770 and a COV of 0.615.
The probability distribution for the member out-of-straightness is taken as lognormal
(Chapter 3). The storey out-of-plumb is fitted to a normal distribution with a mean of
H/625 and a COV of 0.313 (Chapter 3). Out-of-straightness values of the standards and
the storey out-of plumb are applied to the analysis models, treating each standard
independently (uncorrelated) and assigning random magnitude and direction of out-of-
straightness in the x-axis for each standard, as shown in Figure 8.5. Table 8.4 shows the
statistical data for the initial geometric imperfections used in the simulations.

Table 8.4: Statistical data for initial geometric imperfections used in the simulations

Random variable Mean COV Distribution

Out-of-straightness of the standards


Lh/2500 0.75 Lognormal
without spigot joints

Out-of-straightness of the standards with


Lh/770 0.615 Lognormal
spigot joints

Storey out-of-plumb H/625 0.313 Normal


Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 163
______________________________________________________________________

8.3.5 Loading Eccentricity

The strength of support scaffold systems is highly influenced by the loading eccentricity
between the timber bearer and the U-head, as examined in the parametric studies
(Chapter 7). As a result, loading eccentricity must be included in the probabilistic
assessment of the strength of support scaffold systems. The Australian standard for
formwork (Standard Australia 1995) limits the eccentricity at 40 mm or 25% of the
bearer width, whichever is less (18.75 mm for the Cuplok systems investigated herein).

It can be seen from the data obtained from on-site measurements of the loading
eccentricity in Chapter 3 that the mean value (18.10 mm) is approximately equal to the
limit imposed by the code; however, for the purpose of Monte Carlo simulations the
loading eccentricity is limited at not more than 55 mm (maximum possible loading
eccentricity for the studied system) with the mean of 18.10 mm and a COV of 0.608 in
the analysis models. The loading eccentricity is modelled by a lognormal distribution in
the simulations and applied to the analysis models as random in magnitude and
direction in the x-axis, direction perpendicular to the timber bearers, as shown in Figure
8.5. Table 8.5 summarises the statistical data for the initial loading eccentricity used in
the Monte Carlo simulations.

It should be noticed that eccentricity can also occur at the bottom due to uneven or
sloped ground. The amount of base eccentricity depends largely on ground surface
irregularities. The Australian standard AS 3610 (Standard Australia 1995) specifies an
expected base eccentricity of no more than 40 mm or bp/4, whichever is less, where bp is
the stiff portion of bearing of an end plate (as shown in Chapter 6). For example, bp/4 is
17 mm for the support scaffold system studied in this research which is less than 40
mm; therefore, the expected base eccentricity of the system is no more than 17 mm.
Since there is no statistical data for the base eccentricity available in the literature and
acquisition of such data is difficult and subjective, the bottom eccentricity is assumed
conservatively to be 15 mm applied in the same direction to each upright with the base
plate modelling concept presented in Chapter 6.
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 164
______________________________________________________________________

Table 8.5: Statistical data for loading eccentricity used in the simulations

Random variable Mean COV Distribution

Loading eccentricity 18.10 mm 0.608 Lognormal

8.4 Ultimate Strength of Support Scaffold Systems

Once the three-dimensional finite element models of 36 configurations of steel support


scaffold systems with Cuplok joints had been created, each model was modified for
each analysis with the generated values of random variables considered herein through
an API subroutine, (see typical subroutine in Appendix D.1), and around 3,000
advanced analyses on each model were performed in Strand7 (2009) to obtain the
statistics of the ultimate strength for the 36 system configurations.

8.4.1 Monte Carlo Simulation Results

The relative frequency histograms of the ultimate strength of the 36 different system
configurations are presented in Figures 8.7-8.42. The mean ( R ) and coefficient of
variation (VR) for the system ultimate strength of the 36 systems studied are summarised
in Table 8.6.
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 165
______________________________________________________________________

0.14

0.12

0.1 Mean 188.2


Relative frequency

STD 20.28
0.08 COV 0.11

0.06

0.04

0.02

0
0

0
12

13

14

15

16

18

19

20

21

22

24
Frame strength (kN)

Figure 8.7: Histogram for the ultimate strength of 1x1 bay, 1 m lift height, 3 lifts, and
100 mm top and bottom jack extensions support scaffold frame

0.25

0.2

Mean 148.7
Relative frequency

0.15 STD 7.68


COV 0.05

0.1

0.05

0
0

0
11

11

12

12

13

14

14

15

15

16

17

Frame strength (kN)

Figure 8.8: Histogram for the ultimate strength of 1x1 bay, 1 m lift height, 3 lifts, and
300 mm top and bottom jack extensions support scaffold frame
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 166
______________________________________________________________________

0.35

0.3

0.25
Relative frequency

Mean 94.0
0.2 STD 3.02
COV 0.03
0.15

0.1

0.05

0
78

80

82

84

86

88

90

92

94

96

98

6
10

10

10

10
Frame strength (kN)

Figure 8.9: Histogram for the ultimate strength of 1x1 bay, 1 m lift height, 3 lifts, and
600 mm top and bottom jack extensions support scaffold frame

0.25

0.2
Relative frequency

Mean 181.1
0.15
STD 12.27
COV 0.07
0.1

0.05

0
0

0
12

13

14

15

16

18

19

20

21

22

24

Frame strength (kN)

Figure 8.10: Histogram for the ultimate strength of 3x3 bays, 1 m lift height, 3 lifts, and
100 mm top and bottom jack extensions support scaffold frame
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 167
______________________________________________________________________

0.2
0.18
0.16
0.14
Relative frequency

0.12 Mean 141.9


0.1 STD 4.70
COV 0.03
0.08
0.06
0.04
0.02
0
8

8
11

12

12

13

13

13

14

14

15

15

15
Frame strength (kN)

Figure 8.11: Histogram for the ultimate strength of 3x3 bays, 1 m lift height, 3 lifts, and
300 mm top and bottom jack extensions support scaffold frame

0.5
0.45
0.4
0.35
Relative frequency

0.3
Mean 90.7
0.25 STD 2.29
0.2 COV 0.03

0.15
0.1
0.05
0
94

96

98
90

92
78

80

82

84

86

88

6
10

10

10

10

Frame strength (kN)

Figure 8.12: Histogram for the ultimate strength of 3x3 bays, 1 m lift height, 3 lifts, and
600 mm top and bottom jack extensions support scaffold frame
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 168
______________________________________________________________________

0.25

0.2
Relative frequency

0.15
Mean 184.5
STD 11.30
0.1 COV 0.06

0.05

0
0

0
12

13

14

15

16

18

19

20

21

22

24
Frame strength (kN)

Figure 8.13: Histogram for the ultimate strength of 3x6 bays, 1 m lift height, 3 lifts, and
100 mm top and bottom jack extensions support scaffold frame

0.14

0.12

0.1
Relative frequency

0.08 Mean 145.7


STD 6.55
0.06 COV 0.04

0.04

0.02

0
0

0
12

12

12

13

13

14

14

14

15

15

16

Frame strength (kN)

Figure 8.14: Histogram for the ultimate strength of 3x6 bays, 1 m lift height, 3 lifts, and
300 mm top and bottom jack extensions support scaffold frame
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 169
______________________________________________________________________

0.6

0.5
Relative frequency
0.4

Mean 94.5
0.3 STD 1.69
COV 0.02
0.2

0.1

0
78

80

82

84

86

88

90

92

94

96

98

6
10

10

10

10
Frame strength (kN)

Figure 8.15: Histogram for the ultimate strength of 3x6 bays, 1 m lift height, 3 lifts, and
600 mm top and bottom jack extensions support scaffold frame

0.25

0.2
Relative frequency

0.15 Mean 184.3


STD 11.87
COV 0.06
0.1

0.05

0
0

0
12

13

14

15

16

18

19

20

21

22

24

Frame strength (kN)

Figure 8.16: Histogram for the ultimate strength of 9x9 bays, 1 m lift height, 3 lifts, and
100 mm top and bottom jack extensions support scaffold frame
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 170
______________________________________________________________________

0.3

0.25

0.2
Relative frequency

Mean 155.5
0.15 STD 4.66
COV 0.03

0.1

0.05

0
0 4 8 2 6 0 4 8 2
13 13 13 14 14 15 15 15 16
Frame strength (kN)

Figure 8.17: Histogram for the ultimate strength of 9x9 bays, 1 m lift height, 3 lifts, and
300 mm top and bottom jack extensions support scaffold frame

0.5
0.45
0.4
0.35
Relative frequency

0.3
0.25 Mean 95.5
STD 1.77
0.2
COV 0.02
0.15
0.1
0.05
0
88

90

92

94

96

98
78

80

82

84

86

6
10

10

10

10

Frame strength (kN)

Figure 8.18: Histogram for the ultimate strength of 9x9 bays, 1 m lift height, 3 lifts, and
600 mm top and bottom jack extensions support scaffold frame
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 171
______________________________________________________________________

0.16

0.14

Relative frequency 0.12

0.1
Mean 143.0
0.08 STD 10.11
COV 0.07
0.06

0.04

0.02

0
86

94

6
10

11

11

12

13

14

15

15

16
Frame strength (kN)

Figure 8.19: Histogram for the ultimate strength of 1x1 bay, 1.5 m lift height, 3 lifts,
and 100 mm top and bottom jack extensions support scaffold frame

0.25

0.2
Relative frequency

0.15 Mean 116.3


STD 4.49
COV 0.04
0.1

0.05

0
88

92

96

8
10

10

10

11

11

12

12

12

Frame strength (kN)

Figure 8.20: Histogram for the ultimate strength of 1x1 bay, 1.5 m lift height, 3 lifts,
and 300 mm top and bottom jack extensions support scaffold frame
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 172
______________________________________________________________________

0.35

0.3

0.25
Relative frequency

0.2
Mean 74.7
STD 2.73
0.15 COV 0.04

0.1

0.05

0
58

60

62

64

66

68

70

72

74

76

78

80

82

84

86
Frame strength (kN)

Figure 8.21: Histogram for the ultimate strength of 1x1 bay, 1.5 m lift height, 3 lifts,
and 600 mm top and bottom jack extensions support scaffold frame

0.2
0.18
0.16
0.14
Relative frequency

0.12 Mean 143.4


0.1 STD 8.27
COV 0.06
0.08
0.06
0.04
0.02
0
86

94

6
10

11

11

12

13

14

15

15

16

Frame strength (kN)

Figure 8.22: Histogram for the ultimate strength of 3x3 bays, 1.5 m lift height, 3 lifts,
and 100 mm top and bottom jack extensions support scaffold frame
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 173
______________________________________________________________________

0.35

0.3

0.25
Relative frequency

0.2 Mean 114.4


STD 3.03
0.15 COV 0.03

0.1

0.05

0
88

92

96

8
10

10

10

11

11

12

12

12
Frame strength (kN)

Figure 8.23: Histogram for the ultimate strength of 3x3 bays, 1.5 m lift height, 3 lifts,
and 300 mm top and bottom jack extensions support scaffold frame

0.6

0.5

0.4
Relative frequency

Mean 70.6
0.3 STD 1.60
COV 0.02
0.2

0.1

0
58

60

62

64

66

68

70

72

74

76

78

80

82

84

86

Frame strength (kN)

Figure 8.24: Histogram for the ultimate strength of 3x3 bays, 1.5 m lift height, 3 lifts,
and 600 mm top and bottom jack extensions support scaffold frame
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 174
______________________________________________________________________

0.25

0.2
Relative frequency

0.15
Mean 140.7
STD 7.72
0.1 COV 0.05

0.05

0
86

94

6
10

11

11

12

13

14

15

15

16
Frame strength (kN)

Figure 8.25: Histogram for the ultimate strength of 3x6 bays, 1.5 m lift height, 3 lifts,
and 100 mm top and bottom jack extensions support scaffold frame

0.35

0.3

0.25
Relative frequency

0.2
Mean 120.8
STD 3.42
0.15
COV 0.03

0.1

0.05

0
92

96

2
10

10

10

11

11

12

12

12

13

Frame strength (kN)

Figure 8.26: Histogram for the ultimate strength of 3x6 bays, 1.5 m lift height, 3 lifts,
and 300 mm top and bottom jack extensions support scaffold frame
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 175
______________________________________________________________________

0.5
0.45
0.4
0.35
Relative frequency

0.3 Mean 73.8


STD 1.38
0.25
COV 0.02
0.2
0.15

0.1
0.05

0
58

60

62

64

66

68

70

72

74

76

78

80

82

84

86
Frame strength (kN)

Figure 8.27: Histogram for the ultimate strength of 3x6 bays, 1.5 m lift height, 3 lifts,
and 600 mm top and bottom jack extensions support scaffold frame

0.3

0.25
Relative frequency

0.2

Mean 146.2
0.15
STD 7.38
COV 0.05
0.1

0.05

0
2

6
86

94

10

11

11

12

13

14

15

15

16

Frame strength (kN)

Figure 8.28: Histogram for the ultimate strength of 9x9 bays, 1.5 m lift height, 3 lifts,
and 100 mm top and bottom jack extensions support scaffold frame
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 176
______________________________________________________________________

0.4

0.35

Relative frequency 0.3

0.25
Mean 124.1
0.2 STD 3.42
COV 0.03
0.15

0.1

0.05

0
92

96

2
10

10

10

11

11

12

12

12

13
Frame strength (kN)

Figure 8.29: Histogram for the ultimate strength of 9x9 bays, 1.5 m lift height, 3 lifts,
and 300 mm top and bottom jack extensions support scaffold frame

0.7

0.6

0.5
Relative frequency

0.4
Mean 75.1
STD 1.36
0.3 COV 0.02

0.2

0.1

0
58

60

62

64

66

68

70

72

74

76

78

80

82

84

86

Frame strength (kN)

Figure 8.30: Histogram for the ultimate strength of 9x9 bays, 1.5 m lift height, 3 lifts,
and 600 mm top and bottom jack extensions support scaffold frame
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 177
______________________________________________________________________

0.25

0.2
Relative frequency

0.15 Mean 98.3


STD 5.85
COV 0.06
0.1

0.05

0
60

66

72

78

84

90

96

0
10

10

11

12
Frame strength (kN)

Figure 8.31: Histogram for the ultimate strength of 1x1 bay, 2 m lift height, 3 lifts, and
100 mm top and bottom jack extensions support scaffold frame

0.18

0.16

0.14

0.12
Relative frequency

0.1
Mean 85.7
0.08 STD 5.08
COV 0.06
0.06

0.04

0.02

0
56

60

64

68

72

76

80

84

88

92

96

Frame strength (kN)

Figure 8.32: Histogram for the ultimate strength of 1x1 bay, 2 m lift height, 3 lifts, and
300 mm top and bottom jack extensions support scaffold frame
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 178
______________________________________________________________________

0.3

0.25

0.2
Relative frequency

Mean 60.8
0.15 STD 3.05
COV 0.05

0.1

0.05

0
38

42

46

50

54

58

62

66

70

74

78
Frame strength (kN)

Figure 8.33: Histogram for the ultimate strength of 1x1 bay, 2 m lift height, 3 lifts, and
600 mm top and bottom jack extensions support scaffold frame

0.3

0.25
Relative frequency

0.2

Mean 98.5
0.15 STD 6.12
COV 0.06
0.1

0.05

0
60

66

72

78

84

90

96

0
10

10

11

12

Frame strength (kN)

Figure 8.34: Histogram for the ultimate strength of 3x3 bays, 2 m lift height, 3 lifts, and
100 mm top and bottom jack extensions support scaffold frame
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 179
______________________________________________________________________

0.25

0.2
Relative frequency

0.15
Mean 87.9
STD 4.81
0.1 COV 0.05

0.05

0
58

62

66

70

74

78

82

86

90

94

98
Frame strength (kN)

Figure 8.35: Histogram for the ultimate strength of 3x3 bays, 2 m lift height, 3 lifts, and
300 mm top and bottom jack extensions support scaffold frame

0.6

0.5

0.4
Relative frequency

0.3 Mean 58.9


STD 1.99
COV 0.03
0.2

0.1

0
38

42

46

50

54

58

62

66

70

74

78

Frame strength (kN)

Figure 8.36: Histogram for the ultimate strength of 3x3 bays, 2 m lift height, 3 lifts, and
600 mm top and bottom jack extensions support scaffold frame
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 180
______________________________________________________________________

0.3

0.25
Relative frequency
0.2

Mean 98.5
0.15 STD 5.38
COV 0.05
0.1

0.05

0
60

66

72

78

84

90

96

0
10

10

11

12
Frame strength (kN)

Figure 8.37: Histogram for the ultimate strength of 3x6 bays, 2 m lift height, 3 lifts, and
100 mm top and bottom jack extensions support scaffold frame

0.3

0.25
Relative frequency

0.2

Mean 89.6
0.15 STD 4.77
COV 0.05
0.1

0.05

0
62

66

70

74

78

82

86

90

94

98

2
10

Frame strength (kN)

Figure 8.38: Histogram for the ultimate strength of 3x6 bays, 2 m lift height, 3 lifts, and
300 mm top and bottom jack extensions support scaffold frame
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 181
______________________________________________________________________

0.5

0.45
0.4

0.35
Relative frequency

0.3

0.25 Mean 61.3


STD 2.07
0.2 COV 0.03
0.15
0.1
0.05
0
38

42

46

50

54

58

62

66

70

74

78
Frame strength (kN)

Figure 8.39: Histogram for the ultimate strength of 3x6 bays, 2 m lift height, 3 lifts, and
600 mm top and bottom jack extensions support scaffold frame

0.3

0.25
Relative frequency

0.2

Mean 100.7
0.15
STD 5.06
COV 0.05
0.1

0.05

0
60

66

72

78

84

90

96

0
10

10

11

12

Frame strength (kN)

Figure 8.40: Histogram for the ultimate strength of 9x9 bays, 2 m lift height, 3 lifts, and
100 mm top and bottom jack extensions support scaffold frame
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 182
______________________________________________________________________

0.3

0.25
Relative frequency
0.2
Mean 90.1
0.15 STD 4.87
COV 0.05

0.1

0.05

0
62

66

70

74

78

82

86

90

94

98

2
10
Frame strength (kN)

Figure 8.41: Histogram for the ultimate strength of 9x9 bays, 2 m lift height, 3 lifts, and
300 mm top and bottom jack extensions support scaffold frame

0.7

0.6

0.5
Relative frequency

0.4
Mean 62.0
0.3 STD 2.06
COV 0.03
0.2

0.1

0
58

66

70

74

78
38

42

46

50

54

62

Frame strength (kN)

Figure 8.42: Histogram for the ultimate strength of 9x9 bays, 2 m lift height, 3 lifts, and
600 mm top and bottom jack extensions support scaffold frame
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 183
______________________________________________________________________

Table 8.6: Statistical summary of simulation results for system ultimate strengths

Jack
Lift height Number of Number of R (kN) VR
extensions
(m) lifts bays
(mm)
1.0 100 3 1x1 188.2 0.11

1.0 300 3 1x1 148.7 0.05

1.0 600 3 1x1 94.0 0.03

1.0 100 3 3x3 181.1 0.07

1.0 300 3 3x3 141.9 0.03

1.0 600 3 3x3 90.7 0.03

1.0 100 3 3x6 184.5 0.06

1.0 300 3 3x6 145.7 0.04

1.0 600 3 3x6 94.5 0.02

1.0 100 3 9x9 184.3 0.06

1.0 300 3 9x9 155.5 0.03

1.0 600 3 9x9 95.5 0.02

1.5 100 3 1x1 143.0 0.07

1.5 300 3 1x1 116.3 0.04

1.5 600 3 1x1 74.7 0.04

1.5 100 3 3x3 143.4 0.06

1.5 300 3 3x3 114.4 0.03

1.5 600 3 3x3 70.6 0.02

1.5 100 3 3x6 140.7 0.05

1.5 300 3 3x6 120.8 0.03

1.5 600 3 3x6 73.8 0.02

1.5 100 3 9x9 146.2 0.05

1.5 300 3 9x9 124.1 0.03

1.5 600 3 9x9 75.1 0.02


Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 184
______________________________________________________________________

Jack
Lift height Number of Number of R (kN) VR
extensions
(m) lifts bays
(mm)
2.0 100 3 1x1 98.3 0.06

2.0 300 3 1x1 85.7 0.06

2.0 600 3 1x1 60.8 0.05

2.0 100 3 3x3 98.5 0.06

2.0 300 3 3x3 87.9 0.05

2.0 600 3 3x3 58.9 0.03

2.0 100 3 3x6 98.5 0.05

2.0 300 3 3x6 89.6 0.05

2.0 600 3 3x6 61.3 0.03

2.0 100 3 9x9 100.7 0.05

2.0 300 3 9x9 90.1 0.05

2.0 600 3 9x9 62.0 0.03

The histograms are shown to be skewed to the left and appear more scattered when the
jack extension becomes shorter, resulting in higher coefficients of variation. Also, with
shorter lift height, the COV appears to be larger. These results could be expected since
with shorter lift height and jack extension, the nonlinear material effects become more
pronounced, and hence the variation in yield stress becomes a main controlling factor,
which added to other uncertainties, would lead to higher variability in system strength.
This assertion has been confirmed by the stress and strain results obtained at the
ultimate load in the advanced analysis of the systems with 1 m lift height or 100 mm
jack extension, which show that the stress and strain of the standards and jacks are in
the plastic region at ultimate. On the other hand, for the systems with 2 m lift or 600
mm jack extension, the ultimate stress and strain results are still in the elastic range and
the failure of these systems occurs essentially by elastic buckling. Figures 8.43 and 8.44
show typical failure modes and stresses at the ultimate loads of the 1 m lift and 2 m lift
systems, respectively. According to the results with similar lift height and jack
extension, the mean strength increases by a small amount when the system becomes
larger from 3x3 bays to 9x9 bays; however, with the 1x1 bay system, the mean strength
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 185
______________________________________________________________________

can show somewhat higher value than that of the 3x3 bays with the same lift height and
jack extensions, since the system with 3x3 bays does not have braces at the corners
which can weaken the system strength. In all cases, the variability (COV) of the systems
with the same lift height and jack extensions decreases when the system becomes larger
due to the structural redundancy and stochastic averaging of uncorrelated member
properties.

Figure 8.43: Typical failure mode of 1.0 m lift system with 100 mm jack extensions

Figure 8.44: Typical failure mode of 2.0 m lift system with 600 mm jack extensions
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 186
______________________________________________________________________

8.4.2 Nominal Design Models

The nominal design models are created in the same way as the analysis models (see
Sections 8.2 and 9.2), except that all inputs are applied at nominal or proposed values
and all material models are assumed to be elastic-perfectly plastic (Figure 8.45) at the
nominal yield stress (fy). The pattern of geometric imperfections and loading
eccentricity is set to be at the worst case scenario, i.e. first buckling mode shape. These
nominal models are intended for engineers to use in practice when designing scaffold
systems by advanced analysis. The input parameters for the nominal design models are
presented in Table 8.7. Thirty six nominal design models for the system configurations
analysed in the simulations presented in Section 8.4.1 were analysed by advanced
analyses to obtain the nominal ultimate strength (Rn) for each. The results are shown in
Table 8.8.

Table 8.7: Input parameters for nominal design models

Parameter Nominal value Description

Storey out-of-plumb H/600 H = height of the system

Out-of-straightness of the standards Lh/700 Lh = lift height

Loading eccentricity 19 mm Mean value from survey

Bottom eccentricity 15 mm See Section 8.3.5

Section properties Nominal See Section 8.2

Joint stiffness Mean See Tables 8.1 and 8.2

Yield stress for the standard 450 MPa

Yield stress for the jack 430 MPa


Nominal steel grades (see
Yield stress for the spigot 400 MPa Figure 8.45 for elastic-
Yield stress for the ledger 350 MPa perfectly plastic material
model)
Yield stress for the brace 400 MPa

Yield stress for the base plate 250 MPa


Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 187
______________________________________________________________________

Table 8.8: Nominal ultimate strength for nominal design models

Jack
Lift height Number of Number of Rn (kN)
extensions
(m) lifts bays
(mm)
1.0 100 3 1x1 141.0

1.0 300 3 1x1 126.3

1.0 600 3 1x1 87.7

1.0 100 3 3x3 137.0

1.0 300 3 3x3 119.3

1.0 600 3 3x3 83.0

1.0 100 3 3x6 137.0

1.0 300 3 3x6 119.3

1.0 600 3 3x6 83.0

1.0 100 3 9x9 140.0

1.0 300 3 9x9 122.0

1.0 600 3 9x9 87.0

1.5 100 3 1x1 121.7

1.5 300 3 1x1 106.0

1.5 600 3 1x1 70.2

1.5 100 3 3x3 120.2

1.5 300 3 3x3 107.5

1.5 600 3 3x3 69.0

1.5 100 3 3x6 119.9

1.5 300 3 3x6 107.7

1.5 600 3 3x6 71.0

1.5 100 3 9x9 124.1

1.5 300 3 9x9 116.2

1.5 600 3 9x9 74.1


Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 188
______________________________________________________________________

Jack
Lift height Number of Number of Rn (kN)
extensions
(m) lifts bays
(mm)
2.0 100 3 1x1 87.7

2.0 300 3 1x1 77.0

2.0 600 3 1x1 56.5

2.0 100 3 3x3 88.7

2.0 300 3 3x3 80.5

2.0 600 3 3x3 57.0

2.0 100 3 3x6 88.8

2.0 300 3 3x6 81.3

2.0 600 3 3x6 59.5

2.0 100 3 9x9 91.0

2.0 300 3 9x9 83.5

2.0 600 3 9x9 61.2

fy fy is the nominal yield stress.

E is the Youngs modulus.


E
1

Figure 8.45: Elastic-perfectly plastic model for steel

8.4.3 Modelling Uncertainty and Statistics of System Resistance

As mentioned in Section 8.3, the Monte Carlo simulations consider only the inherent
randomness in basic variables, often referred to as aleatory in structural engineering,
which relates to luck or chance (Ayyub and McCuen 2003, and Bulleit 2008). Another
type of uncertainty, which arises from simplifications and assumptions in the modelling
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 189
______________________________________________________________________

of support scaffold systems, must also be accounted for in the reliability analysis in
order to develop probabilistic-based design by advanced analysis. This source of
uncertainty, also called epistemic or modelling error, includes simplified boundary
conditions, modelised initial geometric imperfections, idealised moment-rotation curves
for joint stiffness, and inherent inaccuracy in the discretisation of the structure into a
finite element model. The mean error and coefficient of variation (COV) of the
advanced analysis model can be obtained from the ratios of fifteen comparisons
presented in Section 6.2.8 between the measured ultimate loads (Rtest) and the ultimate
load predictions from advanced analysis (Rpredict). The mean error is 1.014 with a COV
of 0.1 (Chapter 6), suggesting that the advanced analysis of finite element models
proposed in this research provides good indication of the true system strength. The
variability of Rtest / Rpredict (Vtest/predict) comes from three causes (Ellingwood et al. 1980),
as explained by:

/ predict V M VT V S
2 2 2 2
Vtest (8.1)

where VM represents the modelling uncertainty, VT represents the uncertainties in testing


procedures and readings, and VS represents the variations in test environments and
actual test specimen dimensions from those measured. In general, VT and VS are in the
range of 0.02-0.04 for members (Melchers 1999). For the structural system studied
herein, the effects from VT and VS are difficult to quantify, and are conservatively
ignored, thus taking VM equal to Vtest/predict. To incorporate the effect of modelling error,
the predicted ultimate load result from each Monte Carlo simulation for the same
system configuration is modified by multiplying the ultimate load by a randomly
generated value of modelling error, which is assumed to be normally distributed with a
mean of 1.014 and a COV of 0.1 (Chapter 6). After combining the modelling
uncertainty, the statistics for the true system resistance of support scaffold systems are
obtained. The mean-to-nominal ratio or bias factor for system resistance ( R / Rn or R)
and corresponding coefficient of variation (VR) for each system configuration are
presented in Table 8.9. Since the modelling uncertainty has higher COV than that of
simulated strength, the overall variability of system resistance is highly influenced by
the modelling error. The lognormal distribution is conservatively assumed as a
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 190
______________________________________________________________________

probabilistic model for the system resistance in the reliability analysis when using the
first-order reliability method (FORM).

