You are on page 1of 229

University of Wollongong

Research Online
University of Wollongong Thesis Collection University of Wollongong Thesis Collections
1993
Dynamic investigation of semi-rigid tubular T-
joints
Iraj Hoshyari
University of Wollongong
Research Online is the open access institutional repository for the
University of Wollongong. For further information contact Manager
Repository Services: morgan@uow.edu.au.
Recommended Citation
Hoshyari, Iraj, Dynamic investigation of semi-rigid tubular T-joints, Doctor of Philosophy thesis, Department of Civil and Mining
Engineering, University of Wollongong, 1993. http://ro.uow.edu.au/theses/1240
UNIVERSITY OF WOLLONGONG
DYNAMIC INVESTIGATION OF SEMI-RIGID
TUBULAR T-JOINTS
A thesis submitted in fulfilment of the
requirement for the award of the degree
Doctor of Philosophy
by
Iraj Hoshyari
Department of Civil and Mining Engineering
August 1993
To my family
Declaration
This is to certify that the work presented in this thesis was carried out by the author in the
Department of Civil and Mining Engineering, University of Wollongong, and has not been
submitted to any other university or institute for a degree except where specifically
indicated.
Iraj Hoshyari
i
Abstract
Offshore structures are fabricated from tubes or circular hollow sections. Tubular joints
which are created at the intersection of the circular members experience complicated
structural problems such as stress concentration, ultimate strength, joint stiffness and fatigue
life. S o m e aspects of tubular T-joints regarding joint stiffness and stress concentration are
investigated in this thesis. The methods employed in these investigations are based on
dynamic theoretical and experimental modelling techniques.
Semi-rigidity of tubular joints leads to different results when compared with the c o m m o n
assumption of rigid joint analysis. However, due to the lack of an efficient modelling
method, tubular joints are still assumed to be rigid in most analyses. A method based on the
model of a beam with end springs is used in this study to derive the bending and axial
stiffness of joints. This method can be distinguished from other techniques, as it employs the
natural frequencies rather than static measurements.
A n extensive Finite Element analysis is carried out to establish a series of stiffness equations
for inplane bending, out of plane bending and axial deformation of brace in tubular T-joints.
The proposed parametric equations include diameter to thickness ratio which is not
presently considered in any other formulae. The analysis results show that including
diameter to thickness ratio makes a significant difference in the stiffness of tubular T-joints.
The validity of the Finite Element analyses is investigated by experimentally testing eleven
steel T-joints.
The effect of dynamic loading on strain or stress concentration factors in tubular joints are
also investigated in this thesis. Stress concentration factors are used to calculate the
maximum stress values at the tubular joints. The results show that strain or stress
concentration factor is frequency dependent, when strains at tubular joints are subjected to
dynamic forces. This parameter has not been considered in the other studies of stress
concentration of tubular joints.
Another aspect of tubular joints investigated herein is the effect of joint stiffness
consideration on the fatigue life estimation of offshore towers. Results of the analyses of a
100m tower in this study show an average difference of 3 0 % in fatigue life due to the
flexibility consideration of tubular joints. Further detailed investigation though is required
into the joint stiffness effect on fatigue life estimate of tubular structures.
ii
Acknowledgments
The generosity of the Iranian Ministry of Culture and Higher Education in providing the
author with a scholarship is gratefully acknowledged.
This work could not have been completed without the support of a number of people to
w h o m I am deeply grateful:
Dr Richard Kohoutek, m y supervisor, whose guidance was constructive in maintaining a
proper direction towards the completion of this thesis.
M r Ian Bridge, a professional technician, who fabricated the test specimens and related
hardware.
Mr. Charles Mitchell and Tino Ferrero for their assistance in carrying out the experiments.
Other members of technical staff in the Department of Civil and Mining Engineering
especially M s . Lyn Middleton for proof reading of m y thesis, and Mr. Richard W e b b for his
general assistance in the laboratory.
Mrs. J. Kohoutek for drawing the sketches of the test specimens.
I wish to thank m y wife, Mahbobeh, for her unending patience and understanding through
all the difficulties w e encountered when this work was being carried out.
Iraj Hoshyari,
August 1993
iii
Table of Contents
1 INTRODUCTION
1.1 Preamble 2
1.2 Brief history of offshore platforms 2
1.3 Terminologies used in steel template platforms and tubular joints 4
1.4 W h y joint modelling is important? 8
1.5 Platform tower global analysis 9
1.6 Scope of this thesis 11
1.7 Summary 12
2 LITERATURE SURVEY
2.1 Introduction 14
2.2 General 14
2.3 Tubular joints 15
2.3.1 Analytical methods 15
2.3.2 Experimental and semi-experimental methods 18
2.3.3 Numerical methods 21
2.4 Effects of tubular joint flexibility on tower analysis 25
2.5 Model of joint in stiffness method 37
2.5.1 Bending rigidity factor 39
2.6 Summary 40
3 THEORETICAL MODEL
3.1 Introduction 42
3.1.1 Beam theory 42
3.2 Matrix formulation of beams 43
3.2.1 Beam with semi-rigid ends (bending deformations only) 45
3.2.2 Effect of shear deformations on bending stiffness 48
3.2.3 Stiffness determination of a joint 48
3.3 Bending rigidity factor 50
3.4 Stiffness of a joint due to brace axial loading 52
3.4.1 Axial rigidity factor 54
3.5 Summary.. 55
E X P E R I M E N T A L DETERMINATION O F JOINT STIFFNESS
4.1 Introduction 57
4.2 Test specimens 57
4.2.1 Survey of non-dimensional parameters of T-joints 58
4.2.2 Material properties and fabrication of test specimens 59
4.3 Test set-up 60
4.3.1 Test procedure 62
4.3.2 Test results and discussion 65
4.4 Calculation of rigidity factor and joint stiffness 67
4.5 Effect of joint support conditions 69
4.5.1 Provisions of end supports for test specimens 70
4.6 Summary 72
DETERMINATION OF JOINT RIGD3ITY USING FINITE ELEMENT METHOD
5.1 Introduction 74
5.2 A review of the Finite Element method 74
5.2.1 Galerkin Method (a finite element formulation) 75
5.2.2 Shell Finite Element 77
5.3 Description of the Finite Element Package: A L G O R 78
5.3.1 Shell Element in A L G O R 78
5.3.2 A Bench Mark on A L G O R 79
5.3.3 Comparison of shell and three dimensional element performances 80
5.4 Mesh generation procedure used to model T-joints 81
5.5 Boundary conditions for the FE models 82
5.5.1 Effect of joint support conditions 86
5.6 Finite Element analysis of the tested T-joints 88
5.6.1 Comparison of the Finite Element analysis and test results 89
5.7 Summary 91
RIGDDITY AS FUNCTION OF JOINT GEOMETRICAL PARAMETERS
6.1 Introduction 94
6.2 Finite Element analysis results 94
6.2.1 Inplane bending mode 96
6.2.2 Out of plane bending mode 102
6.2.3 Axial deformation of brace 106
6.3 Derivation of stiffness equations 111
6.4 Comparison with other stiffness equations 114
6.4.1 D N V (1977) 115
6.4.2 UEG(1985) 115
6.4.3 Efthymiou(1985) 116
6.4.4 Fessler(1986) 116
6.4.5 Uedaetal.(1990) 117
6.4.6 Chen B.etal. (1990) 117
V
6.4.7 Kohoutek and Hoshyari (1992) 117
6.4.8 Graphical Comparison of different stiffness formulae 117
6.5 Summary 123
7 DYNAMIC STRESS CONCENTRATION FACTOR
7.1 Introduction 126
7.2 Application of stress concentration factor in fatigue analysis 128
7.3 Methods of SCF determination 131
7.3.1 Experimental methods 132
7.3.2 Interpretation of stress concentration factor from measurements 133
7.3.3 The Finite Element Method (FEM) 134
7.4 Reliability approach for stress concentration factor 135
7.5 Dynamic strain reading 136
7.5.1 Test set-up 138
7.5.2 Test results 139
7.6 Dynamic strain calculation using stiffness method 145
7.7 Dynamic strain reading and SCF/SNCF reliability 147
7.8 Summary 148
8 SEMI-RIGIDITY EFFECTS ON BEHAVIOUR OF OFFSHORE TOWERS
8.1 Introduction 150
8.2 Tower geometry and joint specifications 150
8.3 Stiffness determination of the joints of the tower 152
8.5 Loading 152
8.6 Results 155
8.6.1 Deflections 155
8.6.2 Axial forces 157
8.6.3 Bending moments 158
8.6.4 Dynamic characteristics 160
8.7 Fatigue life estimation 167
8.8 Summary 172
9 SUMMARY AND CONCLUSIONS
9.1 Summary 175
9.2 Conclusions 181
9.3 Future research work
183
vi
REFERENCES 184
APPENDICES 192
Appendix A
Stiffness matrix of beam with semi-rigid ends 193
Appendix B
Stiffness matrix of truss member with semi-rigid ends 198
Appendix C
Mode shapes 1,2, and 3 for the tested T-joint specimens 201
Appendix D
List of sub-programs in A L G O R (used in the analysis of joints) 204
Appendix E
Description of TBC3 and GCS8 finite elements 205
Appendix F
Effect of chord length 207
Appendix G
Parametric SCF Formulae for inplane bending of T-joints 209
vii
List of Figures
1.1 A concrete gravity platform in the North Sea 3
1.2 A tension leg platform 3
1.3 Bullwinkle offshore oil drilling platform 4
1.4 Principal elements of an eight-leg steel template platform 5
1.5 Simple tubular T-joint and the common non-dimensional parameters 6
1.6 Some examples of complex joints 7
1.7 Concrete grouted joint 7
1.8 Example of a joint substructure constructed from shell finite elements 8
2.1 Kellogg's models of a tubular joint 16
2.2 Model of cylindrical vessel used by Bijlaard 16
2.3 Shell element used by Holmas
2.4 Extra D O F to express local joint behaviour used by Holmas 17
2.5 Test rig used by Fessler (1981) 19
2.6 Stress distribution assumed in Punching Shear Model 20
2.7 Model of joint substructure used by Bouwkamp (1980) 22
2.8 Rotations measured by Efthymiou for calculation of joint flexibility (1985) 22
2.9 Joint model proposed by Ueda (1986) 23
2.10 Joint super-element used by Souissi (1990) 24
2.11 The results of Kawashima's analyses on the frames with rigid and semi-rigid joints.. 26
2.12 K-braced frame analysed by Ueda and its load cases 27
2.12 Frame models analysed in U R 2 2 report by U E G (1984) 28
2.13 Nodal points considered in U R 2 2 Study to represent a joint 28
2.14 Predicted buckling load by Jong (1987) 32
2.15 Tower analysed by T. Chen (1990) 33
2.16 The frame analysed by Souissi (1990) 34
2.17 The structures analysed by Recho (1990) 35
2.18 Moment-rotation relationship for a semi-rigid joint 37
2.19 Joint model developed by Kawashima (1984) 38
3.1 DOFs for two dimensional beam element 43
3.2 Stiffness coefficients of beam with continuous mass due to
the rotation of left end 44
3.3 Beam with semi-rigid ends (after Kohoutek 1991b) 45
3.4 Dynamic methods to determine rigidity of a joint 49
viii
3.5 B e a m model for definition of rigidity factor (unit length) 50
3.6 Relationship between rotational stiffness and beam end flexibility 51
3.7 Axially loaded member with semi-rigid ends 52
3.8 Relationship between axial stiffness and axial end flexibility 54
4.1 Distribution of f$ parameter for T-joints. The survey was carried out by UEG (1985)
on six offshore structures 58
4.2 Distribution of y parameter for T-joints. The survey was carried out by U E G (1985)
on six offshore structures 58
4.3 Distribution of y parameter for T-joints. The survey was carried out on the database
in U E G (1985) 59
4.4 Distribution of P parameter for T-joints. The survey was carried out on the database
in U E G (1985) 59
4.5 Distribution of T parameter for T-joints. The survey was carried out on the database
in U E G (1985) 59
4.6 A typical weld cross section of a T-joint specimen 60
4.7 Diagrammatic test set-up for frequency measurement of T- joints (IPB) 61
4.9 Typical frequency spectrum of a tubular T-joint 62
4.10 Lumped mass model used to establish the mode shapes of joints 63
4.11 First two mode shapes of T-joint T10 from Dynamic Modal Analysis 65
4.12 Relationship between natural frequencies of specimens and diameter ratio 66
4.13 Stiffness calculation of a T-joint using measured natural frequency 67
4.14 Analytical model used for rigidity calculation of tested T-joints (TPB) 68
4.15 Support arrangement for inplane bending test 70
4.16 Support arrangement for out of plane bending test 71
5.1 Thin plate and shell element in ALGOR 78
5.2 L C C T 9 element assembly 79
5.3 Geometry and meshes used for cylinder under pressure 79
5.4 Typical mesh patterns used for the analysis of T-joints 81
5.5 Data generation procedure for tubular T-joints 82
5.6 Boundary conditions applied to F E models of the T-joints 83
5.7 T-joint T2, first three mode shapes in IPB mode 88
5.8 Weld cut-back at the saddle on joints with P = 1 90
5.9 Correlation between natural frequencies obtained from F E analysis and experiment. 91
6.1 Geometry of Finite Element models used for flexibility analysis of T-joints 95
6.2 Relationship between P and JPB stiffness 98
6.3 Relationship between y and IPB stiffness 99
ix
6.4 Relationship between x and IPB stiffness 100
6.5 Relationship between P and O P B stiffness 103
6.6 Relationship between y and O P B stiffness 104
6.7 Relationship between x and O P B stiffness 105
6.8 Relationship between P and Axial stiffness 108
6.9 Relationship between y and axial stiffness 109
6.10 Relationship between x and axial stiffness 110
6.11 Predicted values against data used for curve fitting 112
6.12 Model adopted for IPB joint stiffness with the absolute errors of curve fitting 113
6.13 Model adopted for O P B joint stiffness with the absolute errors of curve fitting 113