Table 8.9: Statistics for the system resistance of support scaffold systems

Jack extensions Number of R / Rn VR


Lift height (m)
(mm) bays

1.0 100 1x1 1.35 0.15

1.0 300 1x1 1.19 0.11

1.0 600 1x1 1.09 0.10

1.0 100 3x3 1.34 0.12

1.0 300 3x3 1.21 0.10

1.0 600 3x3 1.11 0.10

1.0 100 3x6 1.37 0.12

1.0 300 3x6 1.24 0.11

1.0 600 3x6 1.15 0.10

1.0 100 9x9 1.33 0.12

1.0 300 9x9 1.29 0.10

1.0 600 9x9 1.11 0.10

1.5 100 1x1 1.19 0.12

1.5 300 1x1 1.11 0.11

1.5 600 1x1 1.08 0.11

1.5 100 3x3 1.21 0.12

1.5 300 3x3 1.08 0.10

1.5 600 3x3 1.04 0.10

1.5 100 3x6 1.19 0.11

1.5 300 3x6 1.14 0.10

1.5 600 3x6 1.05 0.10

1.5 100 9x9 1.19 0.11


Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 191
______________________________________________________________________

Jack extensions Number of R / Rn VR


Lift height (m)
(mm) bays

1.5 300 9x9 1.08 0.10

1.5 600 9x9 1.03 0.10

2.0 100 1x1 1.14 0.12

2.0 300 1x1 1.13 0.12

2.0 600 1x1 1.09 0.11

2.0 100 3x3 1.13 0.12

2.0 300 3x3 1.11 0.11

2.0 600 3x3 1.05 0.10

2.0 100 3x6 1.13 0.11

2.0 300 3x6 1.12 0.11

2.0 600 3x6 1.04 0.10

2.0 100 9x9 1.12 0.11

2.0 300 9x9 1.10 0.11

2.0 600 9x9 1.03 0.10

8.5 Reliability Analysis with Design by Advanced Analysis

The basic structural reliability concept is to find the probability of failure (Pf) of the
structure, as defined by:

Pf P( R Q 0) P(G ( R, Q) 0) FR ( x) f Q ( x)dx (8.2)

in which R is the structural system resistance or system capacity and Q is the total load
effect. Both R and Q are modelled as random variables. G(R, Q) represents the limit
state function and G(R,Q) 0 defines the unsafe or failure region. FR(x) is the
cumulative distribution function of R, and fQ(x) is the probability density function of Q.
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 192
______________________________________________________________________

For practical use in the design by advanced analysis, Eq. (8.2) is complemented by the
familiar LRFD formula, which when applied to a structural system, takes the form:

system Rn i Qi (8.3)

in which Rn is the nominal system resistance determined according to the design


procedure (here advanced analysis), Qi are the nominal design load effects, system is the
system resistance factor, and i are the load factors. Since the load factors, i, are known
and can be set equal to the values in the design standard, then the system resistance
factor, system, is determined using reliability analysis (FOSM or FORM) to achieve a
specific system target reliability index.

In reliability analysis by FOSM, the random variables are assumed to be all normally
distributed and uncorrelated. A closed-form FOSM expression for the reliability index
derived in terms of mean-to-nominal values (bias factors) and variances of the random
variables, considering only gravity load combination (dead and live loads), can be
written as (Melchers 1999):

R L L
D L n Dn D Dn L n Dn
G system Dn Dn
(8.4)
G R
2
L
2
L
D L n Dn VR D Dn VD 2 L n Dn V L
Dn
system Dn

in which G is the mean value of the limit state function, G is the standard deviation of
the limit state function, R is the bias factor for system resistance, D is the bias factor for
dead load, L is the bias factor for live load, D is the load factor for dead load, L is the
load factor for live load, system is the system resistance factor, VR is the coefficient of
variation of system resistance, VD is the coefficient of variation of dead load, VL is the
coefficient of variation of live load, Dn is the nominal dead load, and Ln is the nominal
live load. From Eq. (8.4), the system resistance factor system can be solved for any
specified system target reliability when all other parameters are known.
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 193
______________________________________________________________________

When the actual distributions of the random variables are taken into account, including
distributions which are non-normal, the reliability index can be more accurately
approximated using FORM. The basic theory is presented in Section 2.6.6. The FORM
analysis is performed in iterative manner. In order to carry out code calibration to obtain
system with this method, the following step-by-step procedure is used. Again, only the
gravity load combination (dead and live loads) is considered in this research.

(a) Formulate the limit state function G R D L where R is system resistance,


D is dead load effect, and L is live load effect, and determine the probability
distributions with appropriate parameters for all random variables (R, VR, D, VD,
L, and VL). For example, assume R is lognormal, D is normal, L is type I
extreme value distribution, and L/D = 0.5, (however, when Ln/Dn = 0.5, L/D =
0.4762 is used in Section 8.5.2, according to the load statistics in Section 8.5.1).

(b) Obtain an initial design point by assuming d* = D and then l* = L = 0.5D. Set
G = 0, r* = 1.5D then using G = 0 evaluate at the mean values, R = 1.5D.

(c) Determine the equivalent normal parameters for R ( Re and Re ) and L ( Le and

Le ) only since D is assumed normal.

(d) Calculate the partial derivatives of the limit state function with respect to the
reduced variates and define a column vector Gpartial (see Eq. (2.22)).

(e) Calculate a column vector of the sensitivity factors for R, D and L (see Eq.
(2.24)).

(f) Choose a value for the target reliability index .

(g) Find a new design point in reduced variates for n 1 = 2 variables. Choosing the
variables for D and L, then z D* D and z L* L (from Eq. (2.25)).
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 194
______________________________________________________________________

(h) Compute the values of corresponding design point in original coordinates for D
and L, then d * D z *D D and l * Le z L* Le (from Eq. (2.26)).

(i) Solve for the value of the remaining random variable R by setting the limit state
function G = 0, then r * d * l * and update the mean value of R
r*
as R (derived from Eq. (2.25) and Eq. (2.26)).
1 R VR

(j) Repeat steps (b) to (i) with successive new design points until the design points
converge.

(k) Choose load factors: D and L.

D
D L L
(l) Compute system resistance factor as system D L from Eq.
R
R
(8.3).

More information on the code calibration procedure by FOSM and FORM can be found
in Melchers (1999), and Nowak and Collins (2000).

8.5.1 Load Statistics

Steel support scaffold systems are normally designed to withstand both gravity and
wind loads. In the case of wind loads, the top of the support scaffold systems consists of
formwork that is usually tied or braced to permanent structures such as lift cores and
columns, thus transferring most of the lateral loads into the permanent structures. Also,
some of the lateral loads are resisted by the braces on the scaffold systems. The design
of support scaffold systems is therefore normally governed by gravity loads from
formwork, steel reinforcement, wet concrete, equipment, workers, and stacked materials.
In this research, only the gravity load combination of live and dead loads is considered
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 195
______________________________________________________________________

as 1.2Dn+1.6Ln and 1.25Dn+1.5Ln, where Dn is nominal dead load and Ln is nominal live
load, based on the minimum design loads proposed by ASCE (2005) and Standard
Australia (1995), respectively. The construction dead loads generally consist of the self-
weight of concrete, reinforcement, and formwork. In this research, the probability
distribution of the construction dead loads is considered to be comparable to that of
occupancy dead loads, proposed by Ellingwood et al. (1980), i.e. normal distribution
with a mean-to-nominal value ( D / Dn or D) of 1.05 and a coefficient of variation (VD)
of 0.1.

The construction live loads normally include the weight of labourers, equipment, and
stacked materials. The probabilistic data on construction live loads is limited and
dependent on the stage of construction; moreover, the fitted distribution can be complex.
Karshenas and Ayoub (1994b) reported statistical data of the surveyed live loads
conducted before concrete placement in the period of 1987 to 1990 on over twenty two
concrete building sites in the Unites States. They recommended a mixed probability
distribution consisting of a discrete part at zero loads and a continuous exponential
distribution at nonzero loads. The same researchers (Karshenas and Ayoub 1994a) also
studied construction live loads on slab formworks after concrete placement. The mean
of the construction live loads on newly poured slabs, (surveyed 1 week after the pour),
was calculated to be 0.3 kPa with a very high coefficient of variation of 1.10. The result
was later used in the research by Rosowsky and Stewart (2001) to simulate the
distribution of maximum construction live loads over the period of 6 months using a
Monte Carlo method. A type I extreme value distribution was found to give the best fit
to the simulated distributions of measured construction live loads. Also, Rosowsky
(1996) reported on the development of the new ASCE standard on load combinations
and load factors for construction. Rosowsky implied that since there was little data
available on probabilistic models for construction loads, the code had been developed
based on conservative judgment and consensus among engineers. Ellingwood and
Galambos (1982) reported that statistical data on live loads for buildings and
conventional structures may be modelled as a type I extreme value distribution with
L / Ln 1.0 and a coefficient of variation (VL) of 0.25. In this research, the probability
distribution of the construction live loads is assumed to be type I extreme value
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 196
______________________________________________________________________

distribution with a mean-to-nominal value ( L / Ln or L) of 1.0 and a conservative


coefficient of variation (VL) of 0.6. Table 8.10 summarises the construction load
statistics used in this research.

Table 8.10: Statistics for construction loads

Mean-to-nominal Coefficient of
Load Probability distribution
value variation
Dead 1.05 0.1 Normal

Live 1.0 0.6 Type I extreme value

8.5.2 System Resistance Factors

Based on the LRFD design formulae of systemRn = 1.2Dn + 1.6Ln (ASCE 2005) and
systemRn = 1.25Dn + 1.5Ln (Standard Australia 1995), the system resistance factors
system can be computed for various system configurations using the code calibration
procedure discussed in Section 8.5. The nominal live-to-dead load ratio (Ln/Dn) is
assumed to be 0.5. Typical Ln/Dn ratios for reinforced concrete structures are in the
range of 0.5-1.5 (Ellingwood et al. 1980), and the design of support scaffold systems is
usually critical in the stage of concrete pouring as suggested by Hadipriono and Wang
(1987). An Ln/Dn ratio of 0.5 is a reasonably representative value since the construction
live loads such as the weight of labourers, small equipment, and mounding of concrete
are relatively small compared to the dead loads from the weight of wet concrete, steel
reinforcement, and formwork. Other probabilistic parameters for the code calibration
have been presented earlier in Sections 8.4.3 and 8.5.1. Because the Ln/Dn ratio is
assumed to be 0.5, both LRFD design formulae herein give identical results of system
resistance factor through a reliability analysis, which can be verified by Eq. (8.4) and
exemplified in Figure 8.46. Since the mean-to-nominal ratios and coefficients of
variation of the system resistances of different system sizes (1x1, 3x3, 3x6, and 9x9
bays) with the same lift height and jack extensions (see Table 8.9) are comparable, then
the system resistance factors can be categorised based on the lift height and top and
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 197
______________________________________________________________________

bottom jack extensions. The results of the system resistance factor for a system
reliability index in the range of 3.0-5.0, as obtained from an FORM reliability analysis
for different system configurations, are presented in Figures 8.47-8.49.

4.50

4.00
Reliability index

3.50

3.00

2.50

2.00
0 0.5 1 1.5 2
L n /D n

Standard Australia (1995) ASCE (2005)

Figure 8.46: Comparison between Australian and ASCE design formulae based on
Ln/Dn

0.90
100 mm jack extensions
0.80 300 mm jack extensions
System resistance factor

600 mm jack extensions


0.70

0.60

0.50

0.40

0.30
3.0 3.5 4.0 4.5 5.0
Reliability index

Figure 8.47: System resistance factor for 1.0 m lift scaffold system
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 198
______________________________________________________________________

0.90
100 mm jack extensions
0.80 300 mm jack extensions
System resistance factor 600 mm jack extensions
0.70

0.60

0.50

0.40

0.30
3.0 3.5 4.0 4.5 5.0
Reliability index

Figure 8.48: System resistance factor for 1.5 m lift scaffold system

0.90
100 mm jack extensions
0.80 300 mm jack extensions
System resistance factor

600 mm jack extensions


0.70

0.60

0.50

0.40

0.30
3.0 3.5 4.0 4.5 5.0
Reliability index

Figure 8.49: System resistance factor for 2.0 m lift scaffold system

It appears from the figures that the system resistance factor system can be as low as 0.37
if the system target reliability index is chosen as 5.0, and as high as 0.81 if the system
target reliability index is chosen as 3.0. In addition, longer jack extensions and higher
lift heights have lower system resistance factors for the same system target reliability
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 199
______________________________________________________________________

index. In terms of system reliability index, some researchers have recommended values
in the range of 2.5 to 4 depending on the risk of failure and type of structures (see Table
2.1). In this research, the system target reliability of 3.5 is deemed appropriate as
recommended by Galambos (1990) for a frame suffering complete damage (no expected
loss of life) when buckling under the action of gravity loads with little assistance from
claddings. Table 8.11 presents the system resistance factors system, determined for the
system target reliability index of 3.5 for support scaffold systems and categorised by the
lift height and top and bottom jack extensions. The mean-to-nominal ratio for system
resistance ( R / Rn ) and corresponding coefficient of variation (VR) for each system
configuration are also shown. The load statistics are as described in Section 8.5.1. The
values of system resistance factor are presented for both FOSM and FORM calibration
procedures; however, the system resistance factor system determined by the FORM is
recommended in view of enhanced accuracy.

Table 8.11: System resistance factors for support scaffold systems ( = 3.5)

Jack
Lift height system system
extensions R / Rn VR
(m) (FOSM) (FORM)
(mm)
1.0 100 1.35 0.13 0.78 0.70
1.0 300 1.23 0.11 0.77 0.66
1.0 600 1.12 0.10 0.72 0.61
1.5 100 1.20 0.12 0.72 0.63
1.5 300 1.10 0.10 0.71 0.60
1.5 600 1.05 0.10 0.68 0.57
2.0 100 1.13 0.11 0.70 0.61
2.0 300 1.11 0.11 0.69 0.60
2.0 600 1.05 0.11 0.65 0.56

In all cases, the system resistance factors based on FORM are shown to be smaller than
those based on FOSM because the probability distributions assumed herein are more
conservative than those (normal distributions) of the FOSM. To calculate the system
resistance factor for other jack extensions, linear interpolation can be applied, i.e. system
Chapter 8 - Probabilistic Assessment of Support Scaffold Systems 200
______________________________________________________________________

= 0.66 + (400 300) (0.61 0.66) / (600 300) = 0.64 for 1.0 m lift height with 400
mm jack extensions. In the design of support scaffold systems, it might not be possible
to specify the same jack extension for all standards since the ground or formwork might
not be level. In such situations, the system resistance factor system can be based on only
the lift height by averaging the values for different jack extensions such that for 1.0 m
lift, system = 0.66, for 1.5 m lift, system = 0.60, and for 2.0 m lift, system = 0.59.

8.6 Conclusions

This chapter studies the effects of uncertainties in material properties, initial geometric
imperfections, and joint stiffness on the ultimate strength of steel support scaffold
systems through a rational statistical framework and a second-order inelastic finite
element (advanced) analysis. The chapter also presents a reliability analysis of support
scaffold systems to determine system resistance factors that can be used in the direct
design by advanced analysis according to the LRFD framework. Having procured the
statistical data for the main random variables affecting the strength of support scaffold
systems in previous chapters, Monte Carlo simulations are carried out to determine the
statistical distributions of strength for a range of configurations of support scaffold
systems. The load statistics are considered for the gravity load combination. The first-
order second-moment method (FOSM) and first-order reliability method (FORM) are
then employed for determining the system resistance factors for support scaffold
systems in various system configurations. Finally, the system resistance factors, system,
determined by FORM for a system target reliability index of 3.5 and categorised by the
lift height and jack extensions, are proposed to be used in the LRFD design of support
scaffold systems by advanced analysis (Chapter 9). The proposed system resistance
factors vary from 0.56 to 0.70, depending on the lift height and jack extensions.
Chapter 9 - Design of Support Scaffold Systems 201
______________________________________________________________________

CHAPTER 9 - DESIGN OF SUPPORT SCAFFOLD


SYSTEMS

9.1 Introduction

Current best industry practice in the design of steel support scaffold systems is to use
load capacity design charts recommended by the manufacturers, which are based on a
number of scaffold load tests under some standardised conditions, and then apply a
judgmental factor of safety which is typically around 2.0-3.0. Also, a member-based
allowable stress design (ASD) method has been used in the design of steel support
scaffold systems, as shown in the code of practice for falsework from the British
Standards Institution (1996). An alternative design method is the member-based load
and resistance factor design (LRFD) approach, as included in AISC (2005) and
Standard Australia (1998). In recent years, with the availability of powerful computers
and finite element software packages, design by advanced analysis has become feasible.
The design by advanced analysis is allowed in many structural design codes, such as in
AISC (2005) and Standard Australia (1998). In the literature (Chan et al. 2005 and
Buonopane et al. 2003), some researchers have already proposed the use of advanced
analysis design method for steel structures. Nonetheless, there is a need to determine the
system resistance factors in the LRFD framework to be used with advanced analysis for
different types of structures. In Chapter 8 of this thesis, the system resistance factors
have been proposed to be used for the design by advanced analysis of support scaffold
systems.

This chapter provides a structural design guideline for the LRFD-based advanced
analysis method for support scaffold systems. The chapter gives recommendations for
how to create practical three-dimensional nonlinear finite element models of support
scaffold systems to use with the design by advanced analysis. A step-by-step structural
analysis and design procedures for support scaffold systems are presented. General
comparisons of the capacities obtained by advanced analysis design and by an existing
Chapter 9 - Design of Support Scaffold Systems 202
______________________________________________________________________

member-based LRFD design from an industry design chart are presented. Two design
case studies are used to demonstrate the practicality of the direct design by advanced
analysis of support scaffold systems. Without any tedious member and connection
capacity checks and the exclusion of member effective length calculations, this design
method becomes an attractive tool for the design of support scaffold systems.

9.2 Practical Advanced Analysis Models

The main attributes that must be considered in creating practical three-dimensional


finite element models for the advanced analysis of support scaffold systems include
initial geometric imperfections, inelastic material modelling, joint stiffness, loading
eccentricity, and eccentricity at supports. The analysis has to consider geometric 2nd
order effects associated with member P- and frame P- effects as well as material
nonlinearity. In order to promote the use of advanced analysis for the design of support
scaffold systems among professional engineers, the following subsections provide
recommendations for practical three-dimensional finite element modelling, which is
identical to the nominal model defined in Section 8.4.2, but explained in more detail.
Figure 9.1 summarises the main key features of the nominal design model and Figure
9.2 shows a typical three-dimensional advanced analysis model. It should be noticed
that the braces are intentionally left out in Figure 9.1 for a clearer view.

40 kNm/rad y -axis bending Loading eccentricity (19 mm)


Out- of-plumb (H/600)

Out-of-straightness (Lh/700)
Lh Spigot model
Tri-linear joint model

Lh Out-of-plumb

Base plate model (15 mm eccentricity)

Figure 9.1: Idealised nominal design model (elevation view)


Chapter 9 - Design of Support Scaffold Systems 203
______________________________________________________________________

Figure 9.2: Typical three-dimensional finite element (advanced) analysis model

9.2.1 Initial Geometric Imperfections

The initial geometric imperfections of a scaffold frame can be modelled explicitly by


changing the coordinates of the nodes of the finite element model according to known
imperfection values. The member out-of-straightness as well as storey out-of-plumb
should be applied to the standards; however, only the storey out-of plumb is required for
the jacks since the jacks are made of relatively short and stocky members, and so the
effect from the member out-of-straightness is negligible. It should be noticed that the
storey out of plumb has insignificant effect on the system strength because the frame P-
effect is reduced by the braces and the horizontal supports at the top. The member
out-of-straightness is applied directly at the mid height of the standard in each scaffold
lift and the frame out-of-plumb () is applied at each ledger-standard connection point
and at the U-head at the top of the scaffold. Ideally, many subdivided elements might be
preferred for the member out-of-straightness so that the imperfections follow the
sinusoidal shape; however, it has been found that only two elements with a maximum
Chapter 9 - Design of Support Scaffold Systems 204
______________________________________________________________________

deflection at the mid-height of a member are sufficiently reflecting the imperfection


effects (Chen and Kim 1997). The pattern of geometric imperfections and loading
eccentricity is set to be at the worst case scenario, as shown in Figure 9.1. The storey
out-of-plumb of H/600 and the member out-of-straightness of Lh/700, where H is the
height of the system and Lh is the lift height, are proposed for the nominal design model.
These imperfection values are approximately the mean values of those obtained from
the procured on-site data presented in Chapter 3. Since the standards of support scaffold
systems can contain spigot joints, which increase member out-of-straightness, as
discussed in Chapter 3, the proposed member out-of-straightness value is higher than
the value of Lh/1000 commonly imposed by steel design codes (Standard Australia 1998
and AISC 2005).

The nominal design models also include spigot joint modelling, as illustrated in Figure
9.1. There are no spigots in the bottom lift based on general construction practice. The
spigot joints are assumed to be located at the mid-height of the second and higher lifts,
depending on the design. The spigot joint model is discussed in Section 6.2.1. Figure
9.3 shows an illustration of a spigot model in Strand7 (2009).

Figure 9.3: Illustration of a spigot model in Strand7


Chapter 9 - Design of Support Scaffold Systems 205
______________________________________________________________________

9.2.2 Loading Eccentricity

The loading eccentricity that occurs between the timber bearer and the U-Head is
modelled by a rigid link with length equal to the loading eccentricity, connected to the
top of the jack in the direction perpendicular to the bearer. A vertical point load is
applied at the far end of the link (Figure 9.1). The rigid link behaves as a short, stiff
cantilever that introduces vertical force and additional moment into the jack. The
loading eccentricity of 19 mm, which is close to the mean loading eccentricity obtained
from the collected on-site data (Chapter 3), is proposed for the nominal design models
and applied to each U-head screw jack in the same direction. Figure 9.4 shows an
illustration of a loading eccentricity on a screw jack in Strand7 (2009). Eccentricity at
the base may occur as a result of uneven or sloped ground, and is included in the
nominal design models by using the base plate model, described in Section 6.2.4. Since
there is no statistical data for the base eccentricity available in literature and acquisition
of such data is difficult and subjective, the bottom eccentricity is assumed to be 15 mm,
based on the eccentricity limit stated in AS 3610 (Standard Australia 1995). The
eccentricity is applied to each upright in the same direction. The base plate model is
illustrated in Figure 9.5. More details on the base plate model are presented in Section
6.2.4.

Figure 9.4: Illustration of a loading eccentricity in Strand7


Chapter 9 - Design of Support Scaffold Systems 206
______________________________________________________________________

Figure 9.5: Illustration of a base plate model in Strand7

9.2.3 Joint Stiffness

The semi-rigid joints of support scaffold systems are modelled accurately by applying
the tri-linear joint model developed in Chapter 4. The model caters for the large
rotations and plastic deformations that may develop near the ultimate load, as well as
joint looseness. The idealised tri-linear joint model is presented as a moment-rotation
curve in Figure 9.6. The slope k1 represents the looseness of the joint, while the slopes
k2 and k3 represent elastic and plastic joint stiffness, respectively. For the Cuplok
connections studied in this research, the proposed parameters defining the moment-
rotation curves for various joint configurations and bending axes are summarised in
Tables 9.1-9.2. The joints can be modelled by a connection element available in Strand7
(2009) or other finite element software packages where stiffness for any of the six
degrees of freedom (axial, shear in 2 directions, and bending about 3 axes) can be
defined. In the present model, only bending stiffness in the vertical and horizontal
planes is incorporated and the remaining degrees of freedom are assumed to provide
rigid connections between standard and ledger.
Chapter 9 - Design of Support Scaffold Systems 207
______________________________________________________________________

Figure 9.6: Idealised tri-linear joint model

Table 9.1: Cuplok joint stiffness (kNm/rad) for nominal design model

Bending stiffness about horizontal Bending stiffness about vertical


axis (kNm/rad) axis (kNm/rad)
Joint
k1 k2 k3 k1 k2 k3
configuration

4-way 39 102 5.3 15 7.5 0.8

3-way 36 87 5.1 14 7 1

2-way 41 77 4.6 7.5 5 1.5

Table 9.2: Rotation for Cuplok joints (rad) for nominal design model

Rotation about horizontal axis Rotation about vertical axis


(rad) (rad)
Joint
1 2 3 1 2 3
configuration

4-way 0.014 0.036 0.16 0.02 0.04 0.1

3-way 0.012 0.036 0.16 0.02 0.04 0.1

2-way 0.007 0.036 0.16 0.02 0.04 0.1


Chapter 9 - Design of Support Scaffold Systems 208
______________________________________________________________________

In this research, the braces are made of telescopic members with hooks at the ends
(Boral Formwork & Scaffolding Pty. Ltd. 2002). They are modelled using two rigidly
connected elements with different cross-sections. Connection elements with only axial
stiffness form the connection between the brace member elements and the ledgers. The
axial spring stiffness is taken as 1.8 kN/mm as obtained from test calibrations on braced
scaffold systems, as described in Chapter 6. The braces are offset 60 mm from the nodal
points between the standards and ledgers, representing actual construction practice.

9.2.4 Boundary Conditions

The top boundary conditions of the advanced analysis model represent the stiffness
between the supported timber bearer and the U-head screw jack. The rotation stiffness
about the x-axis is assumed to be rigid, corresponding to the negligible strong axis
bending of the supported timber bearer; however, the y-axis bending stiffness (Figure
9.2) is taken as 40 kNm/rad as obtained from calibrations to the full-scale subassembly
tests, described in Chapter 5. Also, the model is assumed to be braced laterally at the top
of the U-head screw jacks (since the formwork is usually tied to a permanent structure),
and hence, the translational stiffness is assumed to be rigid in the x and y directions, but
0 in the z direction, based on the axes shown in Figure 9.2. The bottom boundary
conditions consist of the base plate model as discussed previously in Section 9.2.2.

9.2.5 Material Model

The proposed nominal material model for the steel scaffold components is an elastic-
perfectly plastic model with a yield stress corresponding to the nominal yield stress
(grade of steel) of each component, as shown in Figure 9.7. The nominal yield stress of
each scaffold component used in this research is presented in Table 9.3. The proposed
nominal material model is considered to strike an appropriate balance between accuracy
and practicality for use in design by advanced analysis. The effects of residual stresses
are taken into account by the system resistance factor incorporated into the design
procedure. Such effects have already been considered implicitly by using the Ramberg-
Chapter 9 - Design of Support Scaffold Systems 209
______________________________________________________________________

Osgood expression for the material model in performing Monte Carlo simulations to
obtain system strength statistics and system resistance factors, as presented in Chapter 8.
In this research, a nonlinear beam element in Strand7 (2009) is used to incorporate the
nonlinear material effects for all the scaffold components, using the concept of plastic-
zone analysis (Clarke et al. 1992).

fy fy is the nominal yield stress.

E is the Youngs modulus.


E
1

Figure 9.7: Elastic-perfectly plastic material model for nominal design

Table 9.3: Nominal yield stress of steel scaffold components in this research

Component Nominal yield stress

Standard 450 MPa

Jack 430 MPa

Spigot 400 MPa

Ledger 350 MPa

Brace 400 MPa

Base plate 250 MPa


Chapter 9 - Design of Support Scaffold Systems 210
______________________________________________________________________

9.3 Advanced Analysis and Design Procedures

The ultimate limit state design procedure for support scaffold systems starts with
determining the factored loads and creating the nominal three-dimensional nonlinear
finite element model, which includes initial geometric imperfections. The advanced
structural analysis is performed on the model by incrementing the applied loads until the
ultimate capacity of the system (Rn) is reached. The adequacy of the structural system to
support the applied ultimate loads can then be readily evaluated by comparing the
system ultimate strength systemRn with the applied factored loads iQi. From Eq. (8.3)
of Chapter 8, the design is satisfactory if iQi systemRn. It should be noticed that Rn is
the nominal system strength determined by advanced analysis and system is the system
resistance factor, which may be based on the required system reliability index ( = 3.5
herein) and system configuration, as presented in Table 8.11. If the design is
unsatisfactory, the system can be redesigned and the procedure reapplied. The step-by-
step details of the design procedure for support scaffold systems by advanced analysis
are presented in the following subsections.

9.3.1 Load Combination

The dead and live load (gravity) combination is usually critical in the design of support
scaffold systems. According to the minimum design loads proposed by ASCE (2005),
the combination is 1.2Dn+1.6Ln, where Dn is nominal dead load and Ln is nominal live
load. The dead loads generally include the self-weight of concrete, reinforcement, and
formwork. The construction live loads normally include the weight of labourers,
equipment, and stacked materials. Three stages of construction, i.e. prior to placement
of concrete, during placement, and after placement, should be considered in the design;
however, the most critical stage usually occurs during the placement of concrete due to
higher vertical loads. The calculations of the vertical loads are usually based on a
tributary area for each upright.
Chapter 9 - Design of Support Scaffold Systems 211
______________________________________________________________________

9.3.2 Preliminary Design of Support Scaffold Systems

This step of the design is to select an appropriate system configuration to support the
formwork according to the construction plan. Since the components of support scaffold
systems normally consist of specific pre-manufactured sizes and lengths, the task is then
to select a suitable lift height, bay width, and jack extensions to fit the construction plan.
In this research, the cross-sectional properties of scaffold components are presented in
Section 5.2. Also, it should be noticed that the lift heights are specified as 1.0 m, 1.5 m,
and 2.0 m. The bay width is chosen at 1829 mm, which is frequently used in
construction. The top and bottom jack extensions are between 100 mm and 600 mm.
Nevertheless, the design concept herein also applies to other system configurations.

9.3.3 Modelling of Elements

The proposed nonlinear finite element model of support scaffold systems includes the
modelling of geometric imperfections, loading eccentricity, spigot joints, standard-to-
ledger connections, base plates, and nonlinear material effects, as described in Section
9.2. The factored vertical loads are applied to the U-head screw jacks of the nominal
advanced analysis model as point loads with proposed eccentricity (see Section 9.2.2).

9.3.4 Determination of Load Increments

Before performing an advanced analysis using the nonlinear finite element model, the
input of incremental nominal loads needs to be determined. The load increments are
obtained by dividing the nominal factored loads by an appropriate number of
increments in the range from 20 to 50, which proves to be adequate for tracing nonlinear
load-displacement responses of support scaffold systems. The incremental load input
starts from zero and increases in steps of load amplifier (a). The applied loads may
exceed the nominal factored loads. In Strand7 (2009), an automatic scaling of the load
increments can be applied to help the convergence when changes in the element
stiffness parameter exceed a predefined convergence tolerance. The recommended
Chapter 9 - Design of Support Scaffold Systems 212
______________________________________________________________________

methods, which are available in the software package, are automatic load scaling and
displacement control (arc length method).

9.3.5 Ultimate Strength Design Check

The ultimate strength design check is carried out by performing a nonlinear (advanced)
analysis that considers both nonlinear geometric and material effects. In the analysis, the
factored nominal loads are multiplied by a load amplifier (a), i.e. when a = 1, the
applied loads equal the factored nominal loads. The design is satisfied if the maximum
value of a is larger than or equal to 1/system, where system is the system resistance factor
for support scaffold systems, derived in Chapter 8 and summarised in Table 9.4 for
various lift heights and jack extensions. To estimate the system resistance factor for
other jack extensions, linear interpolation shall be applied. It should be noticed that the
product of the maximum value of a and the nominal factored loads is referred to as the
nominal system strength Rn. There is no need for checking member and connection
strengths to a structural design standard because the strengths of members and
connections are verified in the course of running the advanced analysis.

Table 9.4: Summary of system resistance factors for support scaffold systems ( = 3.5)

Jack
Lift height system
extensions
(m) (FORM)
(mm)
1.0 100 0.70
1.0 300 0.66
1.0 600 0.61
1.5 100 0.63
1.5 300 0.60
1.5 600 0.57
2.0 100 0.61
2.0 300 0.60
2.0 600 0.56
Chapter 9 - Design of Support Scaffold Systems 213
______________________________________________________________________

9.3.6 Adjustment of System Configuration

According to the previous step, if the design is unsatisfactory (i.e. a < 1/system), the
system configuration can be adjusted and the same procedure is reapplied. In the case
that the maximum value of a is found to be much larger than 1/system, the adjustment of
system configuration, i.e. an increase in lift height or bay width, is plausible so as to
minimise the amount of material used to provide the required system strength.

9.4 Design Comparisons

The design strengths of support scaffold systems obtained by advanced analysis are
compared with strengths obtained using design charts, published by Boral Formwork &
Scaffolding Pty. Ltd. (2002). The design chart strengths have been determined
according to the Australian steel structures standard AS 4100 (Standard Australia 1998)
in combination with a proprietary design model of a single standard. The system studied
in this thesis is identical to the Supercuplok support system produced by Boral
Formwork & Scaffolding Pty. Ltd. The values of the system ultimate strength systemRn
obtained by advanced analysis are compared with the ultimate strengths per upright,
extracted from the Boral design charts for different lift heights and the top and bottom
jack extensions of 300 mm. In using the design charts, the top and bottom loading
eccentricities are chosen to be 19 mm and 15 mm, respectively, corresponding to those
of the advanced analysis design models. The results are presented in Table 9.5.

The percentage differences shown in Table 9.5 indicate that the design by advanced
analysis of support scaffold systems results in approximately 13% higher ultimate
strengths than the published design chart values. It should be noted that the system
resistance factor applied in the design by advanced analysis of support scaffold systems
is based on the system target reliability of 3.5. Obviously, if the system target reliability
is chosen to be higher than 3.5, the system resistance factor (system) would be lower and
the percentage difference between the two design methods would be closer. In this
research, the system target reliability of 3.5 is considered to be appropriate in view of
the fact that the differences in the ultimate load capacity resulting from the two design
Chapter 9 - Design of Support Scaffold Systems 214
______________________________________________________________________

methods are not substantial and support scaffold systems are only used as temporary
structures in construction.

Table 9.5: Design comparisons between advanced analysis and member-based design to
AS 4100 (Boral design charts)

Ultimate
Lift Jack strength Ultimate
from strength
height extensions system Difference
Rn (kN) advanced from Boral
(%)
(m) (mm) analysis design
systemRn charts (kN)
(kN)

1.0 300 0.66 126.3 83.3 76.4 9%

1.5 300 0.60 106.0 63.6 54.2 17%

2.0 300 0.60 77.0 46.2 40.9 13%

9.5 Design Examples

Two design cases are presented to demonstrate the practicality of designing support
scaffold (Cuplok) systems by advanced analysis. Case study 1 concerns the design of a
support scaffold system for a typical reinforced concrete flat slab in a residential
building with a clear floor-to-floor height of 3.80 m. The slab thickness is taken to be
250 mm. The required area to be supported by formwork is 7.0 m x 7.0 m. Case study 2
presents the design of a support scaffold system for a reinforced concrete rectangular
box girder of a bridge section, spanning 7.0 m at the clear height of 6.8 m. This design
case presents a scaffold system with numerous lifts that support heavy loadings. The
step-by-step details of the design procedure are presented for both case studies in the
following subsections. The dimensions and material properties of scaffold components
are as described in Section 5.2.
Chapter 9 - Design of Support Scaffold Systems 215
______________________________________________________________________

9.5.1 Case Study 1

(a) The problem is to design a support scaffold system that supports a reinforced
concrete flat slab of 250 mm thickness, which covers an area of 7.0 m x 7.0 m in a
residential building. The clear floor-to-floor height is 3.80 m. The load combination of
dead and live loads is considered as 1.2Dn+1.6Ln, where Dn is nominal dead load and Ln
is nominal live load. The critical second stage of construction, occurring during concrete
placement, is being considered for the design.

(b) Load calculations

The calculations of the vertical point load on each standard are based on a tributary area.
Dead load calculations:
Assume concrete density with reinforcement = 2500 kg/m3
Reinforced concrete slab depth = 250 mm
Estimated reinforced concrete dead load = (2500 x 10 x (250/1000))/1000 = 6.25 kPa
A tributary area per upright based on 1.829 m x 1.829 m (Boral) bays = 3.345 m2
Estimated reinforced concrete dead load per upright = 6.25 x 3.345 = 20.9 kN
Estimated wooden formwork load = 0.3 kPa
Estimated wooden formwork load per upright = 0.3 x 3.345 = 1 kN
Total dead loads per upright = 20.9 + 1 = 21.9 kN
Factored nominal dead loads = 1.2 x 21.9 26 kN
Live load calculations:
Weight of 10 workers in the area = 10 kN
Weight of vibrators, screed, shovels, etc. = 2 kN
Weight of pump line (say about 60 kg/m x 10 m) = 6 kN
Total weight in the area = 10 + 2 + 6 = 18 kN
Average load over working area = 18/(7 x 7) = 0.37 kPa
Additional estimated load from mounding of the wet concrete = 3 kPa
A tributary area per upright based on 1.829 m x 1.829 m bays = 3.345 m2
Total live loads per upright = (0.37 + 3) x 3.345 = 11.3 kN
Factored nominal live loads = 1.6 x 11.3 18 kN
Chapter 9 - Design of Support Scaffold Systems 216
______________________________________________________________________

(c) Preliminary design

Clear floor-to-floor height = 3.80 m


Required height from the base plate up to the U-head (excluding primary and secondary
wooden bearers and wooden formwork) = 3.80 0.15 0.15 0.017 = 3.483 m 3.5 m
Required area (7.0 m x 7.0 m) = 49 m2
Choose a suitable system configuration as 3 x 3 bays (1.829 m equal bay width), 1.5 m
lift height, 2 lifts, and 100 mm top and bottom jack extensions (including 190 mm top
sleeve and 110 mm bottom sleeve).

(d) Finite element modelling

The three-dimensional finite element model for the chosen system configuration is
created, as shown in Figure 9.8. The spigot joints are located in the 2nd lift.

Figure 9.8: Advanced analysis model for case study 1


Chapter 9 - Design of Support Scaffold Systems 217
______________________________________________________________________

(e) Determination of load increments

The incremental load input starts from zero and increases by increasing the load
amplifier (a) in steps of 0.05.

(f) Ultimate strength design check

The advanced analysis is performed on the model. The load-deflection response at the
mid height of the 2nd lift is shown in Figure 9.9. The applied load is presented as the
load amplifier (a) of the factored nominal loads. The maximum value of a is found to
be 3.05, which is larger than 1/system = 1/0.63 = 1.6, where system is the proposed
system resistance factor (see Table 9.4). Therefore, the chosen system configuration is
satisfied by the ultimate strength design check.

3.50

3.00

2.50
Load amplifier

2.00

1.50

1.00

0.50

0.00
0 5 10 15 20 25 30
Deflection (mm)

Figure 9.9: Load-deflection response at the mid-height of the 2nd lift for case study 1
Chapter 9 - Design of Support Scaffold Systems 218
______________________________________________________________________

(g) Adjustment of system configuration

In this design case, the maximum value of a of 3.05 is substantially larger than the
1/system value of 1.6, and so it might be possible to refine the design. However, the
adjustment of system configuration is not possible because the 2.0 m lift system would
not fit the required height. Consequently, the system configuration of 3 x 3 bays (1.829
m equal bay width), 1.5 m lift height, and 2 lifts with 100 mm top and bottom jack
extensions (including 190 mm top sleeve and 110 mm bottom sleeve) is suitable.

9.5.2 Case Study 2

(a) The problem is to design a support scaffold system that supports a reinforced
concrete rectangular box girder of a bridge section, spanning 7.0 m at the clear height of
6.8 m. The cross section of the girder is shown in Figure 9.10. The load combination of
dead and live loads is considered as 1.2Dn+1.6Ln, where Dn is nominal dead load and Ln
is nominal live load. The critical second stage of construction, occurring during concrete
placement, is being considered for the design.

0.25 m

0.25 m 0.25 m 2.5 m

0.25 m

6.0 m

Figure 9.10: Cross section of the box girder for case study 2
Chapter 9 - Design of Support Scaffold Systems 219
______________________________________________________________________

(b) Load calculations

The calculations of the vertical point load on each standard are based on a tributary area.
Dead load calculations:
Assume concrete density with reinforcement = 2500 kg/m3
A tributary area per upright based on 1.829 m x 1.829 m (Boral) bays = 3.345 m2
Reinforced concrete volume supported by one upright (worst case) = (0.25 x 1.829 x 2 x
1.829) + (0.25 x (2.5-0.5) x 1.829) = 2.587 m3
Estimated reinforced concrete dead load per upright = (2.587 x 2500 x 0.01) = 64.68 kN
Estimated wooden formwork load = 0.5 kPa
Estimated wooden formwork load per upright = 0.5 x 3.345 = 1.67 kN
Total dead loads per upright = 64.68 + 1.67 = 66.35 kN
Factored nominal dead loads = 1.2 x 66.35 = 79.6 kN
Live load calculations:
Estimated live load = 3 kPa
A tributary area per upright based on 1.829 m x 1.829 m bays = 3.345 m2
Total live loads per upright = 3 x 3.345 = 10 kN
Factored nominal live loads = 1.6 x 10 = 16 kN

(c) Preliminary design

Clear height = 6.80 m


Required height from the base plate up to the U-head (excluding primary and secondary
wooden bearers and wooden formwork) = 6.80 0.15 0.15 0.017 = 6.483 m 6.5 m
Required area (6.0 m x 7.0 m) = 42 m2
Choose a suitable system configuration as 3 x 3 bays (1.829 m equal bay width), 1.0 m
lift height, 6 lifts, and 100 mm top and bottom jack extensions (including 190 mm top
sleeve and 110 mm bottom sleeve).

(d) Finite element modelling

The three-dimensional finite element model for the chosen system configuration is
created, as shown in Figure 9.11. The spigot joints are located in the 2nd, 4th, and 6th lifts.
Chapter 9 - Design of Support Scaffold Systems 220
______________________________________________________________________

Figure 9.11: Advanced analysis model for case study 2

(e) Determination of load increments

The incremental load input starts from zero and increases by increasing the load
amplifier (a) in steps of 0.05.

(f) Ultimate strength design check

The advanced analysis is performed on the model. The load-deflection response at the
mid height of the 6th lift is shown in Figure 9.12. The applied load is presented as the
load amplifier (a) of the factored nominal loads. The maximum value of a is found to
be 1.43, which is equal to the required value of 1/system = 1/0.7 = 1.43, where system is
the proposed system resistance factor (see Table 9.4). Therefore, the chosen system
configuration is satisfied by the ultimate strength design check.
Chapter 9 - Design of Support Scaffold Systems 221
______________________________________________________________________

1.60
1.40

Load amplifier 1.20


1.00
0.80
0.60
0.40
0.20
0.00
0 4 8 12 16 20 24
Deflection (mm)

Figure 9.12: Load-deflection response at the mid-height of the 6th lift for case study 2

(g) Adjustment of system configuration

In this design case, the maximum value of a equals to 1/system value of 1.43, and so the
design is optimised. Hence, the system configuration of 3 x 3 bays (1.829 m equal bay
width), 1.0 m lift height, and 6 lifts with 100 mm top and bottom jack extensions
(including 190 mm top sleeve and 110 mm bottom sleeve) is suitable.

9.6 Conclusions

This chapter presents a design method for support scaffold systems based on advanced
analysis, which requires no recourse to a design specification for checking member and
connection strengths. The chapter also provides details for creating nominal three-
dimensional finite element nominal design models of support scaffold systems to be
used in practice in the design by advanced analysis. The design formulation is based on
the load and resistance factor design (LRFD) format, incorporating the system
resistance factor derived in Chapter 8. The step-by-step design procedure of support
scaffold systems by advanced analysis is proposed. Finally, two design case studies are
Chapter 9 - Design of Support Scaffold Systems 222
______________________________________________________________________

presented to exemplify the practical use of design by advanced analysis of support


scaffold systems.
Chapter 10 - Conclusions 223
______________________________________________________________________

CHAPTER 10 - CONCLUSIONS

10.1 Summary

The thesis provides comprehensive studies into the direct design of steel support
scaffold systems by advanced geometric and material nonlinear analysis. The particular
system studied, referred to as Cuplok scaffold system, is primarily used to support
formwork and the weight of wet poured concrete.

The research program encompasses the collection and statistical evaluation of field data
of loading eccentricity, member out-of-straightness (member crookedness), and storey
out-of-plumb, as well as eighteen full-scale tests of 3x3 bay subassemblies and a
comprehensive series of tests on Cuplok joints in various configurations to determine
statistical distributions for the joint strength, semi-rigid stiffness and looseness.
Advanced finite element models are calibrated using the full-scale subassembly tests,
thus providing statistical data for the modelling error. Parametric studies are performed
to identify the main factors influencing the system strength.

Having procured the statistical data from site survey and experimental tests for the main
random variables affecting the strength of support scaffold systems, Monte-Carlo
simulations using advanced analysis are carried out to determine the statistical
distributions of strength for a range of geometries of support scaffold systems. From
these studies, the failure mode is mainly determined by the extension of the jacks at the
top and bottom of the frame as well as the lift height. The first-order second-moment
and first-order reliability methods are used to estimate the system resistance factors for
support scaffold systems in various geometric configurations, leading to the formulation
of a design method for support scaffold systems based on advanced analysis, which
does not require tedious checking of member and connection strengths since the
strengths of these components are verified in the course of running the advanced
analysis.
Chapter 10 - Conclusions 224
______________________________________________________________________

10.2 Remarks

The literature review in Chapter 2 provides guidelines for the modelling, analysis and
design of scaffold systems based on past research. The key findings are as follows:

(a) Initial geometric imperfections that include sway of the frame and out-of-
straightness of the uprights need to be incorporated in the structural model so that
geometric second-order effects are considered in the nonlinear analysis. Three methods
of modelling imperfections are the scaling of eigenbuckling modes (EBM), the
application of notional horizontal forces (NHF), and the direct modelling of initial
geometric imperfections (IGI). The magnitudes of imperfections applied to the model
are usually taken from the available codes of practice.

(b) Many codes of practice allow the use of advanced analysis that takes into account
inelastic material properties and geometric imperfections, provided that a structure has
sufficient section capacity. However, before the design by advanced analysis can be
fully put into practice, a reliability analysis of the structural system is necessary in order
to establish the required system resistance factor in LRFD framework.

(c) Support scaffold systems should be monitored by their axial forces and
displacements of the standards especially during concrete placement, and inspected to
check that bracings are applied correctly and adequately. For access scaffold systems,
sufficient ties to a permanent structure must be provided to prevent excessive lateral
movement.

In Chapter 3, the measurements of geometric imperfections and loading eccentricity of


Cuplok support scaffold systems have been obtained from various construction sites in
the Sydney metropolitan area. The important findings are as follows:

(a) The mean normalised out-of-straightness of the standards including standards with
and without spigot joints is 0.00048 (Lh/2080) with a standard deviation of 0.00042;
however, it is computed that the mean normalised out-of-straightness of the standards
Chapter 10 - Conclusions 225
______________________________________________________________________

with spigot joints is 0.0013 (Lh/770) with a standard deviation of 0.0008 and the mean
normalised out-of-straightness of the standards without spigot joints is 0.0004 (Lh/2500)
with a standard deviation of 0.0003. The mean normalised storey out-of-plumb is
0.0016 (H/625) with a standard deviation of 0.0005 whereas the mean loading
eccentricity is 18.09 mm with a standard deviation of 10.67 mm.

(b) By comparing the mean values to national standards, it is shown that the actual out-
of-straightness of standards with spigot joints is over the limit imposed by the
Australian steel structures standard AS 4100 (Standard Australia 1998), and that the
magnitude of the loading eccentricity may be higher than the expected loading
eccentricity specified by the Australian formwork for concrete AS 3610 (Standard
Australia 1995). Consequently, the structural modelling based on existing codes might
lead to less conservative or inadequate design of support scaffold systems. Nevertheless,
the mean of the storey out-of-plumb is shown to be well within the limit imposed by the
Australian formwork for concrete AS 3610 (Standard Australia 1995).

(c) From the statistical studies, the probability distribution for the member out-of-
straightness (member crookedness) is found to be approximately lognormal. The frame
out-of-plumb is fitted to a normal distribution and the loading eccentricity is fitted to a
lognormal distribution.

In Chapter 4, the Cuplok joint stiffness of support scaffold systems is investigated for
bending in the vertical and horizontal planes. The important findings are as follows:

(a) The joint stiffness for bending in the vertical plane is found to be much higher than
the stiffness for bending in the horizontal plane, and 4-way connections provide greater
stiffness than other joint configurations. However, the surface finish (galvanised or
painted components), and the number of hammer blows used to tighten the connection
have insignificant effect on the joint stiffness.

(b) The joint stiffness data fits well to normal distributions for probabilistic modelling.
Chapter 10 - Conclusions 226
______________________________________________________________________

Chapter 5 presents a summary of the procedure and experimental results pertaining to a


series of formwork subassembly tests performed at the University of Sydney as part of a
consultancy project for Boral Formwork and Scaffolding Pty. Ltd. The main
observations are as follows:

(a) The failure modes are controlled by the jack extension length since when 600 mm
top and bottom extensions are used, the failure mode of the system is an overall sway
with final failure at the jacks. On the contrary, when 300 mm extensions are used,
failure occurs mainly in the standards and spigots with only small sway displacements.
Noticeably, the ultimate load decreases as the jack extension increases.

(b) The bracing arrangement significantly influences the ultimate load of the system.
Also, the higher lift height reduces the ultimate load. In addition, the standards tend to
fail at the top lift and around the perimeter region, especially at the corner where there is
no bracing and only two ledgers are connected at the (2-way) joints, which therefore
have relatively lower stiffness and strength. Final failure occurred in spigots and jacks
in most cases.

In Chapter 6, nonlinear finite element analysis models for support scaffold systems are
developed. The main important features are as follows:

(a) Models for various components of the systems including spigot joints, semi-rigid
upright-to-beam connections and base plate eccentricities are proposed.

(b) Calibrations of these finite element models to the full-scale subassembly tests
(CASE 2006) consisting of three-by-three bay formwork systems with the combinations
of different numbers of lifts, jack extension, and lift height are achieved by adjusting the
top and bottom boundary conditions as well as the brace connection stiffness.

(c) Most of the system strength predictions from the advanced analysis models are
within 10% of the actual failure loads. In addition, advanced analysis gives accurate
results in predicting deformation responses of support scaffold systems. The finite
Chapter 10 - Conclusions 227
______________________________________________________________________

element analysis results of the load-deflection responses fit the test results reasonably
closely with most of the values within 20% of one another.

In Chapter 7, an investigation of the factors affecting the system strength of support


scaffold systems with Cuplok joints is presented. The main findings are as follows:

(a) The loading eccentricity, member out-of-straightness, joint stiffness, and material
yield stresses are found to be the main variables affecting the strength of support
scaffold systems.

(b) The variations in cross-sectional geometry considered in the study have insignificant
effect on the system strength.

In Chapter 8, the effects of uncertainties in material properties, initial geometric


imperfections and joint stiffness on the ultimate strength of steel support scaffold
systems are studied. The chapter also presents a reliability analysis of support scaffold
systems to determine system resistance factor that can be used in the design by
advanced analysis according to the LRFD framework. Some pertinent remarks are as
follows:

(a) A set of system resistance factors, system, determined by a first-order reliability


method (FORM) at the system target reliability index of 3.5, is proposed to be used in
the LRFD design of support scaffold (Cuplok) systems by advanced analysis.

(b) The proposed system resistance factors vary from 0.56 to 0.70, depending on the lift
height and jack extensions.

In Chapter 9, a design method for support scaffold systems based on advanced analysis,
which requires no recourse to a steel design specification for checking member and
connection strengths, is presented. The ultimate limit state design procedure for support
scaffold systems starts with determining gravity loads and a preliminary system
configuration, and then the nominal three-dimensional nonlinear finite element model
including geometric imperfections for the structure is created. An advanced structural
Chapter 10 - Conclusions 228
______________________________________________________________________

analysis is performed with the combined factored loads incrementally applied to the
system until the ultimate load is reached. Finally, the adequacy of the structural system
to support the applied ultimate loads can be readily evaluated by comparing the system
ultimate strength with the applied factored loads, based on the LRFD framework.

10.3 Recommendations for Future Research

The following suggestions are made for possible avenues of future research.

(a) Since the scaffold industry is considering in reducing the number of braces used in
the systems to achieve greater competitiveness in terms of faster erection speed, an
extension of the design by advanced analysis presented in this thesis to cover selective
bracing configurations is needed to ensure that such selectively braced steel scaffold
systems are safe and reliable.

(b) An improvement in the standard-to-ledger joint stiffness is in demand to increase the


system strength of support scaffold systems. A new design of Cuplok joints could be
studied to enhance the joint stiffness, especially in the horizontal plane bending.

(c) A collection and statistical evaluation of field data for the base eccentricity and
ground settlement are needed in order to accurately incorporate these factors into the
design procedure for support scaffold systems.

(d) More probabilistic data for construction loads is needed for improved estimation of
the load statistics for support scaffold systems.

(e) Support scaffold systems with other joint types, notably wedge-type joints, should
be studied using the same modelling and probabilistic concepts as presented in this
thesis.
References 229
______________________________________________________________________

REFERENCES

AISC. (1986). "AISC LRFD specification."


AISC. (1994). "AISC LRFD specification, manual of steel construction."
AISC. (2005). "AISC Specification for structural steel buildings."
ANSYS. (1998). "ANSYS. Release 5.5. ANSYS basic analysis procedures guide."
Swanson Analysis Systems.
ASCE. (2005). "Minimum design loads for buildings and other structures. ASCE 7-05."
American Society of Civil Engineers, New York.
Ayyub, B. M., and McCuen, R. H. (2003). Probability, statistics, and reliability for
engineers and scientists, Chapman Hall/CRC, Boca Raton, Florida.
Boral Formwork & Scaffolding Pty. Ltd. (2002). "Supercuplok support system: general
technical and application manual (http://www.boral.com.au/scaffolding)."
Revesby NSW, Australia.
British Standards Institution. (1996). "BS 5975: Code of practice for falsework."
British Standards Institution. (2000). "BS 5950: Structural use of steelwork in building.
Part1: Code of practice for design - rolled and welded sections."
Bulleit, W. M. (2008). "Uncertainty in structural engineering." Practice Periodical on
Structural Design and Construction, 13(1), 24-30.
Buonopane, S. G., and Schafer, B. W. (2006). "Reliability of steel frames designed with
advanced analysis." Journal of Structural Engineering, 132(2), 267-276.
Buonopane, S. G., Schafer, B. W., and Igusa, T. (2003). "Reliability implications of
advanced analysis in design of steel frames." Advances in Structures: Steel,
Concrete, Composite and Aluminium - ASSCCA'03 Sydney, Australia.
CASE. (2006). "Tests of formwork subassemblies and components. Investigation
Report No. S1499." Centre for Advanced Structural Engineering, School of
Civil Engineering, University of Sydney.
Chan, S. L., Dymiotis, C., and Zhou, Z. H. (2002). "Second-order analysis and design of
steel scaffold using multiple eigen-imperfection modes." Proceedings of the
Third International Conference on Advances in Steel Structures, Hong Kong,
321-327.
Chan, S. L., Huang, H. Y., and Fang, L. X. (2005). "Advanced analysis of imperfect
portal frames with semirigid base connections." Journal of Engineering
Mechanics, 131(6), 633-640.
Chen, W. F., and Kim, S. E. (1997). LRFD steel design using advanced analysis, CRC
Press, New York.
Chu, A. Y. T., Chan, S. L., and Chung, K. F. (2002). "Stability of modular steel
scaffolding systems - theory and verification." Proceedings of International
Conference Advances in Building Technology, Hong Kong, 621-628.
Clarke, M. J., Bridge, R. Q., Hancock, G. J., and Trahair, N. S. (1992). "Advanced
analysis of steel building frames." Journal of Constructional Steel Research,
23(1-3), 1-29.
D'Agostino, R. B., and Stephens, M. A. (1986). Goodness-of-fit Techniques, Marcel
Dekker, New York.
Ellingwood, B., and Galambos, T. V. (1982). "Probability-based criteria for structural
design." Structural Safety, 1, 15-26.
References 230
______________________________________________________________________

Ellingwood, B., Galambos, T. V., MacGregor, J. G., and Cornell, C. A. (1980).