6.14 Model adopted for joint axial stiffness with the absolute errors of curve fitting 114
6.15 Comparison of different equations for axial joint stiffness 120
6.16 Comparison of different equations for IPB joint stiffness 121
6.17 Comparison of different equations for O P B joint stiffness 123
7.1 Stress distribution at the intersection of a simple T-joint 126
7.2 Stress distribution near weld toe 128
7.3 The geometry and forces of the T-joint considered in Example 1 128
7.4 Hot-spot definition for T and DT-joints recommended by E C S C 134
7.5a Variations of S C F for direction of axial loading 136
7.5b Variations of SCF for nominally identical joints 136
7.6 A typical strain gauge location diagram of a test specimen 137
7.7 Diagrammatic test set-up for the dynamic strain measurement 138
7.8 Dynamic strain readings for joint T10 139
7.9 Relationships between strain ratio and load frequency for joints with different p
ratios (IPB) 141
7.10 Relationships between strain ratio and load frequency for joints with different p
ratios (OPB) 144
7.11 Beam model to study dynamic effects on strain ratio 146
7.12 Strain ratios resulted from beam models and test models for joints with a) P=l .0,
b) P=0.52, c) P=0.28 146
8.1 Geometry of the tower analysed 151
8.2 W a v e definition for load cases 2,3 and 4 153
8.3 W a v e forces calculated for load cases 2, 3, and 4 155
8.4 Deflected shape of CP1 and M P 2 frames for load case 1 156
8.5 Axial forces for load cases 2, 3 and 4 and their variations between flexible joint
analysis and conventional analysis (MP2-CP1) 157
X
8.6
a) Bending moment diagram for frame M P 2 (Load Case 4) 158
b) Bending moment difference between M P 2 and CP1 (MP2-CP1) 158
8.7 Relationship between stiffness and frequency for a) CP1 and b) M P 2 towers 161
8.8 Mode shape 1 of towers CP1 and M P 2 163
8.9 Mode shape 2 of towers CP1 and M P 2 163
8.10 Mode shape 3 of towers CP1 and M P 2 164
8.11 Mode shape 4 of towers CP1 and M P 2 164
8.12 Mode shape 5 of towers CP1 and M P 2 165
8.13 Mode shape 6 of towers CP1 and M P 2 165
8.14 Mode shape 7 of towers CP1 and M P 2 166
8.15 Mode shape 8 of towers CP1 and M P 2 166
8.16 Mode shape 9 of towers CP1 and M P 2 167
8.17 Typical relationship between wave period and stress per wave height 168
8.18 Stress range per wave height of different wave periods for joints a) 5 and 6 and
b) 7 and 8 169
xi
List of Tables
2.1a Kawashima's results for L-type frame 26
2.1b Kawashima's results for portal frame 26
2.2 Joint specification in Ueda's analyses 27
2.2 Joint parameters used in U R 2 2 Study 29
2.3 Summary of changes from U E G report on joint flexibility 30
2.4 Effect of flexibility consideration in the analysis, Chen T. (1990) 33
2.5 Effect of joint flexibility on internal forces by Recho (1990) 35
2.6 Fatigue life difference (NR/NF) when joint flexibility is considered 35
2.7 M a x i m u m response change for models with rigid joints obtained in various research
works when joint flexibility was considered 36
4.1 Geometrical dimensions and non-dimensional parameters of tested joints 57
4.2 Metallurgical properties of the pipe sections used for test specimens 60
4.3 Natural frequencies of the T-joints using lumped mass theory (TPB) 64
4.4 Natural frequencies of the T-joints using Dynamic Deformation Method (IPB) 64
4.5 Natural frequency of the T-joints due to cantilever mode of vibration 66
4.6 Measured natural frequencies (co), rigidity factor (v
b
), and stiffness (k) of
the tested T-joints 68
4.7 Joint stiffness calculated for hinged and fixed analytical models using test
results of specimens that have hinged supports 69
5.1 Results of different types of shell elements - cylinder under pressure 80
5.2 Comparison of shell and three dimensional solid elements 81
5.3 Application of internal boundary conditions to the F E model of T-joint T10 (IPB).. 84
5.4 Calculated results for the F E models of the T-joints with different support
conditions (IPB) 87
5.5 Results of F E analyses and experiments 89
6.1 Non-dimensional parameters used for FE models of T-joints 95
6.2 Identification names and individual joint parameters of F E models 96
6.3a Calculated natural frequencies and stiffness of T-joint F E models, IPB (t = 0.2) 97
6.3b Calculated natural frequencies and stiffness of T-joint F E models, IPB (t = 0.5) 97
6.3c Calculated natural frequencies and stiffness of T-joint F E models, IPB (T = 1.0) 97
6.4a Calculated natural frequencies and stiffness of the T-joint F E models, OPB (T = 0.2) 102
6.4b Calculated natural frequencies and stiffness of the T-joint F E models, OPB (t = 0.5) 102
6.4c Calculated natural frequencies and stiffness of the T-joint F E models, O P B ( T =1.0). 102
xii
6.5a Calculated natural frequencies and stiffness of the T-joint F E models, axial
deformation (t = 0.2) 106
6.5b Calculated natural frequencies and stiffness of the T-joint F E models, axial
deformation (t = 0.5) 107
6.5c Calculated natural frequencies and stiffness of the T-joint F E models, axial
deformation (t = 1.0) 107
6.6 T-joint stiffness formulae and the corresponding correlation factors Ill
6.7a Geometrical parameters and stiffness of the joints of Figures 6.15 to 6.17
calculated using the formulae by others 118
6.7b Geometrical parameters and stiffness of the joints of Figures 6.15 to 6.17
measured by others 118
6.7c Geometrical parameters and stiffness of the joints of Figures 6.15 to 6.17
calculated by the formulae developed in this study 119
6.8 Comparison of the results of proposed equation with Kohoutek's equation (IPB).. 122
7.1 Hot-spot Stress Concentration Factors for the joint of Example 1 129
7.2 No. of occurrences for different wave heights used in the fatigue example 130
7.3 Nominal stress ranges in brace member for different deformation modes 130
7.4 Hot spot stress ranges for the T-joint in fatigue example 130
7.2 Dynamic strain readings and SNCFs for IPB mode 140
7.3 Dynamic strain readings and SNCFs for O P B mode 143
8.1 Geometrical properties of joints used in the analysed towers 151
8.2 Natural frequency and stiffness of the T-joint in M P 2 tower 152
8.3 Details of the load cases in the analysis of towers 153
8.4 Linear (Airy) Wave Theory equations 154
8.5 Deflection changes when joint flexibility is incorporated in analysis 156
8.6 Axial forces, their changes and axial stress changes for load case 4 158
8.7 Bending moment amplitude changes between M P 2 and CP1 (MP2-CP1) 159
8.8 Natural frequencies of towers CP1 and M P 2 (Hz) 161
8.9 Individual wave occurrences 'n' in a sea state of H
s
=3m, T^= 7 sec,
normalised to 1 year 168
8.10 Fatigue damage of joint 5 per year for each wave height and wave period 170
8.11 Fatigue damage of joint 6 per year for each wave height and wave period 170
8.12 Fatigue damage of joint 7 per year for each wave height and wave period 171
8.13 Fatigue damage of joint 8 per year for each wave height and wave period 171
8.14 Fatigue life predictions and their changes when joint stiffness is considered
in analysis 172
Nomenclature
fli coefficient of displacement vector
A sectional area
As shear area
C vector of integration constants
C\, C2, C
3
, C
4
.... integration constants
d brace diameter
D chord diameter
d' distance between weld toes at the saddle of a joint
D
s
damage summation ratio
E modulus of elasticity
F. degree of fixity
g acceleration due to gravity
f
a
stress due to axial loading
fiPB stress due to inplane bending
fopB stress due to out of plane bending
h longitudinal displacement h(x,t)
I. moment of inertia
K global stiffness matrix
k local stiffness matrix of an element
ky member of a stiffness matrix
k'iQ-) stiffness of an axial spring at end /(/)
,(,) stiffness of a rotational spring at end /(/)
L chord length
L differential operator
/ brace length, length, direction cosine
M bending moment
Mam added mass
m mass per unit length, direction cosine
M
c
continuous mass matrix
M
L
lumped mass matrix
n direction cosine, number of cycles
Af number of cycles before fatigue failure, shape function
N
t
interpolation functions
N
F
the number of cycles to failure when connection is considered flexible
NR the number of cycles to failure when connection is considered rigid
xiv
P axial force
P force vector
Q matrix that converts end forces P to integration constants C
R discrepancy ratio
p water density
t brace wall thickness, time
T. chord wall thickness
U water particle velocity
U water particle velocity
u longitudinal displacement u(x), general displacement function
u approximated displacement function
Ub displacement function at the boundaries
V shear force
v transverse displacement v(x,f)
X body force per unit volume in x direction
x displacement vector
X surface force per unit area in x direction
Y. body force per unit volume in y direction
y transverse displacement y(x)
y
s
displacement due to the shear deformation
Y surface force per unit area in y direction
Z body force per unit volume in z direction
Z
1
surface force per unit area in z direction
a non dimensional parameter 2L/D
ay) bending flexibility of beam end i(j) (a, = El/kil)
P non dimensional parameter dID
P' non dimensional parameter d'lD
P,(,) axial flexibility of beam end i(j) (P, = El/k'f)
A displacement vector
8 transverse displacement
<I> matrix that converts end displacements A to integration constants C
T rigidity factor, partial factor to give the appropriate level of confidence
y non dimensional parameter DI2T
r) longitudinal displacement
v
a
axial rigidity factor
Vb bending rigidity factor
8 rotation, angle of diagonal brace in a Y-joint
p mass per unit volume
XV
Ox,
y
,z normal stress
x \JT
ixy,yz,zx shear stress
u) frequency
CA ith natural frequency
A / stress range
List of Abbreviations
AD Analog to Digital
DDM Dynamic Deformation Method
Det Determinant
DOF Degree Of Freedom
DSCF Design Stress Concentration Factor
ECSC European Community of Steel and Coal
FE Finite Element
FEM Finite Element Method
FS Factor of Safety
HS Hot Spot
IPB InPlane Bending
NF Natural Frequency
OPB Out of Plane Bending
PC Personal Computer
S-N Stress-Number of cycle curve for fatigue estimation
SCF Stress Concentration Factor
SNCF Strain Concentration Factor
U K DEn United Kingdom Department of Energy
UKOSRP United Kingdom Offshore Steels Research Program
Chapter 1
INTRODUCTION
Chapter 1: Introduction
1.1 Preamble
The first oil drilling platforms in sea were built of wood, with much smaller dimensions than
their descendants which are almost the biggest structures on the earth. U p to the 1990s,
three major categories of fixed offshore platforms were developed: concrete gravity, tension
leg and steel template platforms. Approximately 98 percent of the existing platforms are of
the template type with heights varying from a few meters to more than 400 meters.
Concrete gravity platforms gain their resistance through their very large weight; they have a
heavy base, holding them in place against lateral loads. This type of platform is suitable for
regions with soft soil seabeds. Figure 1.1 shows a concrete gravity platform installed in the
North Sea. A tension leg platform is also shown in Figure 1.2. Tension leg platforms, which
are placed in the compliant category of offshore platforms have a flexible response to the
lateral loads. They adjust themselves horizontally with the water motion, and their
movements are controlled by mooring systems, which typically consist of chains, cables,
ropes and anchors (Demirbilek, 1989).
Template platforms are explained in Section 1.3.
1.2 Brief history of offshore platforms
In the 1890s, man first exploited oil from an offshore field on the coast of California, U S A .
Access to the oil well was through piers extending from shore. By 1900 eleven piers were
constructed in that area and drilling was being carried out in water 150m from the shoreline.
Offshore drilling continued through wooden platforms within sight of land in shallow
waters. In the 1920s timber platforms were erected in an oilfield off Venezuela.
After the Second World W a r with the development of technology, deeper water was
searched for oil and taller platforms were fabricated. The first steel platforms were
constructed in the Gulf of Mexico during 1947 and there were two platforms constructed in
that year which became the design standard for many years (Graff, 1981). One was located
29km offshore in 6 m of water. The total platform size was 53m by 33m. It had 268 steel
piles driven through the jacket legs. The other steel platform had a smaller deck area of
250m
2
and was placed in 5.5m of water offshore. Structural members of the tower had pipe
sections which showed less resistance to the forces generated by water. Pile type
foundations were used for transferring the loads to the supports. The pipes were driven
Chapter 1: Introduction 3
Figure 1.1. A concrete gravity platform in the North Sea Figure 1.2. A tension leg platform
through the main legs to a sufficient depth in the seabed and connected to the deck at the
top. A s a result, fixed steel offshore platforms are also called template platforms.
With the rapid growth of industry and its vital need to the oil products, fabrication of fixed
platforms continued and deeper waters were targeted. In September 1988, Shell Offshore
Incorporation installed Bullwinkle in the Green Canyon area, Gulf of Mexico. This platform,
which is 490 meter high, has the tallest tower in the world. Its design, fabrication and
installation took five years. Seventy thousand tonnes of steel were used to construct the
Bullwinkle platform (Sterling, 1989). Figure 1.3 shows the Bullwinkle oil drilling platform
tower.
Chapter 1: Introduction
4
<J>
(sXD

<a>
,<p? ?
I 140''
2 I WL45- 9"-J>

LAN RT +15' \)
PLRN RT -370'
30'-
s rr
Q> T I
4 19' 3 "
140' 140' 140'
1 1
-30'
/-SZ\
\ /
ROW "RJ.B
PLRN RT -1350'
Figure 1.3. Bullwinkle offshore oil drilling platform (after Digre, 1989)
1.3 Terminologies used in steel template platforms and tubular joints
A steel template platform consists of three main sections: 1) superstructure, 2) jacket, and
3) piles. The different components of an eight leg template platform is shown in Figure 1.4.
The jacket is a tubular steel structure sitting on the seabed and supporting the deck. It
surrounds the piles and holds them in position from the mudline to the deck; supports
conductors, pumps, sumps, risers, etc. and hence is called a jacket. The superstructure or
deck which is placed on top of the jacket provides required operational space. The deck sits
on girders and trusses, and eventually on the piles which extend from beneath the
superstructure into the seabed through the jacket legs.
The structural members that secure the jacket legs horizontally are called braces or
branches. Where one or more braces meet a leg, a joint is created. Figure 1.5 shows a
typical simple T-joint with commonly used notations and non-dimensional parameters.
Chapter 1: Introduction
5
skid beams
barge bumper
horizontal brace
vertical diagonal
jacket leg can
bracing stub
upper or drilling deck
wind girders
longittudinal truss
skirt pilre sleeve
launch truss
Figure 1.4. Principal elements of an eight-leg steel template platform (after Graff, 1981)
Chapter 1: Introduction
Brace or
Branch
l r
Saddle
-H
D
M -
P = 75
=
_D
Y
27
/
a =
21
D
Figure 1.5. Simple tubular T-joint and the common non-dimensional parameters
Tubular joints are classified into four categories as follows (UEG, 1985):
1) simple welded joints,
2) complex welded joints,
3) cast steel joints, and
4) composite joints.
This classification is not strict, but is widely accepted. Simple joints are those without any
stiffener, gusset plate, diaphragm or grout. In multi braced simple joints the braces do not
overlap. To strengthen a simple joint, the chord section is usually thickened in the connec-
tion zone. This section with higher strength is called joint can (see Figure 1.4).
The term complex is assigned to the following joints:
1) joints with uniplanar or multiplanar overlapping brace members,
2) joints with internal stiffeners or diaphragms, and
3) joints with external stiffeners.
Chapter 1: Introduction 7
Complex joints have higher strength and stiffness than simple joints. Some examples of
complex joints are shown in Figure 1.6.
ViewA-A
Figure 1.6. Some examples of complex joints (UEG, 1985)
Cast joints are made by a casting process in which the brace and chord are cast together.
Therefore, there is no welded connection between the brace and chord, creating better
shapes of fillet. Cast joints are potentially stronger than the welded joints (Edwards and
Fessler, 1985).
Composite joints are those filled fully, or partially between the leg and the pile passing
through the leg, with concrete. A double-skin grout reinforced joint which is placed in the
latter group of grouted joints is shown in Figure 1.7.
Chapter 1: Introduction
8
Pile centre m a y also
be grout filled
Figure 1.7. Concrete grouted joint
1.4 W h y joint modelling is important?
A s explained in Section 1.2 the most suitable structural sections for jacket structures are
circular tubes. The best way to connect the tubes is direct welding creating tubular joints,
one of the most critical subjects in the analysis and design of an offshore platform. The
diameter to thickness ratios of the tubes at a joint zone (DIT) are generally between 25 and
60, which pronounces the shell characteristics in the structural behaviour of the joint.
To determine the stress resultants at a joint, nominal stress values are calculated through a
structural analysis, where the structural model of a jacket platform is usually a space frame,
or in simple cases a two-dimensional model. In this regard, it is c o m m o n in practice that
the connection of tubular members be considered as rigid. This means that the angle
between members does not change after the structure is gone under loading. However, a
better estimation to the internal forces in a jacket structure can be obtained by
incorporating the flexibility of joints in the analysis. The significance of stress distribution
at tubular joints is in connection with fatigue life. Because the internal forces are generated
under wind, waves, currents and vortices, stress fluctuation and ultimately fatigue failure
m a y be developed in the structural members of offshore platforms. B y considering the joint
stiffness, a designer can calculate the stress resultants at the joints more realistically and
consequently make a better estimate of the fatigue life. In addition to stress evaluation at
the joints, the deflected shape of a structure is calculated more accurately by employing the
flexibility of the joints. This is especially important for tall platforms. More accurate
estimation of the dynamic behaviour of tower frames is another feature of the joint
flexibility consideration. Furthermore, applying the joint flexibility provides more realistic
data for buckling design of compression members.
Chapter 1: Introduction 9
1.5 Platform tower global analysis
Presently, structural analyses are performed on computers and stiffness method based
programs are used. Regarding joint modelling in computer structural analysis, three basic
methods have been identified (Tebbett, 1982). They are:
1) addition of simple springs in line with brace,
2) entering of an effective length for the brace, and
3) use of a substructure of three dimensional finite elements for the tubular joints.
Tebbett labels the first alternative as the easiest and the third one as the most accurate.
Using the second alternative requires more attention to the length of the members, not to
invalidate wave loading and member stability checks. Figure 1.8 shows an example of the
third alternative, which uses a F E substructuring scheme.
Figure 1.8. Example of a joint substructure constructed from shell finite elements
The same methods as Tebbett identified are still used in studies where in essence most use
static Finite Element analysis as base. There has been a number of studies on the stiffness
or flexibility of tubular joints since the late 1970's. The c o m m o n limitation of these studies
is the lack of experimental backup supporting the results, which is indeed due to the
difficulties involved with the measurements of static deformations. In this regard, Fessler et
al. (1981) has carried out the only conclusive experimental study into the stiffness of
tubular joints and is described in Chapter 2.
Tebbett and Lalani (1986) has considered the local joint flexibility as an important
parameter in structural analysis, especially in the reassessment of offshore structures.
However, they have suggested that a consistent method should be used for flexibility
modelling of all joints in a structure.
Chapter 1: Introduction
10
Lack of efficient methods of structural modelling has caused the consideration of rigid
attachment between the beam members in tower global analyses (Barltrop, 1991). This
assumption introduces two important features in regard to the structural analysis
procedures such as the stiffness method. One feature is the simplicity of force-displacement
relationship used for the beam members when the joints are assumed rigid. Using the
c o m m o n matrix equation of kx = f for uniform beams, which includes a series of simple
relationships between moment of inertia, Modulus of Elasticity and beam geometry, saves
a lot of computation time on computer. More complicated relationships are required when
joint stiffness is implemented in the model of a structure. Another feature for using rigid
joints is that the required data is minimal for structural analysis. Extra data is needed when
joint stiffness is considered in the analysis.
Because of the difficulties due to the joint modelling in structural analysis, standard codes
for design and analysis of offshore structures have not addressed specific recommendations
for including the joint stiffness in the global analysis. Design codes nearly all have made
general comments on using accurate structural analysis methods.
The American Design Code API-RP2A (19th edition, 1991) states that: "Brace axial loads
and bending moments essential to the integrity of the structure should be included in the
calculation of acting punching shear."
The British Code of Standard B S 6325 (1982) gives no guidance for global analysis (UEG,
1985). Both API and B S codes allow the maximum eccentricity of D/4 for multi-braced
joints.
The Norwegian Code of Standard, D N V , requires that "Any local flexibility in the
connection of a member to a joint, which is of importance for the force distribution in the
structure should be accounted for in the stiffness or flexibility matrix of the total structure."
D N V requires the consideration of eccentricity in the analysis and recommends two
formulae, stated in Section 6.4.1, to determine the spring stiffness of T-joints for inplane
and out of plane bending, respectively.
The significance and necessity of conducting research on finding accurate, applicable and
simple methods to incorporate the stiffness of joints in the analysis of structures, especially
offshore platforms, can be observed in the practice. In this thesis a new method (Kohoutek,
1985a) which benefits from the dynamic behaviour of structures is examined to model the
joint stiffness of tubular T-joints. The study therefore has focused on the simplified models
of joints to avoid uncertainties due to complicated modelling.
Chapter 1: Introduction 11
1.6 Scope of this thesis
This study attempts to investigate two structural aspects of tubular joints using dynamic
methods of analysis in theory and experiment. These aspects are joint stiffness (flexibility)
and stress concentration. Joint stiffness is calculated based on a dynamic method using the
measured natural frequencies. According to the author's knowledge, this method is
examined for the first time on the tubular joints in this thesis. Using the same method, a
parametric study into the stiffness of tubular joints is also conducted based on the results of
Finite Element analysis. Finally the effects of joint stiffness on the structural behaviour and
especially fatigue life of an offshore tower is investigated.
Only tubular T-joints are studied here because of the time limit, but the method can be used
for other types of joints. The joints are assumed to have linear material properties, since
many fixed offshore structures respond linearly to the important fatiguing waves (Barltrop
and Adams, 1991).
Another feature of tubular joints, which is the focus of this study is the dynamic strain
measurements of tubular joints, as a means to determine stress concentration factor, and the
errors that can be made by performing static tests instead of dynamic tests. Determination
of SCFs are usually carried out through static tests, where the measurements in this study
show that there are some changes when dynamic loading is considered. It is shown here
that those changes are as significant, if not more, as those created by other parameters that
have been studied before.
The Objectives of this study are:
1) applying a new dynamic method to determine the stiffness of tubular joints
experimentally and analytically,