"Development of a probability-based load criterion for American National
Standard A58, NBS special publication 577." National Bureau of Standards,
Washington, D.C.
Enright, J., Harriss, R., and Hancock, G. J. (2000). "Structural stability of braced
scaffolding and formwork with spigot joints." Proceedings of the Fifteenth
International Specialty Conference on Cold-Formed Steel Structures, St. Louis,
Missouri, USA, 357-376.
Galambos, T. V. (1990). "Systems reliability and structural design." Structural Safety,
7(2-4), 101-108.
Galambos, T. V., and Ravindra, M. K. (1978). "Properties of steel for use in LRFD."
Journal of the Structural Division, American Society of Civil Engineers, 104(9),
1459-1468.
Godley, M. H. R., and Beale, R. G. (2001). "Analysis of large proprietary access
scaffold structures." Proceedings of the Institution of Civil Engineers, Structures
& Buildings, 146(1), 31-39.
Godley, M. H. R., and Beale, R. G. (1997). "Sway stiffness of scaffold structures."
Structural Engineer, 75(1), 4-12.
Gylltoft, K., and Mroz, K. (1995). "Load carrying capacity of scaffolds." Structural
Engineering International, 1, 37-42.
Hadipriono, F. C. (1986). "Approximate reasoning for falsework safety assessment."
Structural Safety, 4(2), 131-140.
Hadipriono, F. C., and Wang, H. K. (1987). "Causes of falsework collapses during
construction." Structural Safety, 4(3), 179-195.
Harung, H. S., Lightfoot, E., and Duggan, D. M. (1975). "The strength of scaffold
towers under vertical loading." Structural Engineer, 53(1), 23-30.
Hasofer, A. M., and Lind, N. C. (1974). "Exact and invariant second-moment code
format." ASCE Journal of Engineering Mechanics Division, 100, 111-121.
Huang, Y. L., Chen, H. J., Rosowsky, D. V., and Kao, Y. G. (2000a). "Load-carrying
capacities and failure modes of scaffold-shoring systems, Part I: Modeling and
experiments." Structural Engineering and Mechanics, 10(1), 53-66.
Huang, Y. L., Chen, W. F., Chen, H. J., Yen, T., Kao, Y. G., and Lin, C. Q. (2000b).
"Monitoring method for scaffold-frame shoring systems for elevated concrete
formwork." Computers and Structures, 78(5), 681-690.
Huang, Y. L., Kao, Y. G., and Rosowsky, D. V. (2000c). "Load-carrying capacities and
failure modes of scaffold-shoring systems, Part II: An analytical model and its
closed-form solution." Structural Engineering and Mechanics, 10(1), 67-79.
Karshenas, S., and Ayoub, H. (1994a). "Analysis of concrete construction live loads on
newly poured slabs." Journal of Structural Engineering, 120(5), 1525-1542.
Karshenas, S., and Ayoub, H. (1994b). "Construction live loads on slab formworks
before concrete placement." Structural Safety, 14(3), 155-172.
Kim, S. E., and Chen, W. F. (1999). "Design guide for steel frames using advanced
analysis program." Engineering Structures, 21(4), 352-364.
Li, J. J., and Li, G. Q. (2004). "Reliability-based integrated design of steel portal frames
with tapered members." Structural Safety, 26(2), 221-239.
Lightfoot, E., and Bhula, D. (1977). "A test rig for scaffold couplers." Materials and
Structures, 10(3), 168-173.
References 231
______________________________________________________________________

Lightfoot, E., and Oliveto, G. (1977). "The collapse strength of tubular steel scaffold."
Proceedings of the Institution of Civil Engineers (London). Part 1 - Design &
Construction, UK, 311-329.
LUSAS. (1998). "LUSAS. User manual version 12.2." FEA Ltd., UK.
Melchers, R. E. (1999). Structural reliability analysis and prediction, John Wiley &
Sons Ltd., West Sussex, England.
Milojkovic, B., Beale, R. G., and Godley, M. H. R. (1996). "Modelling scaffold
connections." Proceedings of the Fourth ACME UK Annual Conference,
Glasgow, UK, 85-88.
Milojkovic, B., Beale, R. G., and Godley, M. H. R. (2002). "Determination of the
factors of safety of standard scaffold structures." Proceedings of International
Conference Advances in Steel Structures, UK, 303-310.
NAF-NIDA. (2001). "NAF-NIDA. Software for nonlinear integrated design and
analysis version 3. User's manual." Department of Civil and Structural
Engineering, Hong Kong Polytechnic University.
Nowak, A. S., and Collins, K. R. (2000). Reliability of structures, McGraw-Hill, New
York.
Peng, J. L. (2002). "Stability analyses and design recommendations for practical shoring
systems during construction." Journal of Construction Engineering and
Management ASCE, 128(6), 536-544.
Peng, J. L. (2004). "Structural modeling and design considerations for double-layer
shoring systems." Journal of Construction Engineering and Management ASCE,
130(3), 368-377.
Peng, J. L., Chan, S. L., and Wu, C. L. (2007). "Effects of geometrical shape and
incremental loads on scaffold systems." Journal of Constructional Steel
Research, 63(4), 448-459.
Peng, J. L., Pan, A. D., Rosowsky, D. V., Chen, W. F., Yen, T., and Chan, S. L. (1996a).
"High clearance scaffold systems during construction --I. Structural modelling
and modes of failure." Engineering Structures, 18(3), 247-257.
Peng, J. L., Pan, A. D., Rosowsky, D. V., Chen, W. F., Yen, T., and Chan, S. L. (1996b).
"High clearance scaffold systems during construction --II. Structural analysis
and development of design guidelines." Engineering Structures, 18(3), 258-267.
Peng, J. L., Pan, A. D., and Chan, S. L. (1998). "Simplified models for analysis and
design of modular falsework." Journal of Constructional Steel Research, 48(2-3),
189-209.
Peng, J. L., Pan, A. D., Chen, W. F., Yen, T., and Chan, S. L. (1997). "Structural
modeling and analysis of modular falsework systems." Journal of Structural
Engineering, 123(9), 1245-1251.
Peng, J. L., Rosowsky, D. V., Pan, A. D., Chen, W. F., Chan, S. L., and Yen, T. (1996c).
"Analysis of concrete placement load effects using influence surfaces." ACI
Structural Journal, 93(2), 180-186.
Peng, J. L., Wu, C. L., and Chan, S. L. (2003). "Sequential pattern load modeling and
warning-system plan in modular falsework." Structural Engineering and
Mechanics, 16(4), 441-468.
Prabhakaran, U., Godley, M. H. R., and Beale, R. G. (2006). "Three-dimensional
second order analysis of scaffolds with semi-rigid connections." Welding in the
World, 50 (SPEC ISS), 187-194.
Rackwitz, R., and Fiessler, B. (1978). "Structural reliability under combined random
load sequences." Computers & Structures, 9(5), 489-494.
References 232
______________________________________________________________________

Ramberg, W., and Osgood, W. R. (1941). "Determination of stressstrain curves by


three parameters. Technical Note No. 503." National Advisory Committee on
Aeronautics (NACA).
Rasmussen, K. J. R. (2003). "Full-range stress-strain curves for stainless steel alloys."
Journal of Constructional Steel Research, 59(1), 47-61.
Rosowsky, D. V. (1996). "Load combinations and load factors for construction."
Journal of Performance of Constructed Facilities, 10(4), 175-181.
Rosowsky, D. V., and Stewart, M. G. (2001). "Probabilistic construction load model for
multistory reinforced-concrete buildings." Journal of Performance of
Constructed Facilities, 15(4), 145-152.
Sivakumar Babu, G. L., and Basha, B. M. (2008). "Optimum design of cantilever
retaining walls using target reliability approach." International Journal of
Geomechanics, 8(4), 240-252.
Standard Australia. (1995). "AS 3610: Formwork for concrete."
Standard Australia. (1996). "AS 3679.1: Structural steel - hot-rolled bars and sections."
Standard Australia. (1998). "AS 4100: Steel structures."
Standard Australia. (2009). "AS 1163: Cold-formed structural steel hollow sections."
Strand7. (2009). "Strand7 release 2.4.1 manual." Strand7 Pty. Ltd.
Vaux, S., Wong, C., and Hancock, G. (2002). "Sway stability of steel scaffolding and
formwork systems." Proceedings of the Third International Conference on
Advances in Steel Structures, Hong Kong, 311-319.
Weesner, L. B., and Jones, H. L. (2001). "Experimental and analytical capacity of frame
scaffolding." Engineering Structures, 23(6), 592-599.
Yu, W. K. (2004). "An investigation into structural behaviour of modular steel
scaffolds." Steel & Composite Structures, 4(3), 211-226.
Yu, W. K., and Chung, K. F. (2004). "Prediction on load carrying capacities of multi-
storey door-type modular steel scaffolds." Steel & Composite Structures, 4(6),
471-487.
Yu, W. K., Chung, K. F., and Chan, S. L. (2004). "Structural instability of multi-storey
door-type modular steel scaffolds." Engineering Structures, 26(7), 867-881.
Zhang, L., Tang, W. H., and Ng, C. W. W. (2001). "Reliability of axially loaded driven
pile groups." Journal of Geotechnical and Geo-Environmental Engineering,
127(12), 1051-1060.
Appendix A.1 - Out-of-Straightness Data 233
______________________________________________________________________

APPENDIX A.1 - OUT-OF-STRAIGHTNESS DATA

Table A.1: Data for out-of-straightness of the standards


(TS = Top lift standard, BS = Bottom lift standard, A = Train station platform, B =
Parking structure, C = Residential building and D = Office building)

Out-of- Normalised out-of-


Lift
Grid straightness straightness
Member height Note
Site location X- Y-
(m) X-Axis Y-Axis
Axis Axis
(mm/mm) (mm/mm)
(mm) (mm)
A 19x15 TS 1 3.36 -1.20 0.00336 0.00120 spigot
A 19x15 BS 1 0.90 0.35 0.00090 0.00035
A 20x14 BS 1 -0.86 0.86 0.00086 0.00086
A 20x14 TS 1 0.85 -0.51 0.00085 0.00051
A 23x15 BS 1 0.14 -0.21 0.00014 0.00021
A 23x15 TS 1 1.61 0.11 0.00161 0.00011 spigot
A 24x14 TS 1 1.81 2.60 0.00181 0.00260 spigot
A 24x14 BS 1 0.16 -0.23 0.00016 0.00023
A 28x17 BS 1 0.38 0.25 0.00038 0.00025
A 28x17 TS 1 0.35 -0.28 0.00035 0.00028
A 27x16 BS 1 -0.18 -0.95 0.00018 0.00095
A 27x16 TS 1 0.84 0.43 0.00084 0.00043
A 29x17 BS 1 -0.26 0.22 0.00026 0.00022
A 29x17 TS 1 -2.13 -1.21 0.00213 0.00121 spigot
A 30x17 BS 1 0.39 0.42 0.00039 0.00042
A 30x17 TS 1 -0.40 -1.60 0.00040 0.00160
A 19x10 BS 1 -0.80 -0.12 0.00080 0.00012
A 19x10 TS 1 0.26 0.72 0.00026 0.00072
A 19x11 BS 1 0.55 -0.90 0.00055 0.00090
A 19x11 TS 1 1.20 -0.22 0.00120 0.00022 spigot
A 20x10 BS 1 -0.14 0.12 0.00014 0.00012
A 20x10 TS 1 0.56 -0.20 0.00056 0.00020
A 30x11 BS 1 0.12 0.21 0.00012 0.00021
A 30x11 TS 1 0.42 0.69 0.00042 0.00069
A 29x11 BS 1 0.13 -0.14 0.00013 0.00014
A 29x11 TS 1 0.72 -0.35 0.00072 0.00035
A 1x1 BS 1 0.31 -0.25 0.00031 0.00025
A 1x2 BS 1 -0.41 0.12 0.00041 0.00012
A 2x1 BS 1 0.45 0.51 0.00045 0.00051
A 2x2 BS 1 0.22 -0.15 0.00022 0.00015
A 1x1 TS 1 -0.30 0.12 0.00030 0.00012
A 1x2 TS 1 -0.43 0.41 0.00043 0.00041
Appendix A.1 - Out-of-Straightness Data 234
______________________________________________________________________

Out-of- Normalised out-of-


Lift
Grid straightness straightness
Member height Note
Site location X- Y-
(m) X-Axis Y-Axis
Axis Axis
(mm/mm) (mm/mm)
(mm) (mm)
A 2x1 TS 1 -0.11 0.52 0.00011 0.00052
A 2x2 TS 1 0.24 -0.48 0.00024 0.00048
A 3x1 BS 1 0.23 0.20 0.00023 0.00020
A 4x1 BS 1 -0.45 0.10 0.00045 0.00010
A 3x2 BS 1 0.10 -0.34 0.00010 0.00034
A 4x2 BS 1 0.25 0.35 0.00025 0.00035
A 3x1 TS 1 0.32 0.42 0.00032 0.00042
A 4x1 TS 1 0.25 -0.41 0.00025 0.00041
A 3x2 TS 1 0.53 -0.10 0.00053 0.00010
A 4x2 TS 1 0.31 0.32 0.00031 0.00032
A 5x1 BS 1 0.46 0.10 0.00046 0.00010
A 5x2 BS 1 0.56 0.07 0.00056 0.00007
A 6x1 BS 1 -0.70 -0.05 0.00070 0.00005
A 6x2 BS 1 -0.20 0.12 0.00020 0.00012
A 5x1 TS 1 0.10 -0.23 0.00010 0.00023
A 5x2 TS 1 0.08 -0.25 0.00008 0.00025
A 6x1 TS 1 0.12 0.36 0.00012 0.00036
A 6x2 TS 1 0.09 0.07 0.00009 0.00007
B 1x1 BS 1 0.26 -0.20 0.00026 0.00020
B 1x2 BS 1 0.20 -0.19 0.00020 0.00019
B 1x3 BS 1 0.25 -0.18 0.00025 0.00018
B 1x4 BS 1 0.19 0.20 0.00019 0.00020
B 1x5 BS 1 0.12 0.08 0.00012 0.00008
B 1x6 BS 1 -0.38 0.20 0.00038 0.00020
B 1x7 BS 1 -0.15 0.21 0.00015 0.00021
B 1x8 BS 1 0.20 0.10 0.00020 0.00010
B 1x9 BS 1 0.10 0.18 0.00010 0.00018
B 1x10 BS 1 0.32 -0.12 0.00032 0.00012
B 1x11 BS 1 0.60 0.14 0.00060 0.00014
B 1x12 BS 1 0.25 0.23 0.00025 0.00023
B 1x13 BS 1 0.32 -0.26 0.00032 0.00026
B 1x14 BS 1 0.50 -0.40 0.00050 0.00040
B 1x15 BS 1 -0.42 0.35 0.00042 0.00035
B 1x16 BS 1 0.32 -0.25 0.00032 0.00025
B 1x17 BS 1 0.23 -0.34 0.00023 0.00034
B 1x18 BS 1 -0.45 0.23 0.00045 0.00023
B 1x19 BS 1 0.30 -0.38 0.00030 0.00038
B 1x20 BS 1 0.39 -0.25 0.00039 0.00025
B 1x1 TS 1 0.15 -0.21 0.00015 0.00021
B 1x2 TS 1 0.18 0.23 0.00018 0.00023
B 1x3 TS 1 0.25 -0.32 0.00025 0.00032
B 1x4 TS 1 -0.32 0.24 0.00032 0.00024
Appendix A.1 - Out-of-Straightness Data 235
______________________________________________________________________

Out-of- Normalised out-of-


Lift
straightness straightness
Grid Member height Note
Site X- Y-
location (m) X-Axis Y-Axis
Axis Axis
(mm/mm) (mm)
(mm) (mm)
B 1x5 TS 1 -0.14 0.19 0.00014 0.00019
B 1x6 TS 1 0.25 0.18 0.00025 0.00018
B 1x7 TS 1 0.50 -0.48 0.00050 0.00048
B 1x8 TS 1 0.15 -0.23 0.00015 0.00023
B 1x9 TS 1 -0.24 0.34 0.00024 0.00034
B 1x10 TS 1 0.32 0.24 0.00032 0.00024
B 1x11 TS 1 0.51 -0.39 0.00051 0.00039
B 1x12 TS 1 0.18 0.28 0.00018 0.00028
B 1x13 TS 1 0.26 -0.25 0.00026 0.00025
B 1x14 TS 1 0.38 -0.41 0.00038 0.00041
B 1x15 TS 1 0.25 0.15 0.00025 0.00015
B 1x16 TS 1 1.00 1.30 0.00100 0.00130 spigot
B 1x17 TS 1 0.26 0.18 0.00026 0.00018
B 1x18 TS 1 -0.12 0.19 0.00012 0.00019
B 1x19 TS 1 0.26 -0.15 0.00026 0.00015
B 1x20 TS 1 0.18 0.24 0.00018 0.00024
C 2x1 BS 1.5 -1.10 0.60 0.00073 0.00040
C 2x2 BS 1.5 -1.30 0.50 0.00087 0.00033
C 2x3 BS 1.5 1.60 -0.50 0.00107 0.00033
C 2x4 BS 1.5 -0.90 1.10 0.00060 0.00073
C 2x5 BS 1.5 1.10 -1.00 0.00073 0.00067
C 2x6 BS 1.5 -0.95 0.10 0.00063 0.00007
C 2x7 BS 1.5 0.35 -0.20 0.00023 0.00013
C 2x8 BS 1.5 1.12 -0.52 0.00075 0.00035
C 2x9 BS 1.5 0.20 -0.24 0.00013 0.00016
C 2x10 BS 1.5 0.98 -0.75 0.00065 0.00050
C 2x11 BS 1.5 -1.13 -1.30 0.00075 0.00087
C 2x12 BS 1.5 -1.23 0.51 0.00082 0.00034
C 2x13 BS 1.5 0.71 -0.43 0.00047 0.00029
C 2x14 BS 1.5 -1.12 0.42 0.00075 0.00028
C 2x1 TS 1.5 3.12 3.50 0.00208 0.00233 spigot
C 2x2 TS 1.5 -1.21 1.35 0.00081 0.00090
C 2x3 TS 1.5 2.10 -2.15 0.00140 0.00143 spigot
C 2x4 TS 1.5 -1.54 -1.12 0.00103 0.00075
C 2x5 TS 1.5 1.24 1.10 0.00083 0.00073
C 2x6 TS 1.5 1.10 -2.50 0.00073 0.00167 spigot
C 2x7 TS 1.5 2.40 -1.10 0.00160 0.00073 spigot
C 2x8 TS 1.5 1.55 -1.45 0.00103 0.00097
C 2x9 TS 1.5 0.54 -1.49 0.00036 0.00099
C 2x10 TS 1.5 0.56 0.12 0.00037 0.00008
C 2x11 TS 1.5 0.95 -0.85 0.00063 0.00057
C 2x12 TS 1.5 0.90 -0.25 0.00060 0.00017
Appendix A.1 - Out-of-Straightness Data 236
______________________________________________________________________

Out-of- Normalised out-of-


Lift
Grid straightness straightness
Member height Note
Site location X- Y-
(m) X-Axis Y-Axis
Axis Axis
(mm/mm) (mm/mm)
(mm) (mm)
C 2x13 TS 1.5 1.40 0.14 0.00093 0.00009
C 2x14 TS 1.5 0.92 -1.30 0.00061 0.00087
C 3x1 TS 1.5 0.43 -0.32 0.00029 0.00021
C 3x2 TS 1.5 1.25 0.50 0.00083 0.00033 spigot
C 3x3 TS 1.5 -0.27 0.50 0.00018 0.00033
C 3x4 TS 1.5 1.10 1.12 0.00073 0.00075
C 3x5 TS 1.5 -0.32 0.95 0.00021 0.00063
C 3x6 TS 1.5 0.15 0.58 0.00010 0.00039
C 3x7 TS 1.5 1.12 -0.15 0.00075 0.00010 spigot
C 3x8 TS 1.5 1.40 0.90 0.00093 0.00060 spigot
C 3x9 TS 1.5 -0.70 0.30 0.00047 0.00020
D 1x1 BS 1.5 0.10 1.42 0.00007 0.00095
D 1x2 BS 1.5 -0.30 0.15 0.00020 0.00010
D 2x1 BS 1.5 0.90 -1.12 0.00060 0.00075
D 2x2 BS 1.5 1.00 1.10 0.00067 0.00073
D 1x1 TS 1 -0.35 0.40 0.00035 0.00040
D 1x2 TS 1 0.45 0.63 0.00045 0.00063
D 2x1 TS 1 0.51 0.62 0.00051 0.00062
D 2x2 TS 1 -0.23 -0.10 0.00023 0.00010
D 3x1 BS 1.5 2.14 -1.12 0.00143 0.00075 spigot
D 4x1 BS 1.5 1.13 -0.95 0.00075 0.00063
D 3x2 BS 1.5 -1.23 0.56 0.00082 0.00037
D 4x2 BS 1.5 1.93 -1.02 0.00129 0.00068
D 3x1 TS 1 -0.94 0.62 0.00094 0.00062
D 4x1 TS 1 0.51 -0.31 0.00051 0.00031
D 3x2 TS 1 0.43 0.78 0.00043 0.00078
D 4x2 TS 1 0.96 -0.42 0.00096 0.00042
D 5x1 BS 1.5 -1.41 0.32 0.00094 0.00021
D 5x2 BS 1.5 0.98 -1.93 0.00065 0.00129
D 6x1 BS 1.5 0.86 1.12 0.00057 0.00075
D 6x2 BS 1.5 1.10 -1.45 0.00073 0.00097
D 5x1 TS 1 0.44 0.61 0.00044 0.00061
D 5x2 TS 1 0.36 -0.79 0.00036 0.00079
D 6x1 TS 1 0.78 0.10 0.00078 0.00010
D 6x2 TS 1 0.21 0.14 0.00021 0.00014
Appendix A.2 - Out-of-Plumb Data 237
______________________________________________________________________

APPENDIX A.2 - OUT-OF-PLUMB DATA

Table A.2: Data for storey out-of-plumb


(C = Residential building and D = Office building)

Bottom Top X- Y- Length Normalised


horizontal horizontal axis axis from H out-of-
Site Grid
angle angle sway sway origin (m) plumb
(degree) (degree) (mm) (mm) (m) (mm/mm)
1x1
C 122.167 122.233 Nil -3.72 3.20 3.00 0.00124
(North)
2x1
C 107.358 107.342 Nil 0.91 3.12 3.00 0.00030
(North)
3x1
C 280.250 279.883 Nil 8.01 1.25 3.00 0.00267
(North)
4x1
C 307.083 307.250 Nil -5.98 2.05 3.00 0.00199
(North)
5x1
C 314.442 314.600 Nil -5.93 2.15 3.00 0.00198
(North)
1x2
C 350.803 350.391 Nil 7.91 1.10 3.00 0.00264
(North)
2x2
C 337.250 337.520 Nil -6.36 1.35 3.00 0.00212
(North)
3x2
C 412.321 412.100 Nil 5.40 1.40 3.00 0.00180
(North)
4x2
C 71.317 71.512 Nil -5.79 1.70 3.00 0.00193
(North)
5x2
C 134.583 134.700 Nil -5.11 2.50 3.00 0.00170
(North)
1x3
C 111.233 111.317 Nil -4.40 3.00 3.00 0.00147
(North)
2x3
C 140.633 140.533 Nil 4.36 2.50 3.00 0.00145
(North)
3x3
C 89.700 89.817 Nil -5.21 2.55 3.00 0.00174
(North)
4x3
C 120.750 120.867 Nil -3.98 1.95 3.00 0.00133
(North)
5x3
C 114.250 114.367 Nil -5.00 2.45 3.00 0.00167
(North)
1x4
C 230.725 230.825 Nil -4.63 2.65 3.00 0.00154
(North)
2x4
C 331.992 331.858 Nil 4.98 2.13 3.00 0.00166
(North)
Appendix A.2 - Out-of-Plumb Data 238
______________________________________________________________________

Bottom Top X- Y- Length Normalised


horizontal horizontal axis axis from H out-of-
Site Grid
angle angle sway sway origin (m) plumb
(degree) (degree) (mm) (mm) (m) (mm/mm)
3x4
C 320.208 320.308 Nil -3.75 2.15 3.00 0.00125
(North)
4x4
C 340.442 340.525 Nil -3.16 2.18 3.00 0.00105
(North)
5x4
C 315.775 315.692 Nil 3.26 2.25 3.00 0.00109
(North)
1x1
C 11.542 11.575 -1.41 Nil 2.45 3.00 0.00047
(East)
2x1
C 21.517 21.417 3.89 Nil 2.23 3.00 0.00130
(East)
3x1
C 50.733 50.833 -4.45 Nil 2.55 3.00 0.00148
(East)
4x1
C 65.416 65.483 -3.25 Nil 2.78 3.00 0.00108
(East)
5x1
C 32.517 32.583 -4.09 Nil 3.55 3.00 0.00136
(East)
1x2
C 154.203 154.192 0.47 Nil 2.43 3.00 0.00016
(East)
2x2
C 170.358 170.408 -2.33 Nil 2.67 3.00 0.00078
(East)
3x2
C 175.517 175.633 -5.85 Nil 2.89 3.00 0.00195
(East)
4x2
C 146.350 146.200 5.68 Nil 2.17 3.00 0.00189
(East)
5x2
C 151.517 151.417 4.10 Nil 2.35 3.00 0.00137
(East)
1x3
C 95.717 95.850 -5.22 Nil 2.25 3.00 0.00174
(East)
2x3
C 83.683 83.817 -5.03 Nil 2.15 3.00 0.00168
(East)
3x3
C 87.850 87.983 -5.08 Nil 2.19 3.00 0.00169
(East)
4x3
C 101.467 101.433 1.16 Nil 1.95 3.00 0.00039
(East)
5x3
C 79.192 79.203 -0.45 Nil 2.35 3.00 0.00015
(East)
1x4
C 65.208 65.308 -4.15 Nil 2.38 3.00 0.00138
(East)
2x4
C 62.458 62.392 2.48 Nil 2.15 3.00 0.00083
(East)
3x4
C 87.258 87.287 -1.40 Nil 2.76 3.00 0.00047
(East)
4x4
C 97.367 97.517 -6.94 Nil 2.65 3.00 0.00231
(East)
Appendix A.2 - Out-of-Plumb Data 239
______________________________________________________________________

Bottom Top X- Y- Length Normalised


horizontal horizontal axis axis from H out-of-
Site Grid
angle angle sway sway origin (m) plumb
(degree) (degree) (mm) (mm) (m) (mm/mm)
5x4
C 109.850 109.983 -5.69 Nil 2.45 3.00 0.00190
(East)
1x1
D 102.200 102.417 Nil -9.47 2.50 5.00 0.00189
(North)
2x1
D 115.767 115.583 Nil 8.74 2.72 5.00 0.00175
(North)
3x1
D 108.700 108.850 Nil -7.85 3.00 5.00 0.00157
(North)
4x1
D 106.192 106.350 Nil -9.65 3.50 5.00 0.00193
(North)
5x1
D 110.321 110.208 Nil 6.61 3.35 5.00 0.00132
(North)
1x2
D 24.083 23.933 Nil 9.32 3.56 5.00 0.00186
(North)
2x2
D 34.367 34.520 Nil -10.3 3.87 5.00 0.00207
(North)
3x2
D 38.200 38.350 Nil -9.56 3.65 5.00 0.00191
(North)
4x2
D 21.600 21.733 Nil -7.52 3.24 5.00 0.00150
(North)
5x2
D 36.650 36.517 Nil 8.70 3.75 5.00 0.00174
(North)
1x3
D 230.183 230.325 Nil -8.55 3.45 5.00 0.00171
(North)
2x3
D 245.633 245.533 Nil 5.45 3.12 5.00 0.00109
(North)
3x3
D 254.716 254.842 Nil -8.58 3.90 5.00 0.00172
(North)
4x3
D 223.350 223.200 Nil 9.53 3.64 5.00 0.00191
(North)
5x3
D 210.433 210.558 Nil -6.81 3.12 5.00 0.00136
(North)
1x4
D 124.808 124.692 Nil 6.17 3.05 5.00 0.00123
(North)
2x4
D 146.100 146.242 Nil -9.42 3.80 5.00 0.00188
(North)
3x4
D 152.542 152.442 Nil 6.37 3.65 5.00 0.00127
(North)
4x4
D 115.708 115.842 Nil -8.07 3.45 5.00 0.00161
(North)
5x4
D 170.458 170.517 Nil -3.45 3.35 5.00 0.00069
(North)
1x1
D 310.683 310.533 9.61 Nil 3.67 5.00 0.00192
(East)
Appendix A.2 - Out-of-Plumb Data 240
______________________________________________________________________