2) establishing a set of parametric formulae for stiffness of tubular T-joints,
3) investigation of dynamic loading effects on measurements of stress or strain
concentration factors in tubular joints, and
4) investigation of joint stiffness consideration on the structural behaviour and
fatigue life of tubular joints in offshore towers.
This thesis examines the structural performance of tubular joints as a crucial component in
the whole platform structure and is not concerned with the load calculation on Offshore
platforms. It assumes that only loading varies with time and therefore causes stress fluctu-
ation in the structural members in time.
Chapter 1: Introduction
12
1.7 Summary
In this chapter a general introduction was given to the different types of offshore platforms
and the terminologies used to identify tubular joints. Tubular joints are the most critical
structural element in offshore structures and modelled to be rigid in today structural
analysis practice. The main reason for this assumption is the lack of an efficient method of
analysis for considering the joint stiffness. D u e to the same reason, the codes of standards
for offshore structures do not generally address any method to include the flexibility of
joints in the analysis.
The main reason that tubular joints impose special attention in the design and analysis of
offshore structures is fatigue problem. This thesis attempt to improve the estimation
techniques for fatigue life of tubular joints through establishing more realistic methods of
analysis.
Chapter 2: Literature Survey 14
2.1 Introduction
Extensive work has been carried out on tubular joints with the trend towards understanding
the fatigue behaviour and providing analysis methods to estimate the fatigue life of joints.
The major contribution has been made by oil companies in U S A and Europe to improve
the design, construction and maintenance procedures applied to offshore structures. With
regard to the stiffness behaviour of tubular joints, however, little published material is
available. Although the study on rigidity of open section joints goes back to the 1910s, the
topic of tubular joints was only attended to in the 1970s.
This chapter begins with the study of semi-rigidity of joints, in general. The rest of the
chapter is specifically about tubular joints, where results of other studies are reported and
discussed.
2.2 General
The importance of the end restraint provided by semi-rigid joints was realised over seventy
years ago. Wilson and Moore (1917) first investigated the flexibility of riveted structural
joints in 1917. Research workers in Britain (SSRC, 1931, 1934,1936), Canada (Young and
Jackson, 1934) and the United States (Rathbun, 1936), in three separate investigations
during the 1930s, measured the relations between end moment and relative angle changes
at beam to column connections in an attempt to provide data for semi-rigid joint design.
Since then numerous tests on riveted, bolted and welded joints have been reported.
Batho and Rowan (SSRC, 1931) proposed a graphical method for predicting the end
restraint provided by a connection for which the experimentally obtained moment-rotation
relationship was known. Young and Jackson (1934) investigated 1) the rotational capacity
and end restraint provided by joints to reduce the beam moment due to gravity loads and
also 2) the ability of joints in frames to resist horizontal deflections due to lateral or wind
loads. Methods of incorporating semi-rigid end restraint into slope-deflection and moment-
distribution methods were proposed by both Baker (SSRC, 1931) and Rathbun (1936)
independently.
From the early investigations into joint behaviour some possible economies were realised.
According to British investigations (SSRC, 1934), savings of as much as 20 per cent could
be achieved on the design of beams in frames by taking advantage of end restraint.
Chapter 2: Literature Survey _
2.3 Tubular joints
Because this thesis is directly related to tubular joints, the literature survey has mostly
concentrated on this type of joints. It should be noted here that only static methods were
used in most of the reviewed articles and references. A m o n g them only the work by
Springfield and Brunair (1989) is based on a dynamic method of analyses.
Investigations on tubular joints are more concerned with stress and strain concentration
than flexibility analysis, nevertheless the same theory and models apply to both cases. The
methods used to study the behaviour of tubular joints can be categorised as:
1) analytical methods,
2) experimental and semi-experimental methods, and
3) numerical methods.
Each of the above methods has some advantages and deficiencies, however experimental
techniques can produce the most accurate results provided the test set-up is made according
to the assumptions adopted for the tests. Experimental accuracy and realisation of actual
conditions are very important for interpreting the experimental results. Analytical methods
based on plate and shell theory become very complicated when dealing with tubular joints.
They can, however, produce fast and relatively accurate results where applicable.
Numerical methods are those procedures which attempt to reach the solution of a problem
by somehow discretizing the domain of the function being studied. It is tried here to
differentiate between an analytical and a numerical method. Analytical procedures are
based on the theories of continuum mechanics and aim at the exact solution. However,
numerical methods approximate the exact solution. The Finite Element method is one of
the most powerful numerical methods for studying the behaviour of structures. However,
the complicated behaviour of tubular joints creates some inaccuracy and difficulty when
the Finite Element method is applied to the joints.
2.3.1 Analytical methods
Because of the complicated geometry of tubular joints, realistic analytical solutions to the
deformation or stress problems at the intersection of the tubes have not become available.
Investigators have tried to solve the problem by applying simplifying assumptions for
various loading cases.
Kellogg (1956) replaced the brace load with an equivalent distributed load shown in Figure
2.1. Based on the theory of beam on an elastic foundation, Kellogg derived the maximum
stress under the equivalent load. This method considers only axial load and/or inplane
bending moment on the brace. It gives approximate stress values for the chord and does not
have any reference to the brace (UEG, 1985).
Chapter 2: Literature Survey
16
Figure 2.1. Kellogg's models of a tubular joint
Another example of this type of analysis is Bijlaard's method (1955) which used a double
Fourier series to show the displacement field of a cylinder subjected to a rectangular
distributed load. Although the moment and deflections were computed for point O in the
model shown in Figure 2.2, equations were introduced for obtaining the moments at the
edges of the loaded area. The method needs to take into account a large number of terms in
the Fourier series to give a relatively accurate result. For example Rodabaugh (1980) used
21 terms in the hoop direction and 81 terms in the axial direction to determine the behaviour
ofK-joints (UEG, 1985).
r
Uniformly distributed radial loading
Q_3
. < ,
Circumferencial moment longitudinal moment
Figure 2.2. Model of cylindrical vessel used by Bijlaard
Chapter 2: Literature Survey - 1_J_
Despite the agreement between experimental data and Bijlaard's results, the method is too
simplified for tubular joints. It may, however, be applied to the joints with small P ratio for
preliminary design purposes.
Dundrova (1965) has presented one of the most complete theoretical studies. She has
analysed a T-joint under axial load based on the classical theory of cylindrical shells. Her
solution finds the distribution of the forces acting on the chord wall by imposing
compatibility condition between brace axial displacement and chord wall deformation.
However, brace bending stiffness is not considered in Dundrova's solution. She was the
first one who considered the brace explicitly in the analysis.
Tubular joint flexibility was studied in a report by Holmas et al. (1985) using the classical
shell theory. The range of p considered by Holmas is between 0.1 to 0.5. The Donnell form
was used to express the forces and moments on a shell element as shown in Figure 2.3.
Bending moment and axial force in the brace were replaced by the equivalent forces,
shown in Figure 2.4.
Figure 2.3. Shell element used by Holmas Figure 2.4. Extra DOF to express local joint
behaviour used by Holmas
It appears that the model by Holmas is similar to the Dundrova's model. Holmas's model
suggest three extra degrees of freedom for every brace attachment, which are one
translational, and two inplane and out of plane bending degrees of freedom. The report
presents the variation of the axial and IPB stiffness of a T-joint for various D/T ratios and p
values. A model is suggested by Holmas for considering the high axial stiffness of the
brace based on the collocation method, but the bending stiffness of the brace is not taken
into account in this model.
B. Chen et al. (1990) have investigated the local joint flexibility of simple T, Y and
symmetrical K-joints for axial and inplane bending loads. They have used the classical
Chapter 2: Literature Survey 18
theory of thin shells and the Finite Element method to analyse tubular joints with the chord
and braces treated as substructures of thin shells while the intersection curve between any
two substructures is discretized into finite elements. Chen et al. have recommended a
formula for the stiffness matrix of a symmetrical simple K-joint. They have reached a good
agreement with other formulae by D N V (1977), Fessler (1986) and Ueda (1990) and some
experimental results by Tebbett (1982).
T. Chen et al. (1990) have introduced a similar analytical method to the method by B. Chen
(1990), using the two models by Holmas (1985) and Ueda (1986) for definition of joint
flexibility. These two models are based on solutions of shell equations and Finite Element
analysis, respectively. The model by T. Chen has the features of simple computations and
low C P U time. T. Chen has studied axial and inplane bending flexibility of T, Y and T Y -
joints.
2.3.2 Experimental and semi-experimental methods
There are some parametric formulae that are referred to in this section. The reader may see
Chapter 6 for those formulae and their ranges of validity.
Experimental methods
The theories of structures and continuous media are not used in these procedures. A
physical model which can range from small to full scale in size is tested under the
conditions similar to the real structure. The model can represent the whole structure or a
component of it.
In the study of tubular joints, test specimens selected earlier were from steel. Synthetic
materials such as acrylic and epoxy resin were used later as substitutes for steel since they
are cheaper, easier to handle and more flexible. Experiments are usually carried out by
loading the joints through static forces and measuring the desired quantity, which can be a
strain in any direction or displacement of a location with respect to a datum. Test
specimens from synthetic materials are on a small scale, whereas those from steel could be
the same scale as the prototype. Numerical methods are usually employed for curve fitting
the test results, where generally an equation or formula is established to be used for