Bottom Top X- Y- Length Normalised


horizontal horizontal axis axis from H out-of-
Site Grid
angle angle sway sway origin (m) plumb
(degree) (degree) (mm) (mm) (m) (mm/mm)
2x1
D 305.200 305.317 -7.94 Nil 3.89 5.00 0.00159
(East)
3x1
D 340.100 340.258 -11.0 Nil 4.00 5.00 0.00221
(East)
4x1
D 345.767 345.933 -10.9 Nil 3.75 5.00 0.00217
(East)
5x1
D 358.675 358.542 7.43 Nil 3.20 5.00 0.00149
(East)
1x2
D 220.992 220.875 7.56 Nil 3.70 5.00 0.00151
(East)
2x2
D 234.242 234.375 -8.05 Nil 3.47 5.00 0.00161
(East)
3x2
D 245.208 245.375 -10.4 Nil 3.56 5.00 0.00208
(East)
4x2
D 257.558 257.708 -10.1 Nil 3.87 5.00 0.00203
(East)
5x2
D 238.692 238.583 6.09 Nil 3.20 5.00 0.00122
(East)
1x3
D 32.300 32.417 -7.84 Nil 3.84 5.00 0.00157
(East)
2x3
D 45.225 45.067 10.37 Nil 3.76 5.00 0.00207
(East)
3x3
D 51.467 51.625 -10.1 Nil 3.65 5.00 0.00201
(East)
4x3
D 60.383 60.275 6.50 Nil 3.45 5.00 0.00130
(East)
5x3
D 23.450 23.283 11.02 Nil 3.78 5.00 0.00220
(East)
1x4
D 103.850 103.992 -9.79 Nil 3.95 5.00 0.00196
(East)
2x4
D 113.433 113.583 -9.01 Nil 3.44 5.00 0.00180
(East)
3x4
D 146.033 146.192 -10.4 Nil 3.74 5.00 0.00208
(East)
4x4
D 134.117 134.200 -5.43 Nil 3.75 5.00 0.00109
(East)
5x4
D 124.558 124.408 10.26 Nil 3.92 5.00 0.00205
(East)
Appendix A.3 - Loading Eccentricity Data 241
______________________________________________________________________

APPENDIX A.3 - LOADING ECCENTRICITY DATA

Table A.3: Data for loading eccentricity


(A = Train station platform, B = Parking structure, C = Residential building and D =
Office building)

Site Grid location Loading eccentricity (mm)


A 1x2 40.00
A 2x1 15.00
A 2x2 25.00
A 3x1 30.00
A 4x1 45.00
A 3x2 12.00
A 4x2 50.00
A 5x1 10.00
A 5x2 7.00
A 6x1 8.50
A 6x2 5.00
A 19x15 14.30
A 19x16 14.10
B 1x1 8.50
B 1x2 15.50
B 2x1 19.50
B 2x2 9.50
B 3x1 7.50
B 4x1 12.50
B 3x2 14.30
B 4x2 16.10
B 5x1 21.10
B 5x2 15.10
B 6x1 14.10
B 6x2 10.10
C 1x1 15.60
C 2x1 14.60
C 3x1 29.10
C 4x1 24.30
C 5x1 37.90
C 1x2 8.20
C 2x2 8.70
C 3x2 30.20
C 4x2 6.70
C 5x2 5.20
C 1x3 20.30
C 2x3 13.00
Appendix A.3 - Loading Eccentricity Data 242
______________________________________________________________________

Site Grid location Loading eccentricity (mm)


C 3x3 25.60
C 4x3 21.70
C 5x3 24.30
C 1x4 13.40
C 2x4 16.90
C 3x4 23.40
C 4x4 12.00
C 5x4 11.60
C 6x4 37.50
C 7x4 14.90
C 8x4 22.30
C 9x4 9.60
D 1x1 15.00
D 2x1 21.00
D 3x1 8.70
D 4x1 6.90
D 5x1 16.40
D 1x2 18.30
D 2x2 22.70
D 3x2 5.70
D 4x2 19.30
D 5x2 16.70
D 1x3 15.20
D 2x3 15.00
D 3x3 20.90
D 4x3 47.60
D 5x3 11.80
D 1x4 29.90
D 2x4 15.00
D 3x4 18.30
D 4x4 13.50
D 5x4 4.60
D 6x4 10.40
D 7x4 11.60
D 8x4 24.20
D 9x4 11.90
D 10x4 51.50
Appendix B.1 - Joint Stiffness Comparisons 243
______________________________________________________________________

APPENDIX B.1 JOINT STIFFNESS


COMPARISONS

The comparisons are shown for initial Cuplok joint stiffness in different joint
configurations based on the number of hammer blows (Figures B.1.1-B.1.14) and the
type of finish (Figures B.1.15-B.1.28). The trend lines are added, passing through the
mean values for better comparisons.

170
Initial joint stiffness (kNm/rad)

150

130

110

90

70

50
1 2 3 4 5 6 7 8
Hammer blows

Figure B.1.1: Initial Cuplok joint stiffness for KzA1 based on the number of hammer
blows applied to tighten the cup
Appendix B.1 - Joint Stiffness Comparisons 244
______________________________________________________________________

170

Initial joint stiffness (kNm/rad) 150

130

110

90

70

50
1 2 3 4 5 6 7 8
Hammer blows

Figure B.1.2: Initial Cuplok joint stiffness for KzB1 based on the number of hammer
blows applied to tighten the cup

170
Initial joint stiffness (kNm/rad)

150

130

110

90

70

50
1 2 3 4 5 6 7 8
Hammer blows

Figure B.1.3: Initial Cuplok joint stiffness for KzC1 based on the number of hammer
blows applied to tighten the cup
Appendix B.1 - Joint Stiffness Comparisons 245
______________________________________________________________________

170

Initial joint stiffness (kNm/rad) 150

130

110

90

70

50
1 2 3 4 5 6 7 8
Hammer blows

Figure B.1.4: Initial Cuplok joint stiffness for KzD1 based on the number of hammer
blows applied to tighten the cup

170
Initial joint stiffness (kNm/rad)

150

130

110

90

70

50
1 2 3 4 5 6 7 8
Hammer blows

Figure B.1.5: Initial Cuplok joint stiffness for KzA2 based on the number of hammer
blows applied to tighten the cup
Appendix B.1 - Joint Stiffness Comparisons 246
______________________________________________________________________

170

Initial joint stiffness (kNm/rad) 150

130

110

90

70

50
1 2 3 4 5 6 7 8
Hammer blows

Figure B.1.6: Initial Cuplok joint stiffness for KzB2 based on the number of hammer
blows applied to tighten the cup

170
Initial joint stiffness (kNm/rad)

150

130

110

90

70

50
1 2 3 4 5 6 7 8
Hammer blows

Figure B.1.7: Initial Cuplok joint stiffness for KzC2 based on the number of hammer
blows applied to tighten the cup
Appendix B.1 - Joint Stiffness Comparisons 247
______________________________________________________________________

170

Initial joint stiffness (kNm/rad) 150

130

110

90

70

50
1 2 3 4 5 6 7 8
Hammer blows

Figure B.1.8: Initial Cuplok joint stiffness for KzD2 based on the number of hammer
blows applied to tighten the cup

30
Initial joint stiffness (kNm/rad)

25

20

15

10

0
1 2 3 4 5 6 7 8
Hammer blows

Figure B.1.9: Initial Cuplok joint stiffness for KyA1 based on the number of hammer
blows applied to tighten the cup
Appendix B.1 - Joint Stiffness Comparisons 248
______________________________________________________________________

30

Initial joint stiffness (kNm/rad) 25

20

15

10

0
1 2 3 4 5 6 7 8
Hammer blows

Figure B.1.10: Initial Cuplok joint stiffness for KyB1 based on the number of hammer
blows applied to tighten the cup

30
Initial joint stiffness (kNm/rad)

25

20

15

10

0
1 2 3 4 5 6 7 8
Hammer blows

Figure B.1.11: Initial Cuplok joint stiffness for KyC1 based on the number of hammer
blows applied to tighten the cup
Appendix B.1 - Joint Stiffness Comparisons 249
______________________________________________________________________

30

Initial joint stiffness (kNm/rad) 25

20

15

10

0
1 2 3 4 5 6 7 8
Hammer blows

Figure B.1.12: Initial Cuplok joint stiffness for KyD1 based on the number of hammer
blows applied to tighten the cup

30
Initial joint stiffness (kNm/rad)

25

20

15

10

0
1 2 3 4 5 6 7 8
Hammer blows

Figure B.1.13: Initial Cuplok joint stiffness for KyC2 based on the number of hammer
blows applied to tighten the cup
Appendix B.1 - Joint Stiffness Comparisons 250
______________________________________________________________________

30

Initial joint stiffness (kNm/rad) 25

20

15

10

0
1 2 3 4 5 6 7 8
Hammer blows

Figure B.1.14: Initial Cuplok joint stiffness for KyD2 based on the number of hammer
blows applied to tighten the cup

170
Initial joint stiffness (kNm/rad)

150

130

110

90

70

50
Galvanised Painted
Materials

Figure B.1.15: Initial Cuplok joint stiffness for KzA1 based on the type of finish
(galvanised or painted)
Appendix B.1 - Joint Stiffness Comparisons 251
______________________________________________________________________

170

Initial joint stiffness (kNm/rad) 150

130

110

90

70

50
Galvanised Painted
Materials

Figure B.1.16: Initial Cuplok joint stiffness for KzB1 based on the type of finish
(galvanised or painted)

170
Initial joint stiffness (kNm/rad)

150

130

110

90

70

50
Galvanised Painted
Materials

Figure B.1.17: Initial Cuplok joint stiffness for KzC1 based on the type of finish
(galvanised or painted)
Appendix B.1 - Joint Stiffness Comparisons 252
______________________________________________________________________

170

Initial joint stiffness (kNm/rad) 150

130

110

90

70

50
Galvanised Painted
Materials

Figure B.1.18: Initial Cuplok joint stiffness for KzD1 based on the type of finish
(galvanised or painted)

170
Initial joint stiffness (kNm/rad)

150

130

110

90

70

50
Galvanised Painted
Materials

Figure B.1.19: Initial Cuplok joint stiffness for KzA2 based on the type of finish
(galvanised or painted)
Appendix B.1 - Joint Stiffness Comparisons 253
______________________________________________________________________

170

Initial joint stiffness (kNm/rad) 150

130

110

90

70

50
Galvanised Painted
Materials

Figure B.1.20: Initial Cuplok joint stiffness for KzB2 based on the type of finish
(galvanised or painted)

170
Initial joint stiffness (kNm/rad)

150

130

110

90

70

50
Galvanised Painted
Materials

Figure B.1.21: Initial Cuplok joint stiffness for KzC2 based on the type of finish
(galvanised or painted)
Appendix B.1 - Joint Stiffness Comparisons 254
______________________________________________________________________

170

Initial joint stiffness (kNm/rad) 150

130

110

90

70

50
Galvanised Painted
Materials

Figure B.1.22: Initial Cuplok joint stiffness for KzD2 based on the type of finish
(galvanised or painted)

30
Initial joint stiffness (kNm/rad)

25

20

15

10

0
Galvanised Painted
Materials

Figure B.1.23: Initial Cuplok joint stiffness for KyA1 based on the type of finish
(galvanised or painted)
Appendix B.1 - Joint Stiffness Comparisons 255
______________________________________________________________________

30

Initial joint stiffness (kNm/rad) 25

20

15

10

0
Galvanised Painted
Materials

Figure B.1.24: Initial Cuplok joint stiffness for KyB1 based on the type of finish
(galvanised or painted)

30
Initial joint stiffness (kNm/rad)

25

20

15

10

0
Galvanised Painted
Materials

Figure B.1.25: Initial Cuplok joint stiffness for KyC1 based on the type of finish
(galvanised or painted)
Appendix B.1 - Joint Stiffness Comparisons 256
______________________________________________________________________

30

Initial joint stiffness (kNm/rad) 25

20

15

10

0
Galvanised Painted
Materials

Figure B.1.26: Initial Cuplok joint stiffness for KyD1 based on the type of finish
(galvanised or painted)

30
Initial joint stiffness (kNm/rad)

25

20

15

10

0
Galvanised Painted
Materials

Figure B.1.27: Initial Cuplok joint stiffness for KyC2 based on the type of finish
(galvanised or painted)
Appendix B.1 - Joint Stiffness Comparisons 257
______________________________________________________________________

30

Initial joint stiffness (kNm/rad) 25

20

15

10

0
Galvanised Painted
Materials

Figure B.1.28: Initial Cuplok joint stiffness for KyD2 based on the type of finish
(galvanised or painted)
Appendix B.2 - Probabilistic Model for Joint Stiffness 258
______________________________________________________________________

APPENDIX B.2 - PROBABILISTIC MODELS FOR


JOINT STIFFNESS

The data of joint stiffness (k1, k2, k3) for different configurations is normalised with the
mean value of its configuration and fitted to normal distribution.

0.3
Normal
0.25 mean = 1.00
Relative frequency

0.2 cov = 0.33

0.15
0.1
0.05
0
0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 More
Normalised k1

Figure B.2.1: Fitted normal distribution of normalised Cuplok joint stiffness k1


(looseness alone) in bending about the horizontal axis
Appendix B.2 - Probabilistic Model for Joint Stiffness 259
______________________________________________________________________

0.4
0.35 Normal
mean = 1.07

Relative frequency
0.3 cov = 0.2
0.25
0.2
0.15
0.1
0.05
0
0.6 0.8 1 1.2 1.4 1.6 More
Normalised k2

Figure B.2.2: Fitted normal distribution of normalised Cuplok joint stiffness k2 in


bending about the horizontal axis

0.25

0.2
Relative frequency

Normal
0.15 mean = 0.99
cov = 0.38
0.1

0.05

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 More
Normalised k3

Figure B.2.3: Fitted normal distribution of normalised Cuplok joint stiffness k3 in


bending about the horizontal axis
Appendix C.1 - Initial Imperfections of Subassembly Tests 260
______________________________________________________________________

APPENDIX C.1 - INITIAL IMPERFECTIONS OF


SUBASSEMBLY TESTS

It should be noted that this data is obtained from CASE (2006).

3
Platform Platform
W for for E
supporting 2 supporting
test rig test rig

A B C D

Figure C.1.1: Gridlines used for imperfection readings

Top
9

1 Bottom

Figure C.1.2: Measuring points along standards for imperfection readings


Appendix C.1 - Initial Imperfections of Subassembly Tests 261
______________________________________________________________________

Table C.1.1: Initial geometric imperfections (mm) for subassembly Test No. 2

Measurements in N-S direction:

N-S = measurements in North-South direction (assuming North as positive)

Note: Values are in mm according to positions in Figures C.1.1 and C.1.2, taken as
displacement from the bottom (point 1).

N-S D1 C1 B1 A1 D2 C2 B2 A2
9 14.4 3.6 0.0 21.0 2.7 -14.0 4.5 11.6
8 15.1 -1.2 -7.8 9.9 8.3 -14.8 -3.7 9.5
7 15.1 -5.5 -11.9 6.2 6.7 -14.8 -7.4 9.2
6 11.8 -7.1 -12.3 9.5 5.4 -15.5 -7.8 9.2
5 11.3 -7.1 -12.3 8.3 3.0 -16.5 -6.5 8.8
4 8.2 -3.2 -13.3 4.6 1.8 -14.0 -3.7 8.8
3 6.1 0.0 -11.0 4.6 0.6 -9.8 -2.9 5.5
2 2.3 1.2 -4.5 1.2 0.0 0.2 -2.9 -3.2
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

N-S D3 C3 B3 A3 D4 C4 B4 A4
9 14.3 -14.6 1.6 10.7 18.4 -6.3 4.1 3.2
8 16.1 -17.9 -5.3 8.8 14.4 -9.6 -3.7 1.9
7 15.7 -20.4 -8.8 7.9 14.4 -9.6 -5.3 2.8
6 14.4 -20.5 -10.7 7.6 14.4 -10.5 -6.6 2.8
5 12.8 -20.5 -11.1 7.4 11.7 -12.1 -4.9 2.8
4 9.4 -17.1 -9.0 7.0 8.4 -8.6 -5.5 2.8
3 6.1 -14.5 -8.2 6.7 6.8 -4.5 -1.6 0.5
2 2.9 -9.8 -0.8 1.9 2.3 -0.9 0.0 -1.4
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Appendix C.1 - Initial Imperfections of Subassembly Tests 262
______________________________________________________________________

Measurements in E-W direction:

E-W = measurements in East-West direction (assuming East as positive)

Note: Values are in mm according to positions in Figures C.1.1 and C.1.2, taken as
displacement from the bottom (point 1).

E-W D1 C1 B1 A1 D2 C2 B2 A2
9 2.3 -8.4 -2.4 0.0 -4.2 3.5 -1.1 2.7
8 -4.6 -8.9 -5.0 4.9 -3.3 4.9 -1.1 6.8
7 -6.5 -7.4 -3.2 9.2 -2.7 5.5 -1.5 8.7
6 -0.7 -4.0 -2.6 9.0 -2.2 6.4 -0.2 6.8
5 2.6 -1.5 1.2 10.3 -1.9 8.2 0.5 6.8
4 3.8 -0.2 0.2 12.6 -1.5 8.2 0.0 5.1
3 4.6 -2.8 -1.1 10.3 -2.2 6.1 -1.1 4.3
2 3.3 -3.4 -0.2 9.0 -1.6 3.9 -1.1 2.4
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

E-W D3 C3 B3 A3 D4 C4 B4 A4
9 -17.5 -14.4 -15.2 -7.1 -20.5 -2.9 -16.7 -13.7
8 -15.1 -8.4 -13.1 -8.5 -17.9 -6.6 -17.2 -13.7
7 -11.4 -6.7 -11.7 -5.0 -17.3 -7.0 -17.7 -10.8
6 -9.8 -6.2 -10.9 -5.6 -16.7 -4.6 -15.4 -10.3
5 -8.2 -4.8 -10.3 -5.9 -14.4 -3.1 -14.5 -10.1
4 -6.8 -3.6 -7.9 -3.2 -10.9 0.0 -11.2 -8.7
3 -6.5 -3.2 -6.3 -2.9 -8.9 -1.3 -8.3 -5.6
2 -4.2 -1.8 -2.4 0.0 -6.8 0.0 -8.3 -5.6
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Appendix C.1 - Initial Imperfections of Subassembly Tests 263
______________________________________________________________________

Table C.1.2: Initial geometric imperfections (mm) for subassembly Test No. 3

Measurements in N-S direction:

N-S = measurements in North-South direction (assuming North as positive)

Note: Values are in mm according to positions in Figures C.1.1 and C.1.2, taken as
displacement from the bottom (point 1).

N-S D1 C1 B1 A1 D2 C2 B2 A2
9 11.9 15.3 16.2 -7.0 27.7 12.8 17.6 -8.3
8 7.5 10.2 24.1 -19.7 21.3 10.3 14.3 -17.1
7 5.9 7.6 25.8 -20.6 15.9 10.7 11.4 -16.6
6 3.1 4.6 14.1 -21.5 9.5 5.3 9.0 -15.7
5 1.2 1.8 10.8 -15.9 5.0 3.2 2.0 -9.7
4 -0.3 0.5 4.4 -10.8 6.2 3.2 -3.7 -7.4
3 -1.6 -1.5 -0.4 -7.0 6.8 2.1 -6.5 -7.4
2 -2.2 -2.5 -2.7 -5.2 5.3 0.0 -7.4 -6.9
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

N-S D3 C3 B3 A3 D4 C4 B4 A4
9 36.5 15.4 41.6 -10.1 28.3 14.9 28.6 -17.7
8 36.2 16.7 39.6 -5.5 12.7 15.6 28.2 -22.4
7 26.3 9.9 35.5 -5.5 4.7 14.9 24.9 -17.7
6 20.2 7.5 29.4 -7.4 2.5 7.2 21.1 -14.7
5 15.4 3.2 24.9 -5.1 2.5 4.2 19.5 -12.6
4 13.0 2.8 16.7 -0.9 2.5 2.5 12.4 -9.3
3 9.4 1.8 12.7 0.5 -2.8 1.1 8.3 -7.9
2 3.3 1.4 2.4 0.0 -2.8 1.1 0.0 -4.7
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Appendix C.1 - Initial Imperfections of Subassembly Tests 264
______________________________________________________________________

Measurements in E-W direction:

E-W = measurements in East-West direction (assuming East as positive)

Note: Values are in mm according to positions in Figures C.1.1 and C.1.2, taken as
displacement from the bottom (point 1).

E-W D1 C1 B1 A1 D2 C2 B2 A2
9 -3.6 3.1 18.1 6.5 22.5 22.4 39.1 36.8
8 -3.1 4.3 12.2 -4.0 21.4 19.5 36.5 28.7
7 -0.7 3.5 10.2 -7.2 19.8 15.4 31.5 23.2
6 1.3 4.6 9.7 -2.7 17.7 17.2 25.0 20.9
5 2.2 4.3 9.9 0.5 13.1 14.3 21.3 20.2
4 2.2 5.1 6.4 2.2 7.1 5.6 13.5 16.6
3 1.8 2.5 3.1 1.4 3.7 1.3 6.5 12.0
2 -1.4 -1.2 0.3 2.5 -0.9 -3.0 0.7 6.7
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

E-W D3 C3 B3 A3 D4 C4 B4 A4
9 18.2 25.4 30.3 22.4 -0.7 -16.8 1.6 -9.0
8 17.9 14.3 25.4 21.0 -2.7 -14.2 -1.6 -9.3
7 15.7 10.8 20.5 19.4 -4.0 -14.2 -5.2 -9.6
6 11.5 11.1 15.9 16.8 -3.6 -12.6 -2.9 -7.0
5 9.2 11.1 11.9 14.6 -3.6 -12.6 -1.3 -9.0
4 6.2 8.4 7.8 12.3 -1.0 -13.9 -4.8 -7.6
3 5.0 3.8 1.6 9.0 0.0 -14.5 -5.2 -6.6
2 2.8 -0.3 -4.6 3.6 2.0 -11.0 -4.5 -3.3
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Appendix C.1 - Initial Imperfections of Subassembly Tests 265
______________________________________________________________________

Table C.1.3: Initial geometric imperfections (mm) for subassembly Test No. 4

Measurements in N-S direction:

N-S = measurements in North-South direction (assuming North as positive)

Note: Values are in mm according to positions in Figures C.1.1 and C.1.2, taken as
displacement from the bottom (point 1).

N-S D1 C1 B1 A1 D2 C2 B2 A2
9 11.2 8.8 -3.9 -8.9 14.6 -6.9 -11.1 -14.9
8 11.5 -2.4 -13.3 -9.8 16.3 -2.3 -15.0 -12.7
7 9.2 -12.2 -18.4 -12.4 16.0 -2.0 -18.4 -12.7
6 8.6 -3.4 -12.5 -11.5 14.6 -4.3 -18.8 -10.9
5 7.5 -0.3 -12.1 -11.1 13.2 -4.3 -19.2 -10.5
4 4.3 1.0 -13.5 -11.1 11.5 0.0 -13.8 -11.4
3 2.2 3.1 -10.6 -9.3 10.7 2.0 -5.4 -11.4
2 -0.3 2.4 -3.9 -4.9 6.3 1.3 -2.3 -10.5
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

N-S D3 C3 B3 A3 D4 C4 B4 A4
9 10.0 -5.0 -18.1 -13.1 7.8 -9.5 -23.5 0.0
8 11.1 -4.0 -17.3 -10.5 6.1 -7.8 -22.0 -7.1
7 11.4 -3.3 -16.9 -8.3 8.4 -4.8 -22.0 -7.1
6 10.0 -2.3 -23.1 -6.6 3.2 -3.4 -25.5 -6.7
5 7.2 -0.3 -23.1 -9.6 0.3 -2.7 -25.9 -6.7
4 5.6 1.0 -20.0 -12.7 0.0 0.3 -22.0 -4.4
3 4.5 3.0 -16.1 -9.2 -0.9 1.7 -17.6 -2.2
2 0.8 1.7 -5.0 -3.9 -2.0 2.4 -11.8 0.0
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Appendix C.1 - Initial Imperfections of Subassembly Tests 266
______________________________________________________________________

Measurements in E-W direction:

E-W = measurements in East-West direction (assuming East as positive)

Note: Values are in mm according to positions in Figures C.1.1 and C.1.2, taken as
displacement from the bottom (point 1).

E-W D1 C1 B1 A1 D2 C2 B2 A2
9 1.8 -10.9 4.7 5.7 -3.3 -2.2 -3.5 -15.7
8 1.5 -6.4 1.0 2.3 -12.7 -2.2 -3.1 -3.0
7 -0.4 -4.7 -2.9 -3.1 -12.7 -2.2 -2.9 0.7
6 -2.0 -9.4 -2.0 -5.1 -8.5 0.0 -2.0 -1.4
5 -2.4 -11.4 -2.2 -5.1 -6.1 -0.2 -2.2 -2.3
4 -0.4 -11.7 -2.7 -3.5 -6.6 -2.4 -1.5 -2.8
3 2.0 -10.7 -2.2 -2.5 -8.5 -2.4 -0.9 -2.8
2 3.3 -6.7 0.5 -2.3 -7.8 -2.7 -2.4 -3.7
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

E-W D3 C3 B3 A3 D4 C4 B4 A4
9 -9.4 -10.4 -7.4 -9.0 -26.6 -19.3 -22.2 -14.9
8 -7.7 -8.5 -9.6 -4.5 -16.2 -12.8 -20.2 -16.7
7 -5.4 -5.5 -9.6 -0.6 -8.8 -5.6 -15.7 -16.0
6 -3.4 -4.9 -4.9 0.0 -5.7 -4.3 -13.7 -10.0
5 -2.6 -4.9 -4.1 -2.0 -2.0 -2.3 -11.7 -7.3
4 0.3 -3.3 -5.5 -0.6 1.3 1.0 -3.9 -3.3
3 0.3 -3.3 -5.7 0.0 1.3 1.6 -5.9 -1.0
2 0.3 -2.7 -3.8 0.8 0.7 2.8 -2.6 -0.3
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Appendix C.1 - Initial Imperfections of Subassembly Tests 267
______________________________________________________________________

Table C.1.4: Initial geometric imperfections (mm) for subassembly Test No. 5

Measurements in N-S direction:

N-S = measurements in North-South direction (assuming North as positive)

Note: Values are in mm according to positions in Figures C.1.1 and C.1.2, taken as
displacement from the bottom (point 1).

N-S D1 C1 B1 A1 D2 C2 B2 A2
9 13.5 3.4 5.6 1.3 0.3 -5.7 -3.1 13.7
8 18.8 0.7 -6.7 0.9 7.6 -4.4 -4.3 7.1
7 15.8 1.0 -8.3 -3.1 8.5 -2.7 -4.7 6.6
6 11.7 4.1 -2.0 -2.7 3.1 -1.0 -2.7 5.8
5 9.7 5.2 -2.4 -3.6 -1.4 -1.0 -4.3 4.9
4 5.3 3.4 -3.2 -5.8 -0.6 -0.3 -3.1 4.9
3 3.8 3.1 -2.8 -4.9 -0.8 0.0 -2.3 5.8
2 1.5 0.7 -2.4 -2.7 -5.7 0.7 -0.8 5.3
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

N-S D3 C3 B3 A3 D4 C4 B4 A4
9 6.5 -16.5 -5.1 1.8 4.7 -7.2 -9.5 -8.1
8 6.5 -9.4 -11.3 0.9 9.1 -8.6 -9.1 -2.2
7 5.1 -5.7 -10.5 0.9 9.9 -7.9 -8.7 -2.7
6 4.0 -3.0 -8.2 0.0 5.3 -4.8 -5.9 -5.4
5 4.0 -2.7 -6.6 -1.3 4.7 -3.1 -4.8 -5.8
4 3.7 -5.0 -7.8 -1.8 4.7 0.7 -4.8 -6.3
3 1.1 -6.4 -6.2 -3.5 6.7 -3.1 -4.4 -6.3
2 -1.1 -4.7 -5.1 -3.5 1.8 -2.1 -2.0 -4.0
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Appendix C.1 - Initial Imperfections of Subassembly Tests 268
______________________________________________________________________

Measurements in E-W direction:

E-W = measurements in East-West direction (assuming East as positive)

Note: Values are in mm according to positions in Figures C.1.1 and C.1.2, taken as
displacement from the bottom (point 1).