analysis and design. Parametric formulae for stress concentration factors is a popular
example of the application of experimental methods to tubular joints.
The photoelasticity method is also an experimental technique involved in the experimental
stress analysis of tubular joints, where three dimensional stress distribution can be
deterrnined. The method is restricted to stress analysis and unless a relationship between
flexibility and stress is employed, it can not be used for the study of joint flexibility.
Chapter 2: Literature Survey 19
Fessler et al. (1981) developed a procedure to define and measure the flexibility of tubular
joints. Three loading modes were considered: 1) axial tension, 2) inplane bending moment,
and 3) out of plane bending moment. Fessler only considered T and non-overlapping Y-
joints by testing 25 joints made of precision-cast epoxy resin tubes. Methods, based on
experimental results, were proposed to determine the joint flexibilities of the different
deformation modes. A n equivalent brace length was proposed to consider the flexibility of
typical joints when the customary line model is used. A line model is constructed of one
dimensional beam elements being connected at the joints.
Fessler concluded that further work should include an analysis of simple frames of typical
structures. It appears that the experimental method proposed by Fessler includes a
relatively time consuming procedure and can be costly in terms of test equipment. Figure
2.5 shows the rig used for loading the test specimens. Deflections were directly measured
at various locations.
=
Load ring fixed-
7
to brace
Channel section
support
Tape
transducer
(H
'-M
m
TL
Out of plane bending
___
Support stand reacts load
Strain gauges
______
XI
%4
Load frame-
Axial tension
Figure 2.5. Test rig used by Fessler (1981)
In another work, Fessler et al. (1986a) developed a set of parametric formulae for IPB,
O P B and axial deformation of brace in single brace tubular joints, using the same method
as in the 1981 paper. There were 27 tests on araldite models covering the c o m m o n range of
parameters in offshore structures. In comparison with experimental results Fessler's
formulae overestimate the bending stiffness of T and Y-joints.
In a companion paper, Fessler et al. (1986b) presented a set of equations for the cross-
flexibility between any two braces which may be in any orthogonal planes at a joint. This
work was also based on the same experimental procedures and actually on the same test
Chapter 2: Literature Survey 20
specimens as the other paper (1986a) by the same authors. The measurements on the end of
fictitious unloaded braces were determined from the measurements of the single brace joint
models. In both papers, the effect of the variations in brace wall thickness on joint
flexibility has been ignored. For non-overlapping joints, the proposed parametric equations
may overestimate the flexibility up to 7 0 % compared to the measured data when the
flexibility is significant.
Semi-experimental methods
W h e n compared with the experimental methods, semi-experimental procedures also benefit
from the analytical methods of structural analysis. In these procedures, a mathematical
model is employed and tuned using test results. Punching shear model, shown in Figure
2.6, is an example of this method. The punching shear stress, v
p
, is assumed to be
uniformly distributed. So it can be written as:
N
v_ =
p
ndt
(2.1)
in which N, d and t are the axial force, diameter and thickness of the brace, respectively.
The axial force in terms of shear stress would be:
N=\
p
ndt (2.2)
Design codes give the allowable values of the punching shear stress for different
geometrical parameters. The values have been derived from experiments on various test
models and then stated in the analytical form of punching shear stress formula.
Figure 2.6. Stress distribution assumed in Punching Shear Model
Chapter 2: Literature Survey 7
The method used in this study is a semi-experimental technique in which an unknown
analytical parameter is determined by experiment. A work on the support flexibility of pre-
tensioned cables has been carried out by Springfield and Brunair (1989) that implements an
approach similar to the method adopted in this thesis. The end fixity, as the main objective,
and the bending stiffness (EL) of an electrical transmission line were determined by
measurements of the displacements at certain locations of the line when it was vibrating
under a certain natural frequency. The theoretical model used by Springfield and Brunair is
an axially loaded, transversely vibrating beam supported at the ends through rotational
springs. Springfield and Brunair have concluded that consideration of end fixity leads to a
conductor bending stress of approximately one-lhird the value given by assuming a rigid
end. It was found in this study that only Springfield and Brunair have used a dynamic
method of analysis.
2.3.3 Numerical methods
Numerical and analytical methods were facilitated with the advent of computer, resulting in
the development of the analysis procedures in the theory of structures. The Finite Element
method, the dynamic deformation method and the flexibility method are some examples.
The dynamic deformation method is an improved version of the slope deflection method
where the inertia forces are also considered (Kolousek, 1939,1973).
The Finite Element substructuring is the most effective and non-expensive methods which
can be effectively used for parametric studies. However, there is always this argument that
the method does not include any deficiency unless purposely assumed in the model.
Furthermore, because the F E M is a numerical procedure its predictions can not be
absolutely reliable without any experimental verification.
Bouwkamp et al. (1980) developed a new procedure involving a modified three
dimensional Finite Element formulation for the modelling of a tubular joint substructure
and its subsequent insertion into a complete offshore platform computer model. The
substructuring technique used by Bouwkamp allows fast modelling of the tower frame
without having to do finite element modelling of each joint, when the super-element is
available. The technique is based on the results of the Finite Element analysis. There are
also some simplifications to easily model the super-element. Figure 2.7 shows a typical
substructure of a joint used by Bouwkamp.
Chapter 2: Literature Survey
22
T-brace.
Primary mesh
Embedded masrri
(-agonal brace
main chord
T-brace intersection
X
__
Diagonal intersection
-meshes
constrained'
_r
Figure 2.7. Model of joint substructure used by Bouwkamp (1980)
Efthymiou (1985) has reported a Finite Element study on the local stiffness of unstiffened
tubular T, Y and K-joints subjected to inplane and out of plane bending. H e has defined the
local joint stiffness as the applied moment at the brace divided by the local joint rotation.
The rotation of the brace end due to the joint flexibility was calculated by deducting the
beam type rotation from the total rotation of the brace end. H e measured the rotation at the
end of the brace as shown in Figure 2.8. This is different from the other methods in which
the measurements are usually made on the chord wall.
M
QL_
4|
'ML.,,
Brace'
ii
r - " _ _ "
r
Chord
M
ib
Brace bending stiffness
Crown
0
Chord bending
stiffness 12EIe/le
a) Physical model b) Theoretical model
Figure 2.8. Rotations measured by Efthymiou for calculation of joint flexibility (1985)
Chapter 2: Literature Survey
23
The F E program used by Efthymiou was P M B S H E L L , which had a thin shell element
implemented. T o verify the performance of P M B S H E L L , Efthymiou re-analysed one
geometry using another program called S A T E , which had a combination of plate and
membrane elements. The results of these two analyses showed a very good agreement. H e
established a set of parametric formulae, based on 24 F E analyses, for T, Y, and K-joints.
The Efthymiou's equations for T and Y-joints predict local stiffness to within 1 5 % of the
stiffness values used for curve fitting. The equations for K-joints are somewhat less
accurate. Their predictions are expected to be within 3 0 % of the measured stiffness. The
parameters considered by Efthymiou were |$ and y. His equations are inclusive of common
joint types used in offshore structures, but the database that he has used to establish the
equations does not seem to have adequate data. Furthermore, Efthymiou's study is based
only on F E analysis results and does not have any comparison with experimental findings.
Ueda et al. (1986) have developed a model for tubular joints. The model takes account of
joint flexibility in elastic as well as elastic-plastic ranges based on elastic fully-plastic load-
displacement relationship. It is stated by Ueda that the geometry of tubular joints makes it
difficult to obtain closed form analytical solutions to evaluate load displacement
relationships. In this respect, the method which is chosen in this thesis has a highly
theoretical base. It determines the natural frequency of a tubular joint from the measurement
and then employs it in the analytical model to produce a relationship between load and
displacement. The reader is referred to Chapter 4 for further explanation.
Ueda has proposed line elements for modelling the joint local behaviour, as shown in Figure
2.9. Elements *c' represent the local behaviour of the chord wall in Ueda's model. The
stiffness matrices for the elements are taken from another reference by the same authors.
The method is used for both elastic as well as plastic zones.
a
a
- - i
c,
- .
Chord
b
*
i
Branch stub
- -<
a
a
- -<
CJ
T-joint Y-joint
Branch stub
Chord
Figure 2.9. Joint model proposed by Ueda (1986)
The proposed method by Ueda for considering local joint behaviour is computationally
simple and does not need a great deal of computer memory. However, it is still based on
preliminary analysis by the Finite Element method to obtain the stiffness of the joints. The
Chapter 2: Literature Survey
24
method actually implements the stiffness results of a finite element analysis into a simpler
line model. Besides, the computational nature of the method allows no modification in the
joint model due to imperfections and other complicating factors involved in manufacturing
and fabrication. Such an ability could remove the approximations introduced by the Finite
Element and generally correct the F E model using experimental data.
In another paper, Ueda et al. (1990) developed a set of formulae for stiffness of T and Y-
joints. The database used for establishing the formulae was taken from F E analysis and
included 11 samples for IPB mode and 7 samples for axial deformation of brace.
The relationships between the stiffness of T and Y-joints used by Ueda et al., especially
axial stiffness, do not appear to be consistent with the results of others. For example Fessler
(1986) shows that:
fcY('
a
') = fcrsin"
2
-
1
^, and k
Y
(tPB) = &
T
sin"
1,22
0 (2.3)
in which ky and kj are the stiffness of Y and T-joints, respectively. 0 is the brace angle in a
Y-joint. Efthymiou has obtained the following relation:
J.
Y
aPB) = Min-
(P+a4)
e (2.4)
Whereas, Ueda used the same axial stiffness for T and Y-joints. To determine the IPB
stiffness of a Y-joint, Ueda used the stiffness value of a T-joint divided by sinG, where 9 is
the angle of diagonal brace.
Souissi (1990) has carried out a study on the flexibility of tubular T-joints using the Finite
Element Method. H e established a super-element to model a joint and attributed its property
to the fictitious centre nodes that were at the end of each tube, chord and brace (Figure
2.10).
Y 3
i i
Figure 2.10. Joint super-element used by Souissi (1990)
Chapter 2: Literature Survey
25
Souissi considered inplane bending, out of plane bending and axial loading and performed
18 analyses for each case. His results showed good agreement with Efthymiou's (1985)
results. H e has recommended that corrective factors can be applied for each loading case to
consider the effects of x on joint flexibility.
2.4 Effects of tubular joint flexibility on tower analysis
This subject has been researched since the late 1970s. F e w references are available about the
effects of joint flexibility on the behaviour of tower structures. The Underwater Engineering
Group (UEG), which is a British non-profit-making organisation produced a comprehensive
report called U R 2 2 on the joint flexibility effects. This section provides a summary of this
report as well as other studies.
B o u w k a m p et al. (1980) summarised the results of a limited study into the effects of tubular
joint flexibility on the structural behaviour of deep water fixed offshore towers. Bouwkamp
produced a model, using Finite Element analysis results, to incorporate the joint flexibility
into the structural analysis. In order to illustrate the procedures used to assess the effect of
flexible joints, a two dimensional 330m high tower frame was analysed under dead and
wave loads, using the developed joint model as well as the so called line model. In the latter
model, joint effects are neglected. It was concluded that the effects of joint flexibility on the
structural behaviour of offshore towers can be significant The nature and magnitude of
these effects are dependent not only on the tower height, but also on its geometrical and
structural configuration. The effects were noted in the higher modes of vibration and in the
deflected shape of the tower under static loads. It was observed that joint flexibility effects
are more pronounced when stiffness of the member intersecting at a joint is relatively high.
The effect of joint flexibility on the deflected shape was seen to be very small for the nodes
at the top of the tower. However, larger displacements were observed for the joint model
below -170m of water level, with a maximum increase of 5 0 % over the line model at -300m.
Regarding member forces and moments, B o u w k a m p has shown: 1) a slight increase in
calculated leg axial forces (up to 2 % higher) and a considerable reduction in calculated
brace axial forces (up to 2 0 % ) ; 2) a modified distribution of pile loads with load transferred
towards piles through main legs; 3) an increase of up to five fold in legs moments.
Joint flexibility consideration in dynamic analysis was shown to lead to lengthening of the
fundamental periods particularly for higher modes, where changes in order of the mode
shapes were also observed.
Tebbett (1982) showed the effectiveness of grouting the legs of fixed jacket offshore
platforms which has become important with regard to the reappraisal of steel jacket
structures. T o do so, he has placed emphasis on considering the flexibility of tubular joints
in the analysis of the jacket structures. Tebbett has concluded that the effects of local joint
Chapter 2: Literature Survey
26
flexibility can be significant and should, if possible, be included in the structural analysis
during the reappraisal of jacket structures. Furthermore, if grouting is being considered, the
reduced local joint flexibility should be accounted for in the analysis.
Kawashima and Fujimoto (1984) checked their model by testing an L-frame and a portal
frame. Kawashima only studied the flexibility effects on the mode shapes and natural
frequencies by conducting dynamic analysis. The test models and the results obtained by
Kawashima and Fujimoto are shown in Figure 2.11. They obtained a good agreement
between the analytical results of the joint model and experimental results, especially for the
lower natural frequencies. The effect of flexibility consideration on the natural frequencies
showed a variation from - 2 5 % to 0 % between the calculated results. The - 2 5 % change
occurred for the first natural frequency of the portal frame. Joint stiffness introduced a
m a x i m u m change of 1 0 % to the calculated natural frequencies of the L-frame.
600
h-
900
Table 2.1a. Kawashima's results for L-type frame
Mode
1
2
3
4
Rigid
Cal.
23.3
51.4
74.5
142.8
Semi-rigid
k-\ 200kg. cm/rad
C=2.5kg.cm.sec
L
Cal.
21.6
47.8
67.2
142.9
Exp.
21
46
65
138
a) L-type frame and its natural frequencies obtained from experiment and analytical models
900
(-
600
Table 2.,
Mode
1
2
3
4
5
6
7
8
b. Kawashima's results for portal frame
Rigid |
|
Cal. |i
9.4
26.6
65.8 I
74.9
100.4 |
173.1 |
220.2 1
232.7 J
Semi-rigid
fc=1200kg.cm/rad
C=2.5kc
Cal.
7.1
19.6
62.6
65.3
73.9
163.1
200.4
204.9
l.cm.sec
Exp.
6.9
20
59
65
73
144
180
217
b) Portal frame and its natural frequencies obtained from experiment and analytical models
Figure 2.11. The results of Kawashima's analyses on the frames with rigid and semi-rigid joints
Chapter 2: Literature Survey
27
Matsui et al. (1984) have studied the behaviour of truss beam columns composed of tubular
sections. Matsui considered the effect of joint flexibility on the buckling behaviour of the
web members with large diameter-thickness ratio. The flexibility analysis of Matsui is
based on a spring model from Sakamoto and Minoshima (1979). The results of Matsui's
analysis indicate a maximum of 1 0 % difference in buckling strength of a truss when only
bending moment is applied to the chords.
Ueda et al. (1986) carried out a parametric study on five K-braced, five-storeyed two
dimensional tubular frames as shown in Figure 2.12. Three horizontal point loads were
considered in Ueda's analyses.
4000
4000
4000
4000
4000
Figure 2.12. K-braced frame analysed by Ueda
and its load cases
Table 2.2. Joint specification in Ueda's analyses
Model Model D T Initial
No. type load (kgf)
R15
R20
R30
R40 R
R50
R59
R67
F15
F20
F30
F40 F
F50
F30D
F40D
R: Rigid joints
F: Flexible joints
Chord
Horizontal brace:
Diagonal brace:
Yield stress:
1000
1000
64
50
33
25
20
17
15
64
50
33
25
20
33
25
DxT (Table 2.2)
400x25
400x25
mm
mm
70 kgtfmm
2
1410000
1100000
730000
550000
440000
374000
330000
1410000
1100000
730000
550000
440000
730000
550000
Ueda investigated the effects of the joint flexibility and strength on the structural behaviour
and collapse loads of the K-braced frames. It was found that joint flexibility may have only
a little effect on buckling of braces, but joint strength may have a great influence upon
collapse modes and strength. The lateral stiffness of the K-frame with D/T = 50 decreased
by up to 4 6 % when joint stiffness was considered in the analysis. There was, however, less
reduction of lateral stiffness when lower D/T ratios were assumed in the analysis.
The significance of the Ueda's study is investigating the effects of tubular joints on the
ultimate strength of tubular frames. The 3 point loads considered by Ueda in his study do
not simulate the loading from waves, current, etc. which exist in sea environment. The
Chapter 2: Literature Survey 28
results of Ueda's study could be more applicable to the offshore structures if different
loading were used.
UEG Report, Node Flexibility and its Effects on jacket Structures (1984, UR22)
This report presents an investigation into the effects of chord wall flexibility at brace
connections on the behaviour of oil production jacket structures. It has considered the
effects of joint flexibility on the inplane deflections, axial forces, bending moments, brace
buckling and natural frequencies of three different 100m tall vertical plane frames. The
overall geometry of the frames are shown in Figure 2.12. They have been modelled using
two dimensional beam elements with three inplane degrees of freedom at each end, two
translations and one rotation.
36.24m 63.094m 36.240m
H H H H
structure 1 structure 2 structure 3
Figure 2.12. Frame models analysed in UR22 report by UEG (1984)
A simple representation of the joints was selected in the study. One nodal point was
provided on the chord and one on each brace at the brace to chord wall intersection as
shown in Figure 2.13. The nodal points 2, 3 or 4 were then all connected by a stiffness
matrix derived from the flexibility matrices provided by Fessler (1981).
Figure 2.13. Nodal points considered in UR22 Study to represent a joint
Chapter 2: Literature Survey
29
T w o types of analysis were carried out, one incorporated flexibility of the joints based on
the Fessler's model, the other did not consider flexibility which was called conventional
analysis as the braces were extended to the chord centre lines.
The various joints used were identified by three characters T I N (Type Intersection
Number). Type may be C-Conventional or M-matrix, T describes the intersection of the
braces and chords which is P-intersect at Point or E as Eccentric. 'N' is the joint Number
corresponding to the geometrical ratios characterising the joint geometry. The joint numbers
of the different geometrical parameters are shown in Table 2.2.
Table 2.2. Joint parameters used in UR22 Study
joint no.
1
2
3
4
D/T
25.3
50.6
25.3
25.3
d/D
0.53
0.53
0.33
0.75
For example M P 3 refers to the analysis, using matrix formulation for the joints, where the
braces are intersecting at a point having D/T=25.3 and _W)=0.33. Four load cases were
applied to each structure. The first load case was a point load applied at the top of the
frame. The other three were distributed wave load cases derived from a representative 100-
year storm wave with different phase angles: 0, 90, and 45. The following results were
obtained from the analyses.
1) Global Deflections
The introduction of the joint flexibility into the analysis, made differences of up to
1 3 % to the overall sway of the structures analysed. A comparison of the
deflections for the structures with different joint types is given in Table 2.3.
2) Effect of Flexibility on Axial Forces
This effect was found to be negligible. The biggest change between the
conventional and flexible analysis was 1.5%. The maximum axial stress change was
less than 1 N/ram
2
.
3) Effect of Flexibility on Bending Moments
The largest change in brace end moment found was in structure 1 with joints M P 2 ,
where a horizontal brace moment increased to about three times the conventional
rigid frame analysis value implying a 2 0 0 % change. The largest variation of
bending stress for structure 2 was 6 0 % . Structure 3 had the largest change of
about 5 0 % . These changes are corresponding to a combination of the analysis
Chapter 2: Literature Survey
30
results of load case 2 and load case 3. The bending stress changes for all the
various structures and joints, under the wave load with a 45 phase angle, are
shown in Table 2.3. The largest stress changes in the structures under the same
loading were:
Structure 1: 30N/mm
2
,
Structure 2: 29 N/mm
2
,
Structure 3: 4 N/mm
2
.
Table 2.3. Summary ofc
Change from
conventional analysis
Deflection change%
Chord
Axial stress 45 brace
change(NmnT
2
) 90 brace
Chord
Bending stress 45 brace
change% 90 brace
Buckling load 45 brace
change% 90 brace
1
2
3
4
5
Natural frequency 6
changes% 7
(rigid/semi-rigid) 8
9
10
11
12
13
hangesfrom UEG re port on joint flexibility
Structure 1
MPl
0
0
0
0
4
15
44
4