E-W D1 C1 B1 A1 D2 C2 B2 A2
9 -14.7 -5.2 -0.5 6.1 -18.6 -24.5 -25.3 -9.9
8 -10.3 -2.8 -5.2 7.7 -19.1 -18.1 -23.3 -1.8
7 -9.8 0.5 -3.1 9.3 -16.8 -14.0 -21.2 0.2
6 -7.4 -0.5 -1.1 7.3 -12.5 -12.5 -13.6 2.3
5 -5.6 1.0 0.0 5.3 -9.3 -10.3 -10.8 1.8
4 -4.2 0.2 0.2 6.6 -9.1 -9.0 -10.6 5.1
3 -4.0 -2.1 -1.3 6.1 -7.3 -8.0 -9.7 7.3
2 -4.0 0.0 -1.5 3.4 -3.2 -5.8 -7.8 6.9
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

E-W D3 C3 B3 A3 D4 C4 B4 A4
9 -42.4 -38.6 -38.2 -24.4 -40.1 -29.5 -35.4 -35.1
8 -35.5 -30.5 -31.7 -28.3 -36.2 -26.0 -23.2 -30.8
7 -33.0 -24.4 -29.3 -24.6 -32.2 -20.2 -21.9 -24.2
6 -25.0 -21.7 -21.5 -17.4 -30.3 -19.3 -19.9 -20.9
5 -19.7 -16.3 -18.3 -12.6 -28.3 -16.7 -17.7 -17.9
4 -13.9 -11.0 -12.1 -9.5 -19.7 -9.6 -11.6 -13.9
3 -11.4 -6.4 -7.8 -6.2 -13.8 -4.5 -6.1 -9.9
2 -5.5 -3.5 -4.8 -1.7 -8.5 1.9 -1.0 -3.3
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Appendix C.1 - Initial Imperfections of Subassembly Tests 269
______________________________________________________________________

Table C.1.5: Initial geometric imperfections (mm) for subassembly Test No. 6

Measurements in N-S direction:

N-S = measurements in North-South direction (assuming North as positive)

Note: Values are in mm according to positions in Figures C.1.1 and C.1.2, taken as
displacement from the bottom (point 1).

N-S D1 C1 B1 A1 D2 C2 B2 A2
9 -12.2 -4.9 -3.6 -10.0 -23.6 -17.6 -15.5 -14.0
8 -11.6 -6.7 -6.0 -10.0 -18.1 -13.1 -6.8 -9.9
7 -9.3 -4.2 -8.9 -8.7 -16.9 -7.6 -6.4 -8.6
6 -8.7 -0.4 -5.6 -7.7 -14.3 -6.9 -6.0 -8.1
5 -7.5 0.7 -3.2 -5.9 -11.4 -6.2 -4.4 -6.8
4 -5.4 2.1 -4.8 -3.6 -11.1 -2.4 -6.0 -4.1
3 -2.7 3.2 -4.0 -2.3 -9.3 0.3 -4.4 -3.2
2 -2.1 4.9 -3.2 -3.6 -6.1 0.3 -3.6 -2.7
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

N-S D3 C3 B3 A3 D4 C4 B4 A4
9 -15.3 -22.6 -23.6 -11.8 -24.9 -25.1 -16.4 -24.5
8 -15.0 -12.9 -19.2 -11.8 -30.4 -21.5 -13.1 -20.3
7 -14.7 -12.2 -16.8 -9.5 -27.3 -19.3 -11.5 -19.8
6 -12.4 -12.2 -16.0 -5.4 -22.7 -15.0 -9.0 -9.7
5 -6.5 -10.4 -16.0 -4.5 -18.4 -11.5 -7.0 -10.6
4 -3.5 -8.0 -14.8 -3.2 -14.1 -8.6 -4.9 -8.3
3 -1.2 -5.6 -14.0 -0.9 -10.4 -7.2 -2.0 -6.0
2 2.4 -2.4 -11.6 0.0 -7.4 -2.9 0.8 -3.7
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Appendix C.1 - Initial Imperfections of Subassembly Tests 270
______________________________________________________________________

Measurements in E-W direction:

E-W = measurements in East-West direction (assuming East as positive)

Note: Values are in mm according to positions in Figures C.1.1 and C.1.2, taken as
displacement from the bottom (point 1).

E-W D1 C1 B1 A1 D2 C2 B2 A2
9 -4.4 -12.7 -16.2 -30.5 -3.7 -1.3 -12.4 -13.3
8 1.1 -12.7 -14.0 -11.7 -0.7 -1.8 -8.2 -14.5
7 1.7 -12.3 -12.8 -4.4 0.2 -1.6 -2.9 -15.7
6 0.6 -10.1 -12.3 -9.3 1.9 1.3 -4.4 -11.0
5 1.9 -7.6 -11.0 -10.2 2.3 3.8 -4.7 -8.7
4 1.1 -8.4 -7.4 -5.7 3.3 4.0 -2.2 -7.5
3 1.7 -9.8 -5.9 -2.8 3.0 4.4 -0.4 -6.3
2 1.1 -6.4 -5.4 -2.6 1.2 3.1 1.3 -5.9
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

E-W D3 C3 B3 A3 D4 C4 B4 A4
9 0.9 -16.7 -9.6 -16.5 -30.2 -31.8 -38.0 -47.1
8 -4.6 -17.3 -8.8 -14.2 -22.2 -25.2 -35.0 -40.3
7 -5.4 -17.6 -6.0 -12.2 -16.8 -22.6 -29.8 -35.3
6 -4.6 -18.9 -7.4 -11.4 -13.8 -21.3 -26.5 -26.9
5 -5.1 -18.7 -6.6 -10.5 -10.7 -18.0 -22.9 -22.9
4 -4.0 -15.6 -6.0 -9.7 -3.7 -9.8 -16.4 -16.8
3 -1.7 -13.5 -3.8 -8.8 0.0 -4.3 -11.5 -10.8
2 -1.7 -11.0 -0.5 -3.7 4.0 0.7 -4.3 -6.1
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Appendix C.1 - Initial Imperfections of Subassembly Tests 271
______________________________________________________________________

Table C.1.6: Initial geometric imperfections (mm) for subassembly Test No. 8

Measurements in N-S direction:

N-S = measurements in North-South direction (assuming North as positive)

Note: Values are in mm according to positions in Figures C.1.1 and C.1.2, taken as
displacement from the bottom (point 1).

N-S D1 C1 B1 A1 D2 C2 B2 A2
9 10.3 3.5 27.5 16.7 10.3 -1.0 23.5 18.7
8 10.9 2.8 21.1 17.6 6.6 2.0 26.2 16.0
7 13.0 3.1 18.7 18.0 4.6 2.0 23.9 13.3
6 8.6 0.0 17.1 18.5 3.7 0.3 18.4 9.3
5 5.3 2.4 14.7 12.6 1.7 -1.4 14.9 9.3
4 3.2 1.4 4.4 11.3 -1.4 -0.7 5.9 5.8
3 2.4 1.4 0.0 7.2 -1.4 1.4 2.7 3.1
2 1.5 1.0 0.0 2.7 -2.0 -0.7 1.2 0.0
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

N-S D3 C3 B3 A3 D4 C4 B4 A4
9 10.0 -14.2 21.2 16.0 -2.4 -10.4 13.2 14.5
8 10.0 -12.9 19.6 13.4 3.0 -7.6 20.4 10.4
7 7.7 -3.4 18.0 12.0 8.6 -7.6 20.8 9.5
6 6.0 -1.7 11.8 8.5 3.0 -5.6 13.6 6.3
5 4.3 -1.7 5.1 5.8 -0.6 -1.7 7.2 6.8
4 1.7 -1.7 3.1 4.5 0.6 -1.0 3.6 2.3
3 1.1 -1.7 0.8 3.6 0.3 -1.7 2.0 0.0
2 0.0 0.0 -0.8 0.0 0.0 0.0 0.0 0.0
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Appendix C.1 - Initial Imperfections of Subassembly Tests 272
______________________________________________________________________

Measurements in E-W direction:

E-W = measurements in East-West direction (assuming East as positive)

Note: Values are in mm according to positions in Figures C.1.1 and C.1.2, taken as
displacement from the bottom (point 1).

E-W D1 C1 B1 A1 D2 C2 B2 A2
9 0.0 -5.0 5.2 11.2 -11.8 -15.8 -13.1 -2.3
8 -5.8 -1.7 4.7 8.4 -13.5 -11.0 -7.9 -2.5
7 -10.9 1.2 0.5 3.3 -11.6 -7.9 -5.3 -1.4
6 -7.6 -4.2 -2.3 2.2 -8.6 -7.0 -6.1 -0.9
5 -6.2 -6.0 -1.5 1.6 -7.0 -5.7 -6.8 -0.9
4 -7.1 -4.0 0.0 1.6 -5.8 -4.4 -4.2 -1.2
3 -5.3 -2.5 1.2 0.0 -3.9 -2.8 -4.2 -2.3
2 -2.5 0.0 0.0 0.0 0.0 -0.7 -2.2 0.0
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

E-W D3 C3 B3 A3 D4 C4 B4 A4
9 -33.9 -36.2 -30.2 -25.4 -40.7 -31.5 -36.7 -34.3
8 -33.9 -34.0 -27.2 -24.8 -36.7 -28.3 -33.1 -29.0
7 -27.7 -29.4 -21.5 -19.7 -30.7 -25.0 -24.0 -22.0
6 -20.6 -23.4 -17.9 -14.1 -24.4 -21.8 -19.2 -16.0
5 -17.5 -19.6 -14.9 -11.8 -15.7 -15.3 -15.9 -12.0
4 -13.8 -15.8 -10.3 -6.2 -10.3 -8.8 -9.4 -10.7
3 -10.7 -12.5 -6.0 -2.5 -5.0 -3.9 -4.2 -6.0
2 0.0 -4.1 -2.2 0.0 0.0 0.0 0.0 0.0
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Appendix C.1 - Initial Imperfections of Subassembly Tests 273
______________________________________________________________________

Table C.1.7: Initial geometric imperfections (mm) for subassembly Test No. 9

Measurements in N-S direction:

N-S = measurements in North-South direction (assuming North as positive)

Note: Values are in mm according to positions in Figures C.1.1 and C.1.2, taken as
displacement from the bottom (point 1).

N-S D1 C1 B1 A1 D2 C2 B2 A2
9 -7.4 -5.4 -10.2 -4.4 -10.5 -9.4 -15.2 -4.4
8 -9.7 -4.4 -9.8 1.8 -2.0 -6.4 -8.6 -3.1
7 -9.5 -3.1 -7.8 3.6 -1.7 -3.7 -7.8 -2.7
6 -12.0 0.3 -9.8 0.4 -8.2 -2.0 -10.5 -2.2
5 -14.3 1.0 -10.6 0.9 -7.4 -0.3 -10.5 -0.9
4 -11.2 0.3 -8.2 2.2 -6.8 0.0 -7.0 -0.4
3 -5.7 0.3 -4.7 0.4 -4.0 -0.3 -2.3 0.0
2 -0.3 0.0 -0.4 0.0 -0.3 -0.3 0.8 0.0
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

N-S D3 C3 B3 A3 D4 C4 B4 A4
9 -8.1 0.0 0.0 -4.5 -3.4 -6.0 -11.8 -6.9
8 -7.8 -5.8 -15.4 -5.4 -4.3 -6.7 -12.2 -5.9
7 -7.5 -4.1 -14.2 -5.4 -3.4 -5.7 -12.6 -5.5
6 -7.2 -2.4 -13.0 -4.0 -6.1 -3.2 -11.8 -3.2
5 -7.2 -0.7 -11.8 -2.7 -5.5 0.0 -8.9 -0.5
4 -6.1 0.0 -11.4 -1.3 -4.0 -2.8 -6.9 -2.3
3 -2.9 2.4 -7.9 -0.4 -0.9 -2.8 -4.1 0.0
2 0.0 0.0 -3.2 0.0 -0.3 0.0 0.0 0.0
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Appendix C.1 - Initial Imperfections of Subassembly Tests 274
______________________________________________________________________

Measurements in E-W direction:

E-W = measurements in East-West direction (assuming East as positive)

Note: Values are in mm according to positions in Figures C.1.1 and C.1.2, taken as
displacement from the bottom (point 1).

E-W D1 C1 B1 A1 D2 C2 B2 A2
9 -9.1 -12.0 -11.3 -16.7 2.1 -7.4 -8.3 -15.7
8 -5.7 -13.4 -16.1 -12.1 -7.3 -7.2 -7.6 -10.7
7 -4.0 -13.4 -15.7 -9.5 -9.9 -6.7 -6.7 -9.0
6 -1.9 -9.3 -9.9 -7.1 -2.6 -5.4 -3.8 -8.1
5 1.9 -6.2 -7.2 -3.5 1.4 -3.4 -1.3 -4.8
4 4.5 -2.4 -2.9 0.2 1.6 -0.2 2.2 -1.4
3 4.0 -0.9 0.3 1.1 1.6 1.1 4.7 -0.5
2 0.2 0.0 0.0 0.0 0.0 1.3 2.9 0.0
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

E-W D3 C3 B3 A3 D4 C4 B4 A4
9 -11.1 -13.8 -16.6 -18.7 -3.7 -16.5 -12.6 -9.8
8 -12.6 -15.5 -16.6 -16.7 -3.7 -18.5 -10.2 -8.1
7 -11.4 -13.8 -13.6 -15.6 -2.4 -16.5 -9.9 -4.8
6 -7.4 -11.3 -11.4 -11.0 1.7 -13.5 -6.3 -4.4
5 -6.3 -10.8 -8.9 -5.2 3.7 -12.9 -3.3 -3.1
4 -4.9 -6.9 -5.8 -3.7 3.7 -8.2 -3.3 -1.4
3 -1.7 -3.3 -3.3 -3.5 3.7 -4.0 -3.3 -1.0
2 0.0 -1.9 -0.6 0.0 0.0 -2.3 0.0 0.0
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Appendix C.1 - Initial Imperfections of Subassembly Tests 275
______________________________________________________________________

Table C.1.8: Initial geometric imperfections (mm) for subassembly Test No. 10

Measurements in N-S direction:

N-S = measurements in North-South direction (assuming North as positive)

Note: Values are in mm according to positions in Figures C.1.1 and C.1.2, taken as
displacement from the bottom (point 1).

N-S D1 C1 B1 A1 D2 C2 B2 A2
9 -29.4 -14.9 0.0 5.9 -28.1 -14.0 0.0 5.5
8 -24.1 -10.8 5.2 11.3 -24.6 -12.6 4.4 11.4
7 -20.6 -8.3 4.0 14.5 -21.0 -10.5 3.2 14.6
6 -15.9 -11.5 2.0 8.2 -13.9 -11.2 2.0 7.7
5 -9.7 -10.4 4.8 4.1 -9.5 -10.1 0.0 5.5
4 -3.2 -6.9 0.8 -0.5 -4.1 -6.3 -2.0 0.9
3 -2.9 -4.9 0.0 0.0 0.0 -3.5 -2.0 1.4
2 -0.6 -1.0 0.0 0.0 1.8 -0.7 -3.6 -0.9
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

N-S D3 C3 B3 A3 D4 C4 B4 A4
9 0.0 0.0 9.4 10.2 -42.2 -18.8 8.5 0.0
8 0.0 -13.3 5.7 12.9 -41.5 -16.1 8.9 11.4
7 -23.3 -15.8 4.1 9.7 -19.0 -8.3 8.9 10.5
6 -18.1 -10.8 2.5 9.2 -12.4 -8.3 7.6 9.5
5 -15.4 -9.3 1.6 4.6 -3.9 -8.3 5.1 5.2
4 -4.6 -2.5 0.8 0.9 -5.6 -10.5 4.2 1.9
3 0.0 -4.3 0.0 0.5 -2.6 -7.5 3.0 1.0
2 0.0 -1.8 -1.6 -1.8 0.0 -4.1 0.0 0.0
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Appendix C.1 - Initial Imperfections of Subassembly Tests 276
______________________________________________________________________

Measurements in E-W direction:

E-W = measurements in East-West direction (assuming East as positive)

Note: Values are in mm according to positions in Figures C.1.1 and C.1.2, taken as
displacement from the bottom (point 1).

E-W D1 C1 B1 A1 D2 C2 B2 A2
9 -11.4 -23.7 -22.7 -21.3 -5.7 117.0 -15.3 -10.0
8 -11.2 -21.1 -21.4 -17.0 -7.6 118.8 -11.1 -3.7
7 -11.2 -17.6 -17.6 -11.7 -5.9 120.4 -9.6 -1.2
6 -10.3 -14.2 -14.4 -11.9 -5.0 124.4 -9.1 -4.0
5 -9.2 -11.0 -14.5 -11.7 -6.9 -6.9 -6.9 -3.5
4 -7.3 -7.8 -9.2 -7.9 -6.6 -6.5 -4.7 -1.2
3 -3.5 -4.1 -5.1 -3.6 -4.0 -3.6 -4.4 0.9
2 0.0 0.0 0.0 0.0 -1.4 -2.5 -2.2 0.0
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

E-W D3 C3 B3 A3 D4 C4 B4 A4
9 -4.6 -7.7 -4.4 -2.6 7.4 6.9 1.3 6.0
8 -4.3 -8.6 -8.8 2.3 8.8 3.3 5.3 6.4
7 -4.6 -9.1 -9.9 5.1 10.1 3.0 7.9 6.4
6 -4.3 -8.8 -6.9 4.0 4.1 1.0 6.9 4.0
5 -4.0 -6.9 -6.6 3.1 3.0 3.3 4.9 3.7
4 -4.6 -5.5 -4.7 3.7 -2.0 0.0 2.3 3.4
3 -4.6 -5.5 -3.9 2.0 -3.7 2.3 1.3 1.7
2 -4.0 -5.2 -1.9 0.0 -5.1 1.0 0.0 0.0
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Appendix C.1 - Initial Imperfections of Subassembly Tests 277
______________________________________________________________________

Table C.1.9: Initial geometric imperfections (mm) for subassembly Test No. 11

Measurements in N-S direction:

N-S = measurements in North-South direction (assuming North as positive)

Note: Values are in mm according to positions in Figures C.1.1 and C.1.2, taken as
displacement from the bottom (point 1).

N-S D1 C1 B1 A1 D2 C2 B2 A2
9 13.0 8.5 0.0 10.7 5.5 5.4 8.7 16.1
8 13.9 5.1 16.6 8.1 5.7 6.5 8.7 8.5
7 14.4 7.2 11.8 5.8 6.9 7.1 7.1 6.7
6 13.6 7.2 7.5 2.2 8.9 6.5 4.7 7.1
5 12.4 5.8 5.5 2.7 11.2 3.7 3.9 7.6
4 11.8 4.8 4.7 3.1 5.2 3.7 2.0 5.4
3 10.7 2.7 1.2 2.7 0.3 5.8 1.2 3.6
2 2.6 0.0 -2.4 2.2 0.0 1.4 0.0 3.1
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

N-S D3 C3 B3 A3 D4 C4 B4 A4
9 0.0 0.0 4.8 1.4 -1.9 -5.8 3.3 0.0
8 12.1 -3.1 1.2 5.0 -0.3 -6.5 0.0 0.0
7 9.2 -1.4 0.4 8.6 1.9 -3.3 -2.1 15.7
6 5.6 0.3 -0.8 3.6 1.9 -4.0 -1.2 9.7
5 2.7 2.1 -2.4 -1.4 0.3 -7.2 -0.4 6.0
4 0.6 -1.7 -4.0 1.4 0.3 -7.2 -4.5 3.2
3 0.3 -1.0 -5.6 1.4 0.3 -5.8 -4.1 2.8
2 0.0 -0.7 -3.6 0.9 -2.2 -1.8 -3.3 0.9
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Appendix C.1 - Initial Imperfections of Subassembly Tests 278
______________________________________________________________________

Measurements in E-W direction:

E-W = measurements in East-West direction (assuming East as positive)

Note: Values are in mm according to positions in Figures C.1.1 and C.1.2, taken as
displacement from the bottom (point 1).

E-W D1 C1 B1 A1 D2 C2 B2 A2
9 -3.5 -3.5 1.4 7.9 -9.8 -8.7 0.0 1.8
8 -10.5 -3.2 -0.3 4.3 -9.1 -6.4 -2.5 2.2
7 -16.9 -4.8 -4.3 -3.0 -8.8 -2.1 -3.4 5.6
6 -12.6 -7.5 -6.9 -6.1 -6.6 -4.7 -0.8 1.8
5 -9.3 -6.5 -8.0 -3.4 -6.1 -6.4 -1.9 1.3
4 -8.2 -2.2 -2.2 -1.1 -3.6 -4.5 -1.7 3.8
3 -7.0 0.6 -0.2 -1.8 -2.0 -4.0 -0.6 4.3
2 -3.3 0.6 -1.8 0.0 -0.2 -1.3 -0.2 2.2
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

E-W D3 C3 B3 A3 D4 C4 B4 A4
9 -25.5 -19.1 -16.2 -16.8 -13.8 -15.9 -12.1 -11.4
8 -17.7 -18.1 -18.8 -12.6 -18.7 -23.9 -15.6 -9.8
7 -8.6 -13.3 -18.5 -9.1 -18.0 -21.3 -14.3 -8.2
6 -9.7 -10.6 -10.3 -3.3 -14.4 -14.7 -9.9 -7.8
5 -8.9 -7.4 -6.9 -2.7 -12.5 -11.8 -9.9 -7.2
4 -3.3 -4.5 -5.3 0.3 -9.2 -8.9 -6.4 -4.6
3 -1.7 -3.2 -3.2 1.6 -5.6 -6.1 -4.5 -1.6
2 0.0 -0.5 -1.1 0.0 -2.0 -2.9 -1.6 0.3
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Appendix C.1 - Initial Imperfections of Subassembly Tests 279
______________________________________________________________________

Table C.1.10: Initial geometric imperfections (mm) for subassembly Test No. 12

Measurements in N-S direction:

N-S = measurements in North-South direction (assuming North as positive)

Note: Values are in mm according to positions in Figures C.1.1 and C.1.2, taken as
displacement from the bottom (point 1).

N-S D1 C1 B1 A1 D2 C2 B2 A2
9 17.7 -0.3 5.6 5.8 12.7 -0.7 10.7 6.3
8 18.6 0.7 5.6 2.7 14.8 1.0 10.7 2.7
7 18.0 1.0 5.6 3.6 13.6 1.4 10.3 2.7
6 11.3 -2.1 5.6 2.7 5.5 1.4 9.9 2.7
5 10.5 -3.1 2.8 0.0 7.0 1.4 6.7 -0.4
4 9.3 -3.8 -1.6 -3.1 5.8 -1.0 4.0 -0.4
3 6.7 -3.1 -4.0 -3.6 4.3 0.0 3.2 0.0
2 1.2 -1.7 0.0 0.0 1.7 1.0 1.6 0.0
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

N-S D3 C3 B3 A3 D4 C4 B4 A4
9 11.9 -7.0 5.6 2.7 5.7 -9.5 4.6 -9.3
8 14.6 -7.0 7.6 3.2 0.6 -10.2 4.6 -9.3
7 11.6 -7.0 8.0 2.3 -1.3 -10.2 5.0 -9.8
6 5.7 -7.0 4.4 2.3 -4.4 -11.0 4.2 -9.3
5 6.3 -4.2 3.6 1.4 -3.5 -8.8 3.3 -6.5
4 7.2 -3.5 2.8 0.0 -3.8 -6.6 0.8 -5.6
3 3.3 -3.5 0.4 2.3 -3.8 -3.7 0.0 -5.1
2 1.2 -3.2 -0.4 1.4 -2.5 -2.9 -1.7 0.0
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Appendix C.1 - Initial Imperfections of Subassembly Tests 280
______________________________________________________________________

Measurements in E-W direction:

E-W = measurements in East-West direction (assuming East as positive)

Note: Values are in mm according to positions in Figures C.1.1 and C.1.2, taken as
displacement from the bottom (point 1).

E-W D1 C1 B1 A1 D2 C2 B2 A2
9 -18.3 -14.6 -13.5 -10.4 -17.4 -18.8 -16.9 -16.7
8 -25.0 -13.2 -12.9 -7.2 -19.9 -17.2 -14.4 -15.2
7 -26.7 -11.5 -11.6 -6.0 -18.5 -16.3 -12.3 -11.2
6 -19.9 -12.6 -9.7 -4.4 -13.6 -15.9 -12.9 -11.0
5 -18.5 -11.7 -7.8 -4.4 -13.4 -13.4 -12.1 -11.9
4 -14.0 -6.8 -5.0 -2.8 -12.1 -10.5 -9.8 -9.7
3 -2.9 -3.0 -3.5 -0.4 -8.5 -8.0 -6.0 -6.6
2 -4.5 -0.5 -2.7 -0.2 -1.8 -3.1 -1.9 -2.2
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

E-W D3 C3 B3 A3 D4 C4 B4 A4
9 -18.9 -8.1 -13.3 -13.3 -26.9 -22.4 -20.7 -15.5
8 -17.8 -12.1 -15.7 -13.3 -26.3 -23.3 -19.8 -16.1
7 -15.9 -11.5 -15.7 -13.3 -21.4 -18.9 -17.9 -14.2
6 -13.7 -10.7 -15.2 -13.0 -16.2 -15.7 -14.1 -13.5
5 -10.9 -7.9 -12.8 -9.2 -12.7 -11.0 -11.0 -7.1
4 -6.8 -2.6 -8.9 -6.8 -8.4 -7.6 -5.0 -3.5
3 -4.6 2.1 -6.0 -4.9 -4.2 -3.5 -3.1 -0.6
2 0.0 1.0 -2.6 -1.6 0.0 -1.6 -1.3 0.0
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Appendix C.1 - Initial Imperfections of Subassembly Tests 281
______________________________________________________________________

Table C.1.11: Initial geometric imperfections (mm) for subassembly Test No. 13

Measurements in N-S direction:

N-S = measurements in North-South direction (assuming North as positive)

Note: Values are in mm according to positions in Figures C.1.1 and C.1.2, taken as
displacement from the bottom (point 1).

N-S D1 C1 B1 A1 D2 C2 B2 A2
9 19.0 1.7 2.4 -4.5 8.3 -6.5 -0.8 -0.9
8 19.3 -0.3 7.2 0.0 16.1 -3.1 5.1 -0.9
7 16.6 -2.4 11.2 0.0 15.5 -4.8 7.5 1.3
6 13.3 -0.3 10.8 0.9 12.6 -4.1 9.0 2.7
5 11.6 1.4 7.6 3.6 9.7 -1.7 9.0 2.7
4 9.8 2.1 6.0 4.5 9.5 -1.0 4.3 4.0
3 4.7 1.7 4.0 0.0 6.6 -1.0 0.8 3.6
2 1.2 -0.3 0.0 -0.9 1.1 -0.3 0.4 1.8
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

N-S D3 C3 B3 A3 D4 C4 B4 A4
9 8.9 -7.5 4.3 -1.8 -0.3 -10.1 -6.0 -10.4
8 8.9 -5.8 4.3 -4.0 4.2 -10.1 -2.8 -6.3
7 8.0 -4.4 7.1 -5.4 5.9 -9.1 2.0 -5.9
6 8.9 -3.4 10.2 -1.3 5.0 -8.4 3.6 -3.2
5 8.6 -1.0 9.0 -0.9 4.7 -7.7 2.8 0.9
4 5.2 -0.7 5.5 0.0 4.2 -3.8 1.2 0.5
3 1.1 -0.3 4.3 -0.4 3.0 -1.4 1.2 -0.5
2 -0.3 -0.3 2.0 -1.8 1.5 -0.3 -1.6 -0.5
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Appendix C.1 - Initial Imperfections of Subassembly Tests 282
______________________________________________________________________

Measurements in E-W direction:

E-W = measurements in East-West direction (assuming East as positive)

Note: Values are in mm according to positions in Figures C.1.1 and C.1.2, taken as
displacement from the bottom (point 1).