MP2 MP3
5 2
0 0
0.5 0
0 0
7 7
20 15
94 50
-9
2
3
1
N
3
3
1
5
15
0
N
MP4
-1
0
0
0
4
15
22


Structure 2
MPl
3
0
0
0
7
11
8


MP2 MP3
13 5
0
0
0
14 14
-25 -29
14 14
-10
-13
6
9
3
26
3
4
3
N
17
15
26
MP4
-1


14
13
6


Structure 3
MPl
1
0
0
5
13



MP2
5
0
0
5
6

-12
2
1
3
82
13
8
1
3
2
0
MP3 MP4
1 -1


5 9
6 6



4) Effect of Joint Flexibility on Brace Buckling
The effect of joint flexibility on buckling load of the braces was determined in the
study. The results are shown in Table 2.3. The buckling load was reduced by about
1 0 % between the conventional CP1 and the most flexible M P 2 analysis. This was
caused by the flexible joints increasing the effective length of the brace.
Chapter 2: Literature Survey
31
5) Effect of Joint Flexibility on the Vibration Characteristics of Jacket Structures
The first few natural frequencies and their corresponding m o d e shapes were
calculated for each structure with the conventional C P 1 and the most flexible M P 2
joints. The natural frequencies of similar m o d e shapes were compared. Table 2.3
summarises the natural frequencies and reports the proportional changes. The
changes in the natural frequencies of corresponding modes were on average 4 % ,
1 2 % and 1 1 % for structures 1, 2 and 3, shown in Figure 2.12, respectively. The
greatest change in natural frequency of similar modes was 8 2 % and occurred for mode
shape 5 in structure 3.
The study showed that the increase in bending stress caused by incorporating joint
eccentricity of DIA in the conventional analysis was similar to that caused by joint flexibility.
It was concluded that the effects of joint eccentricity coupled with those of joint flexibility
could therefore be significant
The report U R 2 2 indicates the significance of incorporating joint flexibility of tubular joints
into the analysis of offshore towers. This report considers only one joint modelling
technique that is using joint stiffness matrices provided by Fessler. Other simulation
techniques could produce different results. Furthermore, it focuses only on the different
aspects of joint flexibility in the structural analysis, whereas an analysis of fatigue life seems
to show the significance of joint flexibility consideration more clearly.
Chapter 8 of this thesis reports the results of a similar analysis to U R 2 2 using the theoretical
model suggested in Chapter 3. The joint flexibility effect on the fatigue life is also
investigated in Chapter 8.
The effect of joint rigidity on the buckling behaviour of tubular members in trusses and
frames has been investigated by Jong and Wardenier (1987), using Finite Element method.
The moment rotation relationship of tubular joints have been employed in the derivation of
buckling load curve suggested by Jong. A formula is presented to determine the critical joint
stiffness of axially compressed members with semi-rigid joints. B y definition, there is not a
notable gain in the buckling load of a member when the joint stiffness is higher than the
critical value. The critical joint stiffness was found to be:
(X-20)
2
/
>ifX<2()thenX = 2() (25)
1
1400 L
C
2
=3Q
(2.6)
Chapter 2: Literature Survey
32
The relationship suggested by Jong and Wardenier for buckling load of tubular members
with semi-rigid joints is:
/ V > ( 1 ) N
m
1
10000
N
2
= 0.98AL
(2.7)
(2.8)
The terms in Equations (2.5) to (2.8) are shown in Figure 2.14, below:
N
0
the buckling load for a pin ended member
N-\ the buckling load belonging to C-\
Afc the buckling load belonging to C^
No. the buckling load for a member with rigid joints
X the slendemess ratio of the pin ended
member in compression
100
Joint rigidity, C
Figure 2.14. Predicted buckling load by Jong (1987)
Joint rigidities were assumed to be the same at both ends of a member in compression. The
parametric formula for the joint stiffness allows the adjustment of the joint geometry to
achieve the desired strength for a compressive member intersecting at the joint
The proposed method by Jong and Wardenier is to some extent complicated for design
purposes. However, it is one of the few works on the buckling of tubular members
considering semi-rigid joints and the accuracy of the method also needs to be verified by
experimental analysis.
T. Chen et al. (1990) analysed a 5-storeyed tower, shown in Figure 2.15, considering
flexible tubular joints based on the data by Holmas and Ueda. The results of Chen's analyses
are reported in Table 2.4.
Chapter 2: Literature Survey
33
P1 = P2 = P3 =10000 kg
D=100cm
d=40cm
7=23 cm
t= 1.0 cm
E=2.1x10 kg/cm
v=0.3
Figure 2.15. Tower analysed by T. Chen (1990)
Table 2.4. Effect of flexibility consideration in the analysis, Chen T. (1990)
Umoaicm)
Vmax(cm)
Qmaxfrad)
Nmaxfkg)
Qmax(kg)
Mmax(kg-cm)
computed results
based on
rigid joints
7.198
0.365
-1.09E-3
3.45E5
1.26E5
1.99E6
computed results
based on
Holmas'joints
8.784
0.386
-1.20E-3
4.27E5
1.54E5
1.71E6
computed results
based on
Ueda's joints
8.722
0.379
-1.21E-3
4.23E5
1.53E5
1.92E6
Chen's results show a good agreement between the two methods by Holmas and Ueda. The
biggest difference was between the maximum bending moments calculated by the two
methods. Holmas's model produced a 1 4 % change in bending moment whereas Ueda's
model caused only a 4 % change.
According to Chen's results there was a maximum change of about 2 0 % in the horizontal
displacements, a 2 3 % change in axial forces and a 1 0 % change in bending moments when
semi-rigid joints were employed in the analysis.
Chen's results are due to a loading composed of three point loads as shown in Figure 2.15.
This type of loading does not occur as frequent as wave loading in the sea environment
Therefore, the results are not very applicable to the offshore structures. However, Chen's
results generally show the effects of joint stiffness on the behaviour of structures.
Souissi (1990) also compared the analysis results of two frames (Figure 2.16), one with
flexible and the other with rigid joints (conventional analysis).
x,u
Chapter 2: Literature Survey
34
H
_,g _____'__.
8" 77 . 29'
senv-rigid
^M v_
7
1 11/
>! -J-_JL_____2J
2a
6<> te < _^
4' 15
3> 10 \
V
18 24
-9 '
M 23
10m
25
22
8m
8m
D=800mm
d=240mm P=-3
T=16mm "
f=25
t=8mm ^
5
L=4800mm " ^
8m
8m
Figure 2.16. The frame analysed by Souissi (1990)
His results are:
Loading N o 1 : H * 0 , V = 0
1) 9% to 11% underestimation of displacements for the conventional analysis, and
2) overestimation of bending moments up to 3 5 % at joints 10 to 12 for the conventional
analysis.
Loading N o 2: H = 0 , V * 0
overestimation of axial force up 37% for joint 10 and bending moment up to 23% at joints
10 to 12 for the conventional analysis.
Souissi has concluded the need of a simple method to assess the flexibility of joints from
analytical or numerical models.
Recho et al. (1990) have investigated the influence of flexibility on the fatigue design of
tubular T-joints. The joint stiffness has been determined by using Finite Element method
with static condensation technique. This method of stiffness calculation is the same as what
Souissi (1990) carried out in his study. Three series of curves, based on the F E analyses,
were established for the three load cases in T-joints (IPB, O P B and axial loading).
Recho analysed two different structures, shown in Figure 2.17, and calculated the fatigue
life change when the joint flexibility was applied in the analysis.
Chapter 2: Literature Survey
35
1000H
"
5m
6D

a

m
3
2\ Brace
l
3fc

Chord
_
h
frtn
Typel
Same T-joint characteristics
u sed in the two frames
".
4m
1000N
i
Am
i
i
8m
3
*2
r
Chord
'JMVA
I
D = 80cm, T= 1.6 cm
d=. .4 cm, d=0.8 cm
Brace
Brace
1 0 m
Type 2
3
2
i
Chord
1000N
1 J
Figure 2.17. The structures analysed by Recho (1990)
The results obtained for the forces at joint 2 in the two frames are:
Table 2.5. Effect of joint flexibility on internal forces by Recho (1990)
Rigid joint Flexible joint Difference %
Typel Fa, = 1739 N F
axial
=1762N F
axial
= +1.3%
A_n>B = 37 N m Mn
B
= 45 N m Mn
B
= +21.6%
Type 2
^axi_l=763N
M I P B = 4.4 N m
Fi_!=724N
A-TPR = 4.8 N m
Fi_i=-5.1%
MIPB = +9.1%
Recho et al. then calculated the fatigue life of joint 2 using the French Standards ( A R S E M ,
1985) and compared the results of the rigid and semi-rigid analyses. The details of the
fatigue calculation is not given by Recho, however, the influence of flexibility on the fatigue
life of the two frames are reported. Table 2.6 shows this influence as the ratio of 7v*
R
(the
number of cycles to failure when connection is considered rigid) to 7v> (the number of cycles
to failure when connection is considered flexible).
Table 2.6. Fatigue life difference (NtfN
F
)
when joint flexibility is considered
Rupture Type 1 Type 2
at the saddle point 1.04 0.85
Chapter 2: Literature Survey
36
According to the above results, the influence of flexibility consideration on the fatigue life
can be negative or positive. Its cause is related by Recho to the geometry and boundary
conditions of a structure.
The study does not include a realistic loading c o m m o n to the offshore structures since the
loading in sea environment is a distributed load and depends on the wave or current
characteristics, whereas Recho considered one or two point loads in his analysis examples.
Furthermore, when comparing the fatigue life of rigid and flexible joints in Table 2.6, the
location of fatigue rupture is not specified. Failure of a joint under fatigue is because of
rupture at either saddle or crown locations, not likely both. Therefore, only two
comparisons out of the four shown in Table 2.6, correspond to the fatigue life of the joints
analysed by Recho.
Table 2.7 summarises the results of different investigations surveyed in this study
regarding the effect of joint stiffness on the behaviour of tubular structures. It is noted that
wave loading which is the dominant force acting on offshore towers is not considered in
most studies. Furthermore, a consistent trend can not be found in the results. The most
changes are in bending moments. This would be expected since tubular joints possess more
flexibility under inplane bending than axial loading. A s offshore structures transfer external
loads in a truss like fashion, introducing rotationally flexible joints will not make
significant changes in lateral deflections and axial forces. This can be also seen in Table
2.7.
The variations observed in the results of different studies indicate that joint stiffness has
different effects on the analysis results of tubular structures. These effects vary with
structural geometry, configuration, height and loading of the structure.
Table 2.7. Maximum response change for models with rigid joints obtained in various research works when
joint flexibility was considered
Bouwkamp (1981)
U E G (1984)
Ueda (1986)
Chen T. (1990)
Souissi (1990)
Loading
dead load + point
loads at top
point load + wave
load
point load
point load
point load + dead
load
Deflection
+1%
at top
+13%
+13%
(DIT) = 26
+22%
+10%
Axial force
chord: +2%
brace: -20%
+1.5%