E-W D1 C1 B1 A1 D2 C2 B2 A2
9 -25.0 -14.7 -12.5 -15.3 -11.5 -6.5 -0.8 -1.7
8 -22.0 -16.2 -11.4 -7.8 -11.0 -4.0 -1.0 -1.7
7 -20.4 -15.0 -9.0 -5.7 -11.3 1.7 3.1 2.2
6 -17.4 -13.0 -5.7 -4.5 -4.5 3.1 5.8 3.7
5 -16.7 -13.0 -5.5 -3.7 -1.6 6.1 6.0 5.7
4 -15.0 -11.4 -5.4 -1.6 -2.0 4.4 7.1 5.0
3 -10.5 -7.1 -3.5 0.4 -3.2 3.1 5.0 2.8
2 -3.5 -2.3 -0.2 2.0 -2.5 0.6 1.7 0.7
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

E-W D3 C3 B3 A3 D4 C4 B4 A4
9 -31.3 -26.0 -31.0 -26.4 -40.7 -34.4 -36.1 -40.2
8 -26.9 -29.9 -25.3 -25.3 -35.2 -30.3 -30.7 -31.8
7 -24.5 -26.0 -20.8 -20.5 -29.6 -23.6 -25.1 -25.4
6 -19.5 -22.3 -18.0 -17.5 -21.2 -17.7 -17.9 -19.0
5 -17.0 -17.3 -14.9 -14.8 -15.3 -12.0 -11.9 -12.5
4 -8.2 -11.5 -7.6 -7.3 -11.1 -7.2 -7.8 -9.0
3 -5.2 -8.7 -3.1 -1.9 -6.8 -4.4 -2.8 -5.8
2 -0.5 -3.7 -1.6 -1.3 -1.6 0.9 0.0 0.0
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Appendix C.1 - Initial Imperfections of Subassembly Tests 283
______________________________________________________________________

Table C.1.12: Initial geometric imperfections (mm) for subassembly Test No. 14

Measurements in N-S direction:

N-S = measurements in North-South direction (assuming North as positive)

Note: Values are in mm according to positions in Figures C.1.1 and C.1.2, taken as
displacement from the bottom (point 1).

N-S D1 C1 B1 A1 D2 C2 B2 A2
9 -5.7 1.4 -6.2 8.4 -5.8 -2.3 -11.9 -2.2
8 -3.4 0.7 -9.4 12.8 -5.8 -2.3 -3.1 1.3
7 2.3 2.7 -9.0 12.8 -2.2 -1.0 2.3 4.4
6 4.0 3.0 -6.2 5.3 3.6 0.3 0.8 7.4
5 7.2 2.0 -6.2 -1.3 7.8 -0.3 0.0 3.1
4 4.9 1.4 -5.9 -2.2 5.0 -2.0 -1.2 -1.3
3 1.7 0.0 -2.7 -1.3 3.6 -2.0 -3.1 -0.4
2 0.0 -1.0 -1.6 0.0 0.0 -2.0 -2.7 0.0
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

N-S D3 C3 B3 A3 D4 C4 B4 A4
9 -7.2 2.3 -4.6 11.4 -4.6 1.4 0.4 14.6
8 -1.7 -0.3 -4.6 11.4 -4.6 3.1 -3.5 12.4
7 1.1 1.3 -4.2 9.6 -3.5 5.8 -3.9 12.0
6 3.1 2.6 -3.8 3.9 1.7 2.0 -3.9 8.4
5 7.2 2.6 -5.0 -0.4 3.8 -1.7 -5.1 2.7
4 5.3 0.0 -3.5 -4.4 2.9 -6.5 -7.1 -1.8
3 1.1 0.0 -0.4 -4.4 1.4 -4.1 -4.3 -0.4
2 0.0 0.0 0.0 -1.3 -0.3 -1.0 0.0 1.8
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Appendix C.1 - Initial Imperfections of Subassembly Tests 284
______________________________________________________________________

Measurements in E-W direction:

E-W = measurements in East-West direction (assuming East as positive)

Note: Values are in mm according to positions in Figures C.1.1 and C.1.2, taken as
displacement from the bottom (point 1).

E-W D1 C1 B1 A1 D2 C2 B2 A2
9 0.5 -1.3 -0.8 0.9 1.8 4.2 8.2 6.9
8 -1.6 -2.2 -0.8 8.2 6.0 9.7 11.1 16.0
7 -4.7 -0.9 -1.6 10.2 6.0 10.1 13.2 20.0
6 0.3 0.3 0.0 6.5 3.4 7.6 9.0 12.5
5 3.0 0.3 1.7 4.9 1.6 6.9 7.1 7.3
4 2.1 0.6 0.2 4.2 -1.6 2.7 5.2 4.2
3 0.2 0.5 -1.7 2.6 -4.7 0.4 0.8 2.4
2 -1.0 -2.2 -1.6 1.2 -3.1 -0.6 -0.6 0.4
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

E-W D3 C3 B3 A3 D4 C4 B4 A4
9 -16.2 -15.0 -15.2 -9.0 15.0 9.5 6.9 11.4
8 -16.7 -17.1 -13.1 -8.7 14.0 12.3 11.1 12.0
7 -14.0 -15.8 -11.0 -8.5 13.0 12.3 13.3 12.7
6 -11.2 -11.0 -10.5 -10.7 4.2 2.2 1.3 4.5
5 -6.3 -8.4 -11.0 -10.9 -3.6 -6.0 -8.8 -2.3
4 -4.1 -5.8 -5.0 -6.0 -10.1 -8.5 -6.9 -7.1
3 -2.7 -3.7 -2.4 -2.7 -7.8 -5.4 0.0 -2.9
2 0.0 -1.6 -0.3 -1.4 0.0 0.0 -1.6 -0.6
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Appendix C.1 - Initial Imperfections of Subassembly Tests 285
______________________________________________________________________

Table C.1.13: Initial geometric imperfections (mm) for subassembly Test No. 15

Measurements in N-S direction:

N-S = measurements in North-South direction (assuming North as positive)

Note: Values are in mm according to positions in Figures C.1.1 and C.1.2, taken as
displacement from the bottom (point 1).

N-S D1 C1 B1 A1 D2 C2 B2 A2
7 6.5 -9.3 0.8 -17.9 12.5 -4.9 -4.5 -9.1
6 4.5 -11.7 -1.5 -16.6 10.1 -6.5 -3.0 -11.2
5 5.6 -9.3 -5.0 -13.1 4.9 -3.6 2.3 -13.4
4 3.1 -3.0 -1.2 -7.9 7.6 2.9 0.4 -9.1
3 -0.3 -0.3 1.2 -4.4 6.3 2.9 0.8 -3.5
2 -0.3 0.0 0.0 -2.2 2.5 0.3 0.0 -3.9
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

N-S D3 C3 B3 A3 D4 C4 B4 A4
7 13.9 -9.8 -3.4 -17.3 1.4 -5.7 -15.5 -19.3
6 13.4 -3.3 -3.4 -13.4 -1.4 -10.4 -11.6 -20.2
5 10.1 0.0 0.4 -9.1 -2.3 -11.4 -3.5 -20.2
4 7.1 2.0 0.8 -8.6 -2.3 1.3 -3.9 -11.9
3 4.9 1.6 0.8 -6.1 -3.4 5.4 -3.5 -6.2
2 1.4 0.0 0.4 -1.7 -2.8 0.0 -0.8 -2.2
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Appendix C.1 - Initial Imperfections of Subassembly Tests 286
______________________________________________________________________

Measurements in E-W direction:

E-W = measurements in East-West direction (assuming East as positive)

Note: Values are in mm according to positions in Figures C.1.1 and C.1.2, taken as
displacement from the bottom (point 1).

E-W D1 C1 B1 A1 D2 C2 B2 A2
7 12.5 19.1 15.1 8.0 -3.7 4.2 3.2 0.0
6 10.9 17.8 16.4 9.7 -2.9 3.1 4.2 2.0
5 6.3 14.6 16.1 11.3 -0.4 -0.2 0.4 2.5
4 3.2 4.5 5.3 2.4 -6.2 -2.1 0.2 -1.8
3 1.8 -1.3 0.6 1.0 -6.8 -2.3 0.4 -1.8
2 0.0 -1.3 -1.1 1.2 -4.2 -0.6 -1.3 -1.1
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

E-W D3 C3 B3 A3 D4 C4 B4 A4
7 -7.0 -6.0 -4.5 -8.8 -40.9 -38.7 -40.9 -33.1
6 -6.5 -3.4 -4.0 -7.2 -37.7 -35.3 -34.5 -33.4
5 -8.7 -3.7 -4.0 -3.0 -28.4 -25.8 -28.2 -29.1
4 -7.3 -5.8 -4.5 -6.4 -24.2 -18.0 -20.6 -16.4
3 -4.9 -3.7 -3.7 -6.1 -18.0 -10.7 -10.5 -6.9
2 -1.9 -1.0 -1.8 -1.1 -5.2 -1.3 -3.2 0.0
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Appendix C.1 - Initial Imperfections of Subassembly Tests 287
______________________________________________________________________

Table C.1.14: Initial geometric imperfections (mm) for subassembly Test No. 16

Measurements in N-S direction:

N-S = measurements in North-South direction (assuming North as positive)

Note: Values are in mm according to positions in Figures C.1.1 and C.1.2, taken as
displacement from the bottom (point 1).

N-S D1 C1 B1 A1 D2 C2 B2 A2
9 -8.7 -1.1 -0.4 -11.0 -11.1 -13.4 -2.4 -10.6
8 -8.4 -5.0 -1.2 -8.7 -4.2 -9.5 2.0 -11.0
7 -9.0 -2.1 -1.2 -8.3 -5.4 -6.0 2.0 -10.6
6 -5.7 2.5 -0.4 -8.3 -6.0 -0.7 4.1 -10.1
5 -4.5 3.5 0.8 -6.4 -4.5 0.4 4.5 -9.2
4 -4.8 4.3 1.2 -5.1 -3.9 1.4 4.9 -7.3
3 -5.7 4.6 0.8 -4.1 -3.9 1.4 4.9 -7.3
2 -5.7 2.1 0.0 -1.8 -3.9 0.7 2.4 -3.7
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

N-S D3 C3 B3 A3 D4 C4 B4 A4
9 -1.8 -6.1 -5.8 -3.7 -12.6 -11.5 -8.9 -9.5
8 1.5 -6.1 -5.8 -3.7 -12.6 -7.8 -8.9 -9.5
7 1.8 -6.1 -4.9 -3.7 -11.6 -5.2 -8.5 -10.4
6 1.5 -5.7 -4.1 -3.7 -9.7 -1.1 -8.0 -8.5
5 1.2 -5.7 -4.1 -2.8 -8.4 0.4 -8.0 -5.2
4 1.2 -4.3 -3.7 0.5 -6.1 3.0 -7.6 -3.8
3 1.5 -4.3 -3.7 0.5 -6.1 3.0 -8.0 -3.3
2 0.6 -2.2 -1.6 0.0 -5.8 1.5 -3.4 -2.8
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Appendix C.1 - Initial Imperfections of Subassembly Tests 288
______________________________________________________________________

Measurements in E-W direction:

E-W = measurements in East-West direction (assuming East as positive)

Note: Values are in mm according to positions in Figures C.1.1 and C.1.2, taken as
displacement from the bottom (point 1).

E-W D1 C1 B1 A1 D2 C2 B2 A2
9 8.8 8.3 22.4 10.2 -4.0 -2.0 -10.3 -7.8
8 15.6 9.1 17.8 13.2 -3.8 -4.6 -11.8 -9.3
7 18.5 7.8 18.3 13.9 -1.8 -4.6 -10.0 -10.5
6 13.6 3.8 16.5 10.6 -1.6 -4.6 -9.0 -13.1
5 12.1 1.8 13.5 8.7 -1.8 -6.0 -8.5 -7.4
4 10.5 -1.0 9.6 5.7 -3.8 -5.2 -7.5 -1.0
3 8.8 -3.1 7.9 3.6 -3.6 -5.4 -6.6 0.2
2 6.6 -5.1 4.7 0.0 -2.0 -4.4 -5.6 1.0
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

E-W D3 C3 B3 A3 D4 C4 B4 A4
9 -22.7 -20.3 -15.1 -17.3 -21.5 -22.7 -30.9 -39.5
8 -19.6 -15.8 -14.3 -10.8 -19.1 -16.0 -24.0 -22.6
7 -18.1 -14.7 -13.5 -10.5 -15.4 -8.9 -20.8 -19.9
6 -16.8 -13.5 -13.5 -11.4 -12.0 -7.4 -17.7 -17.9
5 -16.3 -13.0 -12.7 -10.5 -8.9 -5.8 -15.5 -14.9
4 -15.5 -13.0 -11.1 -6.8 -6.5 -2.8 -14.5 -12.0
3 -16.6 -10.9 -8.5 -5.4 -4.6 -1.2 -11.7 -9.3
2 -16.8 -7.1 -5.8 -2.8 -2.5 0.9 -6.0 -4.7
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Appendix C.1 - Initial Imperfections of Subassembly Tests 289
______________________________________________________________________

Table C.1.15: Initial geometric imperfections (mm) for subassembly Test No. 18

Measurements in N-S direction:

N-S = measurements in North-South direction (assuming North as positive)

Note: Values are in mm according to positions in Figures C.1.1 and C.1.2, taken as
displacement from the bottom (point 1).

N-S D1 C1 B1 A1 D2 C2 B2 A2
11 -5.3 6.9 5.5 3.1 -5.6 11.0 6.6 7.5
10 -7.9 7.2 5.1 4.5 -7.0 10.7 12.4 7.5
9 -8.8 7.2 4.3 3.1 -8.4 10.7 10.9 7.9
8 -9.9 7.2 4.3 3.1 -8.4 10.7 8.9 7.5
7 -10.8 7.2 4.3 3.6 -8.4 10.0 8.9 7.5
6 -8.2 6.9 1.2 4.9 -7.9 9.4 7.0 7.5
5 -6.4 5.8 0.4 4.5 -5.9 7.7 5.8 6.2
4 -4.4 3.4 0.4 4.0 -4.2 5.0 3.5 5.3
3 -2.9 1.0 0.4 2.2 -2.5 4.0 2.7 4.8
2 -0.9 0.3 0.8 0.0 -0.6 2.3 1.6 4.4
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

N-S D3 C3 B3 A3 D4 C4 B4 A4
11 -8.4 3.3 5.4 -3.1 -9.3 1.7 2.8 -4.0
10 -8.7 4.0 5.4 -2.6 -7.0 5.1 3.2 0.4
9 -8.7 9.0 4.7 0.0 -7.0 7.9 3.2 0.9
8 -10.4 8.7 4.7 0.4 -7.3 8.2 3.2 2.2
7 -11.0 8.7 5.4 0.4 -7.9 8.2 3.9 3.1
6 -8.7 8.4 3.9 -1.8 -9.3 7.5 0.8 1.3
5 -6.5 6.7 1.2 -1.8 -9.3 5.1 0.8 1.3
4 -5.6 5.4 0.8 -3.1 -6.7 4.5 0.0 1.3
3 -3.9 4.7 0.4 -2.6 -3.8 2.4 -0.8 2.2
2 -1.1 1.7 -1.9 -2.6 -1.7 1.7 -1.6 0.4
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Appendix C.1 - Initial Imperfections of Subassembly Tests 290
______________________________________________________________________

Measurements in E-W direction:

E-W = measurements in East-West direction (assuming East as positive)

Note: Values are in mm according to positions in Figures C.1.1 and C.1.2, taken as
displacement from the bottom (point 1).

E-W D1 C1 B1 A1 D2 C2 B2 A2
11 1.7 2.2 3.4 6.3 -4.6 -11.1 -9.3 -4.2
10 3.1 0.6 4.5 4.7 -5.5 -10.2 -2.3 -4.2
9 3.6 1.0 3.5 3.6 -4.1 -9.0 -1.7 -2.0
8 3.8 0.3 1.3 3.6 -2.5 -8.8 0.0 0.4
7 4.1 0.3 0.0 3.6 -1.6 -7.7 1.3 1.6
6 1.7 -0.2 0.8 3.6 -0.9 -6.4 1.1 0.4
5 0.3 -0.3 0.0 2.9 -2.1 -5.1 0.0 0.2
4 -1.2 0.0 -1.3 2.2 -3.0 -5.1 0.8 -0.4
3 -1.2 0.2 -0.8 0.5 -3.0 -5.3 1.3 0.2
2 -1.2 -1.3 -0.5 0.4 -1.6 -3.4 1.9 0.2
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

E-W D3 C3 B3 A3 D4 C4 B4 A4
11 -27.8 -30.1 -25.1 -24.6 -16.8 -14.3 -14.3 -12.3
10 -22.8 -28.2 -24.3 -23.2 -15.1 -13.1 -12.1 -12.3
9 -21.1 -26.6 -23.5 -21.0 -13.2 -12.8 -9.8 -12.3
8 -18.6 -24.5 -20.1 -19.4 -9.5 -9.6 -8.6 -9.7
7 -15.8 -22.3 -18.2 -17.2 -6.3 -7.3 -7.3 -7.8
6 -13.9 -19.2 -15.9 -15.6 -6.3 -4.5 -4.1 -5.8
5 -11.4 -16.0 -13.5 -14.2 -6.9 -3.5 -2.5 -3.9
4 -8.1 -11.2 -10.0 -10.7 -6.6 -1.6 -0.6 -2.6
3 -5.6 -8.5 -8.2 -7.7 -4.6 0.3 0.0 -2.3
2 1.9 -4.8 -4.2 -1.6 0.0 1.3 0.0 -0.3
1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Appendix D.1 - Typical Strand7 API Subroutine for Monte Carlo Simulations 291
______________________________________________________________________

APPENDIX D.1 TYPICAL STRAND7 API


SUBROUTINE FOR MONTE CARLO
SIMULATIONS

The following typical API subroutine is written in C++ language that works together
with Strand7 nonlinear finite element model to perform Monte Carlo simulations using
advanced structural analyses in Strand7 software. To compile the subroutine into an
executable file, it requires some associated files (St7APIConst.h, St7APICall.h,
API_code_old.h and St7APILoad.cpp) that are provided in the API folder of the
software (Strand7 2009). Also, this subroutine needs a Boost library that can be
downloaded from http://www.boost.org to generate random values based on statistical
distributions. This subroutine was used for a single (1.829 m x 1.829 m) bay, 1-m lift
height, 3-lifts, 300-mm top and bottom jack extensions scaffold frame.

#include <atlstr.h>
#include "St7APIConst.h"
#include "St7APICall.h"
#include "API_code_old.h"
#include <iostream>
#include <string>
#include <fstream>
#include <Ctime>
#include <math.h>
#include <time.h>

//The following header files are for the random functions of the Boost library

#include <boost/random/linear_congruential.hpp>
#include <boost/random/uniform_real.hpp>
#include <boost/random/uniform_01.hpp>
#include <boost/random/normal_distribution.hpp>
Appendix D.1 - Typical Strand7 API Subroutine for Monte Carlo Simulations 292
______________________________________________________________________

#include <boost/random/lognormal_distribution.hpp>
#include <boost/random/variate_generator.hpp>
#include <boost/random.hpp>
using namespace std;
//typedef boost::mt19937 base_generator_type;
typedef boost::minstd_rand base_generator_type;//switch between mt19937 and
minstd_rand for random generator

int main()
{
if (!LoadSt7API())
{
cout<<"Failed to load St7api.dll. This program will now exit."<<endl;
exit(1);
}
long iErrInit;
iErrInit=St7Init();
long FileUnit=1;
if (St7OpenFile(FileUnit,
"API_nonLinear_3lifts_1.0m_300mmext_1.829mx1.829m_1x1bay.st7","C:\\Program
Files\\Strand7-R24-Beta\\Bin")==0)
cout<<"File open successfully"<<endl;
else
cout<<"File open error"<<endl;
St7GetTotal(FileUnit, tyNODE, &FNodeTotal);//total number of nodes
St7GetTotal(FileUnit, tyBEAM, &FBeamTotal);//total number of beam
elements
St7GetNumLoadCase(FileUnit, &FNumLCases);//total number of load cases
long PropNum;
long Entity;
long NewPropNum;
long NumProps[kMaxEntityTotals], LastProp[kMaxEntityTotals];
Appendix D.1 - Typical Strand7 API Subroutine for Monte Carlo Simulations 293
______________________________________________________________________

St7GetTotalProperties(FileUnit, NumProps, LastProp);


FBeamPropertyTotal=NumProps[ipBeamPropTotal];//total number of beam
properties
base_generator_type generator(static_cast<unsigned int>(std::time(0)));//set
seed for random number generator
boost::uniform_real<> uniform_sign(0,1);
boost::normal_distribution<> normal_dist_2wayK1(41000, 14350);
boost::normal_distribution<> normal_dist_2wayK2(77000, 15400);
boost::normal_distribution<> normal_dist_2wayK3(4600, 2116);
boost::normal_distribution<> normal_dist_3wayK1(36000, 13680);
boost::normal_distribution<> normal_dist_3wayK2(87000, 18270);
boost::normal_distribution<> normal_dist_3wayK3(5100, 1887);
boost::normal_distribution<> normal_dist_4wayK1(39000, 8580);
boost::normal_distribution<> normal_dist_4wayK2(102000, 18360);
boost::normal_distribution<> normal_dist_4wayK3(5300, 1590);
boost::normal_distribution<> normal_dist_E(2E+11, 12E+9);//Young's modulus
for bending stress only
boost::lognormal_distribution<> lognormal_StandardI(137675.8,
6883.79);//Standard I, COV=0.05
boost::normal_distribution<> normal_StandardFy(4.95E+8,
0.495E+8);//Standard Fy(mean)=1.1*Fy(nominal)=1.1*450=495MPa, COV=0.1, i.e.
sigma=49.5MPa
boost::lognormal_distribution<> lognormal_JackI(82447.96, 4122.4);//Jack I,
COV=0.05
boost::normal_distribution<> normal_JackFy(4.73E+8, 0.473E+8);//Jack
Fy(mean)=1.1 *Fy(nominal)=1.1*430=473MPa, COV=0.1, i.e. sigma=47.3MPa
boost::normal_distribution<> normal_dist_OutOfPlumbX(0.0016, 5E-4);//Out
of plumb,
boost::normal_distribution<> normal_dist_OutOfPlumbY(0.0016, 5E-4);//Out
of plumb,
boost::lognormal_distribution<> lognormal_dist_OutOfStraightnessX(0.0013,
8E-4);//Out of straightness with spigot
Appendix D.1 - Typical Strand7 API Subroutine for Monte Carlo Simulations 294
______________________________________________________________________

boost::lognormal_distribution<> lognormal_dist_OutOfStraightnessY(0.0013,
8E-4);//Out of straightness with spigot
boost::lognormal_distribution<>
lognormal_dist_OutOfStraightnessXnoSpigot(0.0004, 3E-4);//Out of straightness no
spigot
boost::lognormal_distribution<>
lognormal_dist_OutOfStraightnessYnoSpigot(0.0004, 3E-4);//Out of straightness no
spigot
boost::lognormal_distribution<> lognormal_loadEcc(18.1,11);//
boost::lognormal_distribution<>
lognormal_StandardA(556.69,27.8345);//Standard cross sectional area, COV=0.05
boost::lognormal_distribution<> lognormal_JackA(1017.876,50.8938);//Jack
cross sectional area, COV=0.05
boost::variate_generator<base_generator_type&, boost::normal_distribution<>
> norm2wayK1(generator, normal_dist_2wayK1);
boost::variate_generator<base_generator_type&, boost::normal_distribution<>
> norm2wayK2(generator, normal_dist_2wayK2);
boost::variate_generator<base_generator_type&, boost::normal_distribution<>
> norm2wayK3(generator, normal_dist_2wayK3);
boost::variate_generator<base_generator_type&, boost::normal_distribution<>
> norm3wayK1(generator, normal_dist_3wayK1);
boost::variate_generator<base_generator_type&, boost::normal_distribution<>
> norm3wayK2(generator, normal_dist_3wayK2);
boost::variate_generator<base_generator_type&, boost::normal_distribution<>
> norm3wayK3(generator, normal_dist_3wayK3);
boost::variate_generator<base_generator_type&, boost::normal_distribution<>
> norm4wayK1(generator, normal_dist_4wayK1);
boost::variate_generator<base_generator_type&, boost::normal_distribution<>
> norm4wayK2(generator, normal_dist_4wayK2);
boost::variate_generator<base_generator_type&, boost::normal_distribution<>
> norm4wayK3(generator, normal_dist_4wayK3);
boost::variate_generator<base_generator_type&, boost::normal_distribution<>
> normE(generator, normal_dist_E);
Appendix D.1 - Typical Strand7 API Subroutine for Monte Carlo Simulations 295
______________________________________________________________________

boost::variate_generator<base_generator_type&,
boost::lognormal_distribution<> > lognormStandardI(generator, lognormal_StandardI);
boost::variate_generator<base_generator_type&, boost::normal_distribution<>
> StandardFy(generator, normal_StandardFy);
boost::variate_generator<base_generator_type&,
boost::lognormal_distribution<> > lognormJackI(generator, lognormal_JackI);
boost::variate_generator<base_generator_type&, boost::normal_distribution<>
> JackFy(generator, normal_JackFy);
boost::variate_generator<base_generator_type&, boost::normal_distribution<>
> normOutOfPlumbX(generator, normal_dist_OutOfPlumbX);
boost::variate_generator<base_generator_type&, boost::normal_distribution<>
> normOutOfPlumbY(generator, normal_dist_OutOfPlumbY);
boost::variate_generator<base_generator_type&,
boost::lognormal_distribution<> > lognormOutOfStraightnessX(generator,
lognormal_dist_OutOfStraightnessX);
boost::variate_generator<base_generator_type&,
boost::lognormal_distribution<> > lognormOutOfStraightnessY(generator,
lognormal_dist_OutOfStraightnessY);
boost::variate_generator<base_generator_type&,
boost::lognormal_distribution<> > lognormOutOfStraightnessXnoSpigot(generator,
lognormal_dist_OutOfStraightnessXnoSpigot);
boost::variate_generator<base_generator_type&,
boost::lognormal_distribution<> > lognormOutOfStraightnessYnoSpigot(generator,
lognormal_dist_OutOfStraightnessYnoSpigot);
boost::variate_generator<base_generator_type&,
boost::lognormal_distribution<> > lognormStandardA(generator,
lognormal_StandardA);
boost::variate_generator<base_generator_type&,
boost::lognormal_distribution<> > lognormJackA(generator, lognormal_JackA);
boost::variate_generator<base_generator_type&,
boost::lognormal_distribution<> > lognormEcc(generator, lognormal_loadEcc);
boost::variate_generator<base_generator_type&, boost::uniform_real<> >
uniformSign(generator, uniform_sign);
Appendix D.1 - Typical Strand7 API Subroutine for Monte Carlo Simulations 296
______________________________________________________________________

long TableIDMomentRotation2Way;
long TableIDMomentRotation3Way;
long TableIDMomentRotation4Way;
long TableIDMomentRotation2WayHorizontal;
long TableIDMomentRotation3WayHorizontal;
long TableIDMomentRotation4WayHorizontal;
long TableIDStressStrainStandard;
long TableIDStressStrainJack;

St7GetTableID(FileUnit, "Moment vs Rotation ZA", ttMomentRotation,


&TableIDMomentRotation4Way);
St7GetTableID(FileUnit, "Moment vs Rotation ZB", ttMomentRotation,
&TableIDMomentRotation3Way);
St7GetTableID(FileUnit, "Moment vs Rotation ZD", ttMomentRotation,
&TableIDMomentRotation2Way);
St7GetTableID(FileUnit, "Moment vs Rotation YA", ttMomentRotation,
&TableIDMomentRotation4WayHorizontal);
St7GetTableID(FileUnit, "Moment vs Rotation YB", ttMomentRotation,
&TableIDMomentRotation3WayHorizontal);
St7GetTableID(FileUnit, "Moment vs Rotation YD", ttMomentRotation,
&TableIDMomentRotation2WayHorizontal);
St7GetTableID(FileUnit, "Stress vs Strain for Standard (Design using Ramberg-
Osgood)", ttStressStrain, &TableIDStressStrainStandard);
St7GetTableID(FileUnit, "Stress vs Strain for Steel Jack (Design using
Ramberg-Osgood)", ttStressStrain, &TableIDStressStrainJack);

//Table related data

long NewTableID, NumEntries;


long NewStandardFyTableID;
long NewJackFyTableID;
NewTableID=100;
Appendix D.1 - Typical Strand7 API Subroutine for Monte Carlo Simulations 297
______________________________________________________________________

NewStandardFyTableID=500;
NewJackFyTableID=800;
int NumMC;//number of Monte Carlo sampling
NumMC=3000;//*****************number of Monte Carlo
simulations*********************
long Solver, Mode, Wait;
//Solver=stLinearStaticSolver;
Solver=stNonlinearStaticSolver;
//Mode=smProgressRun;
Mode=smBackgroundRun;
Wait=1;
int maxIncrement;//the maximum increment including auto sub-increment,
long CaseNum;
double LastConvergedIncrementFactor;
maxIncrement=2000;//100(increment)X20(sub-increment)
double XYZ[3];//node coordinate, the node with loading eccentricity
double OriginalXYZ[3];
double OutOfPlumbX;
double OutOfPlumbY;
double OutOfStraightnessX;
double OutOfStraightnessY;
double OutOfStraightnessXnoSpigot;
double OutOfStraightnessYnoSpigot;
int standardNum;
long BType, SType, MType, slice;
double SData[kNumBeamSectionData];
double SJackData[kNumBeamSectionData];
double BMaterial[kNumMaterialData];
ofstream iss
("API_nonLinear_3lifts_1.0m_300mmext_1.829mx1.829m_1x1bay_random_load_ecc_
imperfections_xsection_joint_stiffness_fy.txt");
int startTime=clock();
St7CloseFile(FileUnit);
Appendix D.1 - Typical Strand7 API Subroutine for Monte Carlo Simulations 298
______________________________________________________________________

double StandardMaterialData[9];
for (int k=0;k<NumMC;k++){
St7OpenFile(FileUnit,
"API_nonLinear_3lifts_1.0m_300mmext_1.829mx1.829m_1x1bay.st7","C:\\Program
Files\\Strand7-R24-Beta\\Bin");//reopen the nominal file

//Random cross sectional area of standard and jack

St7GetBeamProperty(FileUnit, 1, &BType, &SType, &MType, SData,


BMaterial);//get the property of standard
for (int standardNum=13;standardNum<=24;standardNum++){
SData[0]=lognormStandardA();
SData[1]=lognormStandardI();
SData[2]=SData[1];
St7SetBeamSectionPropertyData(FileUnit, standardNum, &slice, SData);
}
St7GetBeamProperty(FileUnit, 4, &BType, &SType, &MType, SJackData,
BMaterial);//get the property of jack
for (int jackNum=25;jackNum<=32;jackNum++){
SJackData[0]=lognormJackA();
SJackData[1]=lognormJackI();
SJackData[2]=SJackData[1];
St7SetBeamSectionPropertyData(FileUnit, jackNum, &slice, SJackData);
}

//Random loading eccentricity

double loadEcc;
int sign;
St7GetNodeXYZ(FileUnit, 132, XYZ);//loading eccentricity, need to specify the
node coordinate
sign=(uniformSign()>=0.5)?-1:1;
loadEcc=lognormEcc();
Appendix D.1 - Typical Strand7 API Subroutine for Monte Carlo Simulations 299
______________________________________________________________________

if (loadEcc>=55.) loadEcc=55.;
XYZ[0]=3658.+sign*loadEcc;//loading eccentricity in x direction only
XYZ[1]=1829.;
XYZ[2]=3900.;
St7SetNodeXYZ(FileUnit, 132, XYZ);
St7GetNodeXYZ(FileUnit, 131, XYZ);//loading eccentricity, need to specify the
node coordinate
sign=(uniformSign()>=0.5)?-1:1;
loadEcc=lognormEcc();
if (loadEcc>=55.) loadEcc=55.;
XYZ[0]=3658.+sign*loadEcc;//loading eccentricity in x direction only
XYZ[1]=3658.;
XYZ[2]=3900.;
St7SetNodeXYZ(FileUnit, 131, XYZ);
St7GetNodeXYZ(FileUnit, 2, XYZ);//loading eccentricity, need to specify the
node coordinate
sign=(uniformSign()>=0.5)?-1:1;
loadEcc=lognormEcc();
if (loadEcc>=55.) loadEcc=55.;
XYZ[0]=1829.+sign*loadEcc;//loading eccentricity in x direction only
XYZ[1]=1829.;
XYZ[2]=3900.;
St7SetNodeXYZ(FileUnit, 2, XYZ);
St7GetNodeXYZ(FileUnit, 1, XYZ);//loading eccentricity, need to specify the
node coordinate
sign=(uniformSign()>=0.5)?-1:1;
loadEcc=lognormEcc();
if (loadEcc>=55.) loadEcc=55.;
XYZ[0]=1829.+sign*loadEcc;//loading eccentricity in x direction only
XYZ[1]=3658.;
XYZ[2]=3900.;
St7SetNodeXYZ(FileUnit, 1, XYZ);
Appendix D.1 - Typical Strand7 API Subroutine for Monte Carlo Simulations 300
______________________________________________________________________

//Random initial geometric imperfections

OutOfPlumbX=normOutOfPlumbX();//assume the out of plumb ratio is constant


for the whole structure
OutOfPlumbY=normOutOfPlumbY();//assume the out of plumb ratio is constant
for the whole structure
for (int iNode=1;iNode<=FNodeTotal;iNode++)
{
double XYZ[3];
int sign;
double XYZnoSpigot[3];
St7GetNodeXYZ(FileUnit, iNode, XYZnoSpigot);
if (XYZnoSpigot[2]>909.999&&XYZnoSpigot[2]<910.001){
OutOfStraightnessX=lognormOutOfStraightnessXnoSpigot();
sign=(uniformSign()>=0.5)?-1:1;
XYZnoSpigot[0]+=sign*1000.*OutOfStraightnessX;//lift height = 1.0m
St7SetNodeXYZ(FileUnit, iNode, XYZnoSpigot);
}
St7GetNodeXYZ(FileUnit, iNode, XYZ);
if((XYZ[2]>1909.999&&XYZ[2]<1910.001)||(XYZ[2]>2909.999&&XYZ[2]<2
910.001)){//out of straightness (the midpoint of one lift)
OutOfStraightnessX=lognormOutOfStraightnessX();
sign=(uniformSign()>=0.5)?-1:1;
XYZ[0]+=sign*1000.*OutOfStraightnessX;//lift height = 1.0m
St7SetNodeXYZ(FileUnit, iNode, XYZ);
double Spigotupper[3];
double Spigotlower[3];
double Uppernode[3];
double Lowernode[3];
St7GetNodeXYZ(FileUnit, iNode,XYZ);
St7GetNodeXYZ(FileUnit, iNode-1.,Spigotupper);
St7GetNodeXYZ(FileUnit, iNode+1.,Spigotlower);
St7GetNodeXYZ(FileUnit, iNode-2.,Uppernode);
Appendix D.1 - Typical Strand7 API Subroutine for Monte Carlo Simulations 301
______________________________________________________________________

St7GetNodeXYZ(FileUnit, iNode+2.,Lowernode);
Spigotupper[0]=(XYZ[0]-Uppernode[0])*(350./500.)+Uppernode[0];
St7SetNodeXYZ(FileUnit, iNode-1., Spigotupper);
Spigotlower[0]=(XYZ[0]-Lowernode[0])*(350./500.)+Lowernode[0];
St7SetNodeXYZ(FileUnit, iNode+1., Spigotlower);
}
}
for (int aNode=1;aNode<=FNodeTotal;aNode++)
{
double XYZ[3];
St7GetNodeXYZ(FileUnit, aNode,XYZ);
if (XYZ[2]>0){//storey out of plumb
sign=(uniformSign()>=0.5)?-1:1;
XYZ[0]+=XYZ[2]*OutOfPlumbX*sign;
//XYZ[1]+=XYZ[2]*OutOfPlumbY*sign;
St7SetNodeXYZ(FileUnit, aNode, XYZ);
}
}

//Random joint stiffness

for (int iBM=1;iBM<=FBeamTotal;iBM++){//random joint stiffness


//search all beam elements to determine if it is a connection
St7GetElementProperty(FileUnit, tyBEAM, iBM, &PropNum);
if (PropNum==9){//4way connection
double NewTableDoubles[14];
long MaxRows, NumRows;
MaxRows=50;
double ConnectionDoubles[6];
double TableDoubles[100];
St7GetTableTypeData(FileUnit, ttMomentRotation,
TableIDMomentRotation4Way, MaxRows, &NumRows, TableDoubles);//get
the data for 4way connection
Appendix D.1 - Typical Strand7 API Subroutine for Monte Carlo Simulations 302
______________________________________________________________________

char NewTableName[50];
sprintf(NewTableName, "New Random Moment Rotation ZA %d", iBM);
for (int j=0;j<NumRows*2;j++)
NewTableDoubles[j]=TableDoubles[j];
//cout<<TableDoubles[j];
NewTableDoubles[0]=-1.6E-1;
NewTableDoubles[2]=-3.6E-2;
NewTableDoubles[4]=-1.4E-2;
NewTableDoubles[6]=0.;
NewTableDoubles[7]=0.;
NewTableDoubles[12]=1.6E-1;
NewTableDoubles[10]=3.6E-2;
NewTableDoubles[8]=1.4E-2;
NewTableDoubles[9]=0.014*abs(norm4wayK1());
NewTableDoubles[11]=NewTableDoubles[9]+0.022*abs(norm4wayK2());
NewTableDoubles[13]=NewTableDoubles[11]+0.124*abs(norm4wayK3());
NewTableDoubles[5]=-NewTableDoubles[9];
NewTableDoubles[3]=-NewTableDoubles[11];
NewTableDoubles[1]=-NewTableDoubles[13];
NewTableID++;
St7NewTableType(FileUnit, ttMomentRotation, NewTableID, 7,
NewTableName, NewTableDoubles);//7: 7 pair of data
FBeamPropertyTotal++;
char NewConnectionName[50];
sprintf(NewConnectionName, "New 4Way Connection %d", iBM);
St7NewBeamProperty(FileUnit, FBeamPropertyTotal, kBeamTypeConnection,
NewConnectionName);//add a new beam (connection) property
char New4WayName[50];
sprintf(New4WayName, "New 4Way Joint %d", iBM);
St7NewBeamProperty(FileUnit, FBeamPropertyTotal, kBeamTypeConnection,
New4WayName);//add a new beam (connection) property
St7SetBeamNonlinearType(FileUnit, FBeamPropertyTotal,
ntNonlinElastic);//set joint as nonlinear elastic
Appendix D.1 - Typical Strand7 API Subroutine for Monte Carlo Simulations 303
______________________________________________________________________

St7GetConnectionData(FileUnit, 9, ConnectionDoubles);
St7SetConnectionData(FileUnit, FBeamPropertyTotal, ConnectionDoubles);
St7SetPropertyTable(FileUnit, ptConnectionBend2, FBeamPropertyTotal,
NewTableID);
St7SetPropertyTable(FileUnit, ptConnectionBend1, FBeamPropertyTotal,
TableIDMomentRotation4WayHorizontal);
St7SetElementProperty(FileUnit, tyBEAM, iBM, FBeamPropertyTotal);
}
if (PropNum==10){//3way connection
double NewTableDoubles[14];
long MaxRows, NumRows;
MaxRows=50;
double ConnectionDoubles[6];
double TableDoubles[100];
St7GetTableTypeData(FileUnit, ttMomentRotation,
TableIDMomentRotation3Way, MaxRows, &NumRows, TableDoubles);//get
the data for 3way connection
char NewTableName[50];
sprintf(NewTableName, "New Random Moment Rotation ZB %d", iBM);
for (int j=0;j<NumRows*2;j++)
NewTableDoubles[j]=TableDoubles[j];
//cout<<TableDoubles[j];
NewTableDoubles[0]=-1.6E-1;
NewTableDoubles[2]=-3.6E-2;
NewTableDoubles[4]=-1.2E-2;
NewTableDoubles[6]=0.;
NewTableDoubles[7]=0.;
NewTableDoubles[12]=1.6E-1;
NewTableDoubles[10]=3.6E-2;
NewTableDoubles[8]=1.2E-2;
NewTableDoubles[9]=0.012*abs(norm3wayK1());
NewTableDoubles[11]=NewTableDoubles[9]+0.024*abs(norm3wayK2());
NewTableDoubles[13]=NewTableDoubles[11]+0.124*abs(norm3wayK3());
Appendix D.1 - Typical Strand7 API Subroutine for Monte Carlo Simulations 304
______________________________________________________________________

NewTableDoubles[5]=-NewTableDoubles[9];
NewTableDoubles[3]=-NewTableDoubles[11];
NewTableDoubles[1]=-NewTableDoubles[13];
NewTableID++;
St7NewTableType(FileUnit, ttMomentRotation, NewTableID, 7,
NewTableName, NewTableDoubles);//7: 7 pair of data
FBeamPropertyTotal++;
char NewConnectionName[50];
sprintf(NewConnectionName, "New 3Way Connection %d", iBM);
St7NewBeamProperty(FileUnit, FBeamPropertyTotal, kBeamTypeConnection,
NewConnectionName);//add a new beam (connection) property
char New3WayName[50];
sprintf(New3WayName, "New 3Way Joint %d", iBM);
St7NewBeamProperty(FileUnit, FBeamPropertyTotal, kBeamTypeConnection,
New3WayName);//add a new beam (connection) property
St7SetBeamNonlinearType(FileUnit, FBeamPropertyTotal,
ntNonlinElastic);//set joint as nonlinear elastic
St7GetConnectionData(FileUnit, 10, ConnectionDoubles);
St7SetConnectionData(FileUnit, FBeamPropertyTotal, ConnectionDoubles);
St7SetPropertyTable(FileUnit, ptConnectionBend2, FBeamPropertyTotal,
NewTableID);
St7SetPropertyTable(FileUnit, ptConnectionBend1, FBeamPropertyTotal,
TableIDMomentRotation3WayHorizontal);
St7SetElementProperty(FileUnit, tyBEAM, iBM, FBeamPropertyTotal);
}
if (PropNum==11){//2way connection
double NewTableDoubles[14];
long MaxRows, NumRows;
MaxRows=50;
double ConnectionDoubles[6];
double TableDoubles[100];
Appendix D.1 - Typical Strand7 API Subroutine for Monte Carlo Simulations 305
______________________________________________________________________

St7GetTableTypeData(FileUnit, ttMomentRotation,
TableIDMomentRotation2Way, MaxRows, &NumRows, TableDoubles);//get
the data for 2way connection
char NewTableName[50];
sprintf(NewTableName, "New Random Moment Rotation ZD %d", iBM);
for (int j=0;j<NumRows*2;j++)
NewTableDoubles[j]=TableDoubles[j];
//cout<<TableDoubles[j];
NewTableDoubles[0]=-1.6E-1;
NewTableDoubles[2]=-3.6E-2;
NewTableDoubles[4]=-7E-3;
NewTableDoubles[6]=0.;
NewTableDoubles[7]=0.;
NewTableDoubles[12]=1.6E-1;
NewTableDoubles[10]=3.6E-2;
NewTableDoubles[8]=7E-3;
NewTableDoubles[9]=0.007*abs(norm2wayK1());
NewTableDoubles[11]=NewTableDoubles[9]+0.029*abs(norm2wayK2());
NewTableDoubles[13]=NewTableDoubles[11]+0.124*abs(norm2wayK3());
NewTableDoubles[5]=-NewTableDoubles[9];
NewTableDoubles[3]=-NewTableDoubles[11];
NewTableDoubles[1]=-NewTableDoubles[13];
NewTableID++;
St7NewTableType(FileUnit, ttMomentRotation, NewTableID, 7,
NewTableName, NewTableDoubles);//7: 7 pair of data
FBeamPropertyTotal++;
char NewConnectionName[50];
sprintf(NewConnectionName, "New 2Way Connection %d", iBM);
St7NewBeamProperty(FileUnit, FBeamPropertyTotal, kBeamTypeConnection,
NewConnectionName);//add a new beam (connection) property
char New2WayName[50];
sprintf(New2WayName, "New 2Way Joint %d", iBM);
Appendix D.1 - Typical Strand7 API Subroutine for Monte Carlo Simulations 306
______________________________________________________________________

St7NewBeamProperty(FileUnit, FBeamPropertyTotal, kBeamTypeConnection,


New2WayName);//add a new beam (connection) property
St7SetBeamNonlinearType(FileUnit, FBeamPropertyTotal,
ntNonlinElastic);//set joint as nonlinear elastic
St7GetConnectionData(FileUnit, 11, ConnectionDoubles);
St7SetConnectionData(FileUnit, FBeamPropertyTotal, ConnectionDoubles);
St7SetPropertyTable(FileUnit, ptConnectionBend2, FBeamPropertyTotal,
NewTableID);
St7SetPropertyTable(FileUnit, ptConnectionBend1, FBeamPropertyTotal,
TableIDMomentRotation2WayHorizontal);
St7SetElementProperty(FileUnit, tyBEAM, iBM, FBeamPropertyTotal);
}
}

//Random Fy for standard

for (int ii=13;ii<=24;ii++){


//cout<<"ii "<<ii<<endl;
double NewStandardFyTableDoubles[24];
long MaxRowsStandardFy, NumRowsStandardFy;
MaxRowsStandardFy=50;
double StandardFyTableDoubles[100];
char NewStandardFyTableName[100];
sprintf(NewStandardFyTableName, "Random Stress vs Strain for Standard %d",
ii);
St7GetTableTypeData(FileUnit, ttStressStrain, TableIDStressStrainStandard,
MaxRowsStandardFy, &NumRowsStandardFy, StandardFyTableDoubles);
for (int j=0;j<NumRowsStandardFy*2;j++)
NewStandardFyTableDoubles[j]=StandardFyTableDoubles[j];
NewStandardFyTableDoubles[0]=0.;
NewStandardFyTableDoubles[1]=0.;
NewStandardFyTableDoubles[3]=abs(StandardFy());
Appendix D.1 - Typical Strand7 API Subroutine for Monte Carlo Simulations 307
______________________________________________________________________

NewStandardFyTableDoubles[2]=(NewStandardFyTableDoubles[3]/2E+11)+(0.
002*pow(NewStandardFyTableDoubles[3]/(NewStandardFyTableDoubles[3]+0.8E+8),
38));
NewStandardFyTableDoubles[5]=NewStandardFyTableDoubles[3]+0.05E+8;
NewStandardFyTableDoubles[4]=(NewStandardFyTableDoubles[5]/2E+11)+(0.
002*pow(NewStandardFyTableDoubles[5]/(NewStandardFyTableDoubles[3]+0.8E+8),
38));
NewStandardFyTableDoubles[7]=NewStandardFyTableDoubles[3]+0.1E+8;
NewStandardFyTableDoubles[6]=(NewStandardFyTableDoubles[7]/2E+11)+(0.
002*pow(NewStandardFyTableDoubles[7]/(NewStandardFyTableDoubles[3]+0.8E+8),
38));
NewStandardFyTableDoubles[9]=NewStandardFyTableDoubles[3]+0.15E+8;
NewStandardFyTableDoubles[8]=(NewStandardFyTableDoubles[9]/2E+11)+(0.
002*pow(NewStandardFyTableDoubles[9]/(NewStandardFyTableDoubles[3]+0.8E+8),
38));
NewStandardFyTableDoubles[11]=NewStandardFyTableDoubles[3]+0.2E+8;
NewStandardFyTableDoubles[10]=(NewStandardFyTableDoubles[11]/2E+11)+
(0.002*pow(NewStandardFyTableDoubles[11]/(NewStandardFyTableDoubles[3]+0.8E
+8),38));
NewStandardFyTableDoubles[13]=NewStandardFyTableDoubles[3]+0.25E+8;
NewStandardFyTableDoubles[12]=(NewStandardFyTableDoubles[13]/2E+11)+
(0.002*pow(NewStandardFyTableDoubles[13]/(NewStandardFyTableDoubles[3]+0.8E
+8),38));
NewStandardFyTableDoubles[15]=NewStandardFyTableDoubles[3]+0.3E+8;
NewStandardFyTableDoubles[14]=(NewStandardFyTableDoubles[15]/2E+11)+
(0.002*pow(NewStandardFyTableDoubles[15]/(NewStandardFyTableDoubles[3]+0.8E
+8),38));
NewStandardFyTableDoubles[17]=NewStandardFyTableDoubles[3]+0.35E+8;
NewStandardFyTableDoubles[16]=(NewStandardFyTableDoubles[17]/2E+11)+
(0.002*pow(NewStandardFyTableDoubles[17]/(NewStandardFyTableDoubles[3]+0.8E
+8),38));
NewStandardFyTableDoubles[19]=NewStandardFyTableDoubles[3]+0.4E+8;
Appendix D.1 - Typical Strand7 API Subroutine for Monte Carlo Simulations 308
______________________________________________________________________

NewStandardFyTableDoubles[18]=(NewStandardFyTableDoubles[19]/2E+11)+
(0.002*pow(NewStandardFyTableDoubles[19]/(NewStandardFyTableDoubles[3]+0.8E
+8),38));
NewStandardFyTableDoubles[21]=NewStandardFyTableDoubles[3]+0.45E+8;
NewStandardFyTableDoubles[20]=(NewStandardFyTableDoubles[21]/2E+11)+
(0.002*pow(NewStandardFyTableDoubles[21]/(NewStandardFyTableDoubles[3]+0.8E
+8),38));
NewStandardFyTableDoubles[23]=NewStandardFyTableDoubles[3]+1E+8;
NewStandardFyTableDoubles[22]=(NewStandardFyTableDoubles[23]/2E+11)+
(0.002*pow(NewStandardFyTableDoubles[23]/(NewStandardFyTableDoubles[3]+0.8E
+8),38));
NewStandardFyTableID++;
St7NewTableType(FileUnit, ttStressStrain, NewStandardFyTableID, 12,
NewStandardFyTableName, NewStandardFyTableDoubles);
if (ii==1) St7SetPropertyTable(FileUnit, ptBeamAxialStressVsStrain, 1,
NewStandardFyTableID);//First standard is beam property #1
if (ii!=1) St7SetPropertyTable(FileUnit, ptBeamAxialStressVsStrain, ii,
NewStandardFyTableID);//Other standard is beam property #ii
}

//Random Fy for jack

for (int ii=25;ii<=32;ii++){


double NewJackFyTableDoubles[24];
long MaxRowsJackFy, NumRowsJackFy;
MaxRowsJackFy=50;
double JackFyTableDoubles[100];
char NewJackFyTableName[100];
sprintf(NewJackFyTableName, "Random Stress vs Strain for Jack %d", ii);
St7GetTableTypeData(FileUnit, ttStressStrain, TableIDStressStrainJack,
MaxRowsJackFy, &NumRowsJackFy, JackFyTableDoubles);
for (int j=0;j<NumRowsJackFy*2;j++)
Appendix D.1 - Typical Strand7 API Subroutine for Monte Carlo Simulations 309
______________________________________________________________________

NewJackFyTableDoubles[j]=JackFyTableDoubles[j];
NewJackFyTableDoubles[0]=0.;
NewJackFyTableDoubles[1]=0.;
NewJackFyTableDoubles[3]=abs(JackFy());
NewJackFyTableDoubles[2]=(NewJackFyTableDoubles[3]/2E+11)+(0.002*po
w(NewJackFyTableDoubles[3]/(NewJackFyTableDoubles[3]+1.1E+8),16));
NewJackFyTableDoubles[5]=NewJackFyTableDoubles[3]+0.05E+8;
NewJackFyTableDoubles[4]=(NewJackFyTableDoubles[5]/2E+11)+(0.002*po
w(NewJackFyTableDoubles[5]/(NewJackFyTableDoubles[3]+1.1E+8),16));
NewJackFyTableDoubles[7]=NewJackFyTableDoubles[3]+0.1E+8;
NewJackFyTableDoubles[6]=(NewJackFyTableDoubles[7]/2E+11)+(0.002*po
w(NewJackFyTableDoubles[7]/(NewJackFyTableDoubles[3]+1.1E+8),16));
NewJackFyTableDoubles[9]=NewJackFyTableDoubles[3]+0.15E+8;
NewJackFyTableDoubles[8]=(NewJackFyTableDoubles[9]/2E+11)+(0.002*po
w(NewJackFyTableDoubles[9]/(NewJackFyTableDoubles[3]+1.1E+8),16));
NewJackFyTableDoubles[11]=NewJackFyTableDoubles[3]+0.2E+8;
NewJackFyTableDoubles[10]=(NewJackFyTableDoubles[11]/2E+11)+(0.002*p
ow(NewJackFyTableDoubles[11]/(NewJackFyTableDoubles[3]+1.1E+8),16));
NewJackFyTableDoubles[13]=NewJackFyTableDoubles[3]+0.25E+8;
NewJackFyTableDoubles[12]=(NewJackFyTableDoubles[13]/2E+11)+(0.002*p
ow(NewJackFyTableDoubles[13]/(NewJackFyTableDoubles[3]+1.1E+8),16));
NewJackFyTableDoubles[15]=NewJackFyTableDoubles[3]+0.3E+8;
NewJackFyTableDoubles[14]=(NewJackFyTableDoubles[15]/2E+11)+(0.002*p
ow(NewJackFyTableDoubles[15]/(NewJackFyTableDoubles[3]+1.1E+8),16));
NewJackFyTableDoubles[17]=NewJackFyTableDoubles[3]+0.35E+8;
NewJackFyTableDoubles[16]=(NewJackFyTableDoubles[17]/2E+11)+(0.002*p
ow(NewJackFyTableDoubles[17]/(NewJackFyTableDoubles[3]+1.1E+8),16));
NewJackFyTableDoubles[19]=NewJackFyTableDoubles[3]+0.4E+8;
NewJackFyTableDoubles[18]=(NewJackFyTableDoubles[19]/2E+11)+(0.002*p
ow(NewJackFyTableDoubles[19]/(NewJackFyTableDoubles[3]+1.1E+8),16));
NewJackFyTableDoubles[21]=NewJackFyTableDoubles[3]+0.45E+8;
NewJackFyTableDoubles[20]=(NewJackFyTableDoubles[21]/2E+11)+(0.002*p
ow(NewJackFyTableDoubles[21]/(NewJackFyTableDoubles[3]+1.1E+8),16));
Appendix D.1 - Typical Strand7 API Subroutine for Monte Carlo Simulations 310
______________________________________________________________________

NewJackFyTableDoubles[23]=NewJackFyTableDoubles[3]+1.2E+8;
NewJackFyTableDoubles[22]=(NewJackFyTableDoubles[23]/2E+11)+(0.002*p
ow(NewJackFyTableDoubles[23]/(NewJackFyTableDoubles[3]+1.1E+8),16));
NewJackFyTableID++;
St7NewTableType(FileUnit, ttStressStrain, NewJackFyTableID, 12,
NewJackFyTableName, NewJackFyTableDoubles);
if (ii==4) St7SetPropertyTable(FileUnit, ptBeamAxialStressVsStrain, 4,
NewJackFyTableID);//Jack is beam property #4
if (ii!=4) St7SetPropertyTable(FileUnit, ptBeamAxialStressVsStrain, ii,
NewJackFyTableID);//Jack is beam property #ii
}

//Run advanced analysis

St7RunSolver(FileUnit, Solver, Mode, Wait);


long NumPrimary, NumSecondary;
St7OpenResultFile(FileUnit,"API_nonLinear_3lifts_1.0m_300mmext_1.829mx
1.829m_1x1bay.nla","",FALSE, &NumPrimary, &NumSecondary);
bool Converged;
bool StrengthRecorded;
StrengthRecorded=false;
long LastConvergedIncrement;//find the last converged increment
LastConvergedIncrement=0;
Converged=false;
for (int j=maxIncrement;j>0;j--){
St7GetResultCaseConvergence(FileUnit, j, &Converged);
if(Converged==true){
char NewCaseName[kMaxStrLen+1];
LastConvergedIncrement=j;
St7GetResultCaseName(FileUnit,LastConvergedIncrement,NewCaseName,
kMaxStrLen);
cout<<"*** MC "<<k<<" is "<<NewCaseName<<endl;
StrengthRecorded=true;
Appendix D.1 - Typical Strand7 API Subroutine for Monte Carlo Simulations 311
______________________________________________________________________

iss<<"MC "<<k<<" "<<NewCaseName<<endl;


break;
}
}
if (StrengthRecorded!=true){//if the ultimate strength can not be recovered by
above method
Converged=false;
for (int j=1;j<=maxIncrement;j++){
St7GetResultCaseConvergence(FileUnit, j, &Converged);
if(Converged!=true){
char CaseName[kMaxStrLen+1];
LastConvergedIncrement=j-1;
St7GetResultCaseName(FileUnit,LastConvergedIncrement,CaseName,
kMaxStrLen);
cout<<"MC "<<k<<" is "<<CaseName<<endl;
iss<<"MC "<<k<<" "<<CaseName<<endl;
break;
}
}
}
St7CloseResultFile(FileUnit);
St7CloseFile(FileUnit);
}
int stopTime=clock();
double timeUsed= double(stopTime - startTime)/CLOCKS_PER_SEC ;
cout<<endl<<"Time used "<<timeUsed<<endl;
system("pause");
FreeSt7API();
return 0;
}

Вам также может понравиться