+24%

Bending
moment
+500%
-90%

-14%
-35%
Natural
frequencies
-10%
-45%



Recho (1990) point load -5% +22%
Chapter 2: Literature Survey
37
2.5 Model of joint in stiffness method
A typical model which is used commonly to represent a joint in a framed structure is a
linear spring. The idea has taken originally from the moment rotation relationship of a
semi-rigid connection, shown in Figure 2.18. For multi brace joints the joint stiffness may
be adopted in the form of a stiffness matrix to include the interaction of the braces upon
each other.
The linearized form of this relationship was for the first time implemented in the slope
deflection method by Rathbun (1936). His results showed that changing the end
connections of a beam from rigid to semi-rigid is equivalent to lengthening the beam by the
amount of3EI/k. k is the spring stiffness of the beam end.
5
Angle of rotation, 8
Figure 2.18. Moment-rotation relationship for a semi-rigid joint
In 1963, Monforton and Wu expanded the semi-rigid formulation introduced by Rathbun to
a matrix form, suitable for using in computer programs. The formulations developed by
Rathbun or Monforton and W u included only the stiffness properties without any reference
to inertia forces.
The works by Grant (1968), Lionberger and Weaver (1969) involved dynamic analysis but
the joint effects were not considered in the mass matrix formulation. They also used the
spring model to represent the joint flexibility.
Kawashima et al. (1984) modelled a beam with two subsidiary elements at each end. In this
model after derivation of the beam stiffness matrix, lengths of the end beams are taken
Test behaviour
of a semi-rigid joint
Design range
Chapter 2: Literature Survey 38
close to zero which virtually simulate two rotational springs at the ends. Dashpots parallel
to the end springs are considered for modelling the energy dissipation.
In the Kawashima's study, the force displacement relationship for a beam with rotary
springs and dashpots is considered in a matrix form. The dynamic stiffness matrix has been
derived through the transfer matrix in an exact form in which a continuous mass property is
assumed. The elements of this matrix then has been expanded in a series form in terms of
circular frequency. This has resulted in the mass, damping and stiffness matrices whose
elements include two types of parameters designated as fixity factors and damping factors.
Fixity factor varies between zero and one, where it is zero for a hinged and one for a fixed
end. The significance of Kawashima's work is inclusion of the joint flexibility in the mass
and damping matrices. The joint model developed by Kawashima is shown in Figure 2.19.
Q~" 9
vy U -U
L
-U J
actual lenght is zero actual lenght is zero
Figure 2.19. Joint model developed by Kawashima (1984)
Kawashima concluded that the effects of semi-rigid connections depend on the mode
shape. H e also obtained a good agreement between the results of analysis and experiment.
Kohoutek (1985a) used a different formulation from the spring model for a beam with
semi-rigid joints. The boundary conditions assumed by Kohoutek are:
y(0) = 0 y(l) = 0 (2.9a)
-(l-r^AO) + r
v
y{0) = o (i-r,,)y'(/) + _>/(/) = o (2.9b)
in which T is the fixity factor and is zero for a hinged or one for a fixed support. Equations
(2.9b) are aiming at balancing the second and first derivatives of the beam axis
displacement curve at the supports such that their sum becomes zero. y" and / correspond
to the bending moment and the slope of the beam axis. The bending moment is zero for a
hinged support whereas the slope of the beam axis is zero for a fixed support.
Using natural frequencies as a means to determine the rigidity of joints was introduced by
Kohoutek (1985a) who applied the method to open section joints.
Chapter 2: Literature Survey
39
2.5.1 Bending rigidity factor
It is preferred in practice to use a factor between 0 and 1 (or 0 % and 100%) to describe the
stiffness of a joint. U K O S R P has used a similar factor, called degree of fixity (F), to
correct for the effect of chord end fixity and chord length on stress concentration (UEG,
1985). Kohoutek (1985) modelled the joint rigidity by imposing the boundary conditions
shown in Equations (2.9a) and (2.9b). He called the joint rigidity F, which varies from 0 to
1. Monforton and W u (1963) defined a fixity factor v
b
as:
v
"
( V i a t j o i n , 0 =
TT3_77r
(X10)
in which E is the modulus of elasticity, / is moment of inertia and k
t
is the stiffness of the
end i of the beam. This definition was also used later by Grant (1968) and Kawashima
(1984). v
b
varies from 0 to 1, corresponding to k from 0 to oo. The base of Equation (2.10)
can be found from the work by Rathbun (1936) which states:
li = l+3EJ/ki (2.11)
in which // is the equivalent beam length to be used in the slope deflection formulae to take
the joint stiffness into account. Dividing both sides of Equation (2.11) by / and inverting
the result one can obtain:
- =
l
- (2.12)
/.. 1 + 3EI/L
which is actually the definition of v
bi
in the studies by Grant (1968) and Kawashima
(1984).
Chapter 2: Literature Survey
40
2.6 S u m m a r y
The report of the survey on the available literature related to the flexibility of tubular joints
was presented in this chapter. The different methods of investigation on tubular joints were
described with the major applications on the stiffness of joints. These are:
1) analytical methods,
2) experimental and semi-experimental methods, and
3) numerical methods.
It was pointed that the early analytical works on tubular joints by Bijlaard, Dundrova and
Kellogg are too simplified for the real and practical examples of tubular joints. Nevertheless,
the theoretical studies are still used and are seen in the works by Holmas (1985), B. Chen
(1990) and T. Chen (1990). The study of Fessler, which based on the surveyed articles here
is the only comprehensive experimental work on tubular joints, was also described. Fessler
has used araldite models in his tests. There was only one study found where a dynamic
method has been used to calculate the support stiffness of cables. Static techniques have
been used in all other works. The literature survey showed that semi-experimental methods
which have a theoretical and experimental base are not used for flexibility formulation of
tubular joints.
Finite Element method is used extensively in the researches on the flexibility of tubular
joints by Bouwkamp, Efthymiou, Ueda, Souissi and Recho. There is no comparison with
any test results in these studies. The effect of joint flexibility on fatigue life has been partially
studied by Recho which was on two very simple frames.
The reviewed studies show that consideration of joint stiffness makes significant changes in
the results of a structural analysis. The major effect has been recognised to be on the
bending moments. The fatigue life has hardly been attended in the studies related to
flexibility of tubular joints. The literature survey shows that the experimental investigations
on the various aspects of joint flexibility have not been carried out on frame specimens. This
is an essential step to verify and substantiate the results and conclusions already established
in this area.
The joint modelling techniques in the stiffness method use the spring concept most
extensively. The main task in the studies is finding the appropriate stiffness of the joints. In
this regard the rigidity factor is used to state the degree of fixity of a joint as a number
between 0 and 1.
Chapter 3
THEORETICAL MODEL
Chapter 3: Theoretical Model
42
3.1 Introduction
The theoretical part of this study considers the dynamic behaviour of beams in order to
establish a model for flexibility of joints. The stiffness matrix of a beam element with two
rotational springs at the ends is derived by exact solution of the differential equation of
displacement (in contrast with approximate or numerical solution). The axial flexibility is
also considered by applying translational springs to a joint. The formulation of axial
stiffness presented in Section 3.4 is derived by the author. The first difference between the
theoretical method incorporated here and those in other studies is the use of inertia forces
as a continuous property in the beam model. This would make the model more
representative of the real structures. Second is the dynamic approach employed in this
study for the calculation of joint stiffness which covers most stiffness and mass properties
of a joint. This is especially important for fatigue life estimation where inertia forces are
also involved.
Only prismatic beams are considered. Damping is not considered herein as this study deals
with steel structures and damping forces are not generally significant in this class of
structures, especially for lower natural frequencies. The stress-strain relationship is also
considered to be linear. These simplifying assumptions are employed herein to avoid a
complicated model. This study is carried out for the first time on tubular joints by the
author and different aspects of it is not known yet. It is, however, necessary to include
damping and especially non-linear material behaviour in the future works to achieve a
more refined model.
3.1.1 Beam theory
The relationship between flexural deformations and bending moments of a beam, is based
on the following assumptions:
1) planes normal to the axis of the beam remain plane and normal, after bending,
2) transverse displacements are attributed only to the longitudinal axis of the beam,
and
3) deformations are small.
The basic differential equation for flexural members then can be derived as:
EIy"=M (3.1)
Chapter 3: Theoretical Model
43
in which the curvature of the longitudinal axis of a beam, y", is related to the bending
moment M. In equation (3.1), _ is the modulus of elasticity and / is the moment of inertia
of the beam section about its neutral axis. In order to solve or analyse a structure that is
composed of beam members, called a framed structure, each individual member should,
within its domain, satisfy Equation (3.1). Also, continuity of displacements and rotations at
the boundaries of the beam members should be satisfied. This solution result in a system of
linear equations, where the unknowns are the integration constants.
Several techniques have been developed for construction and solution of such linear system
of equations as easily and practically as possible, such as: moment distribution method,
slope deflection method formed into matrix analysis procedures. Matrix analysis
formulation is most c o m m o n because of its suitability for computer programming.
Equation (3.1) states the most simplified version of the flexural behaviour of prismatic
beams. Effects of shear and/or axial deformations and rotary inertia on the bending
deformation can also be included which creates a more complicated equation. The matrix
method of analysis is similar for all cases with more elaborate algebraic manipulation
involved in establishing the beam stiffness matrix when other assumptions are introduced.
This study focuses on the beam formulation excluding axial deformation and rotary inertia
effects, as they are not usually considered in practice, yet.
3.2 Matrix formulation of beams
Rigid connections are commonly used in computer programs for analysis of framed struc-
tures. In the stiffness matrix of a beam each nodal point has three degrees of freedom in
plane, that is two translations and one rotation, as shown in Figure 3.1.
__C _}__.
k I
6, 5j
Figure 3.1. DOFsfor two dimensional beam element
The system of equilibrium equations in dynamic analysis of framed structures is usually
established by using two types of mass distribution models called discrete and continuous
models.
Chapter 3: Theoretical Model 44
Discrete model
A discrete mass model assumes that mass and stiffness are decoupled properties in
the equilibrium equations, as shown in Equation (3.2).
K x -
2
M x = P (3.2)
There are two types of discrete mass models known as lumped and consistent
models. Mass of the members are attached to the nodal points in the lumped mass
model and the beam members are assumed to be massless. However, a more accurate
representation of mass is obtained by expressing a uniformly distributed mass in
terms of the end deflections and slopes. This will result in a full matrix which is
called the consistent mass matrix. Equations (3.3) represent the lumped and
consistent mass matrices for a beam element.
c
E, 1, m, I, co
0
a) Degrees of freedom for the mass matrices
M
L
=
mill 0 0 0
0 ml/2 0 0
0 0 0 0
0 0 0 0
mL
M c
420
156
54
22/
-13/
54
156
13/
-22/
22/
13/
4/2
-3/2
-13/
-22/
-3/2
4/2
(3.3)
b) Lumped mass matrix
b) Consistent mass matrix
Continuous model
The stiffness matrix of a beam with continuous mass has been originally developed
by Kolousek in 1939 (Kolousek, 1973). He developed and published the dynamic
stiffness matrix of a beam using Dynamic Deformation method. The mass and
stiffness properties in the continuous model are coupled, therefore the stiffness
matrix includes the mass parameters and stiffness parameters together. A coefficient
of the stiffness matrix, &ij, is the force developed at the beam degree of freedom i due
to the amplitude of vibration one at the degree of freedom j. For example, the
coefficients of the third column of the stiffness matrix correspond to the deformations
and forces shown in Figure 3.2.
Chapter 3: Theoretical Model
45
k =
0 0 k
l3
0
0 0 k
23
0
0 0 k
33
0
L 0 0 k
43
0
Figure 3.2. Stiffness coefficients of beam with continuous mass due to the rotation of left end.
The common form of representing the stiffness matrix of a beam modelled with
continuous mass is shown by Equation (3.4).
FT FT FT FT
T
F
6
(X)
T
F
5
(X)
T
F
4
(X)
T
F
3
(X)
k =
FT FT FT FT
T
F
5
(X) jF
6
(X)
T
F
3
(k)
T
F<(X)
FT FT FT FT
T
F,{X) jF
3
(X)
T
F
2
(X) -pF^X)
L T
F
^ T
F
<
(X)
f
F
*
(X)
T
m)
(3.4)
In this equation the coefficients F
x
are called the Frequency Functions (Kolousek,
1973). A, is a variable which includes the stiffness properties and frequency of the
beam and is defined later in this section. E, /and / are Modulus of Elasticity, moment
of inertia and length of the beam.
3.2.1 Beam with semi-rigid ends (bending deformations only)
The model used in this study is a beam with two rotational springs h and k] at its ends, as
shown in Figure 3.3, where <g) represents a semi-rigid support.
end spring
(for this study)
Figure 3.3. Beam with semi-rigid ends (after Kohoutek 1991b)
The stiffness matrix of a beam with end springs was introduced by Monforton and W u
(1963). They used the conjugate beam method to derive a relationship between the forces
Chapter 3: Theoretical Model
46
and displacements at the ends of a beam member with semi-rigid connections. The elastic
behaviour of the springs was assumed to be linear as given by Equation (3.5).
M = kQ (3.5)
There was no reference to the inertia forces in the work by Monforton and Wu. Grant
(1968) used a similar model to Monforton and Wu's for dynamic analysis of frames but he
assumed a lumped mass formulation. The mass stiffness matrix that Grant used did not
include the spring stiffness of the ends. The stiffness matrix of a beam member with semi-
rigid ends, considering flexural deformations only and continuous mass, is derived in this
section. Equation (3.6) shows the differential equation that expresses the equilibrium
condition on an infinitesimal beam element under no external load.
d
4
v 5
2
v
EI^ + m^- = 0 (3.6)
obc
4
dt
2
v(x,t) is the transverse displacement of the beam axis, and m is the mass per unit length of
the beam. For a harmonic boundary displacement with the frequency of co, the solution
v(x,r) =y(x)sm(ot can be assumed that yields:
EI^4-m(o
2
y = 0 (3.7)
dx
A y
or
^-(=^ = 0 (3.8)
dx EI
Assuming X = l(mG>
2
/Er)
A
the solution of Equation 3.6 will be (Kohoutek, 1985b):
y(x) = Cicos(Xx/l) + C
2
sm(Xx/l) + Cje'^ + Qe'^^
7
(3.9)
in which / is the length of the beam and C/, C_, C3 and Q are the integration constants. For
the degrees of freedom shown in Figure 3.1, the boundary conditions of a beam with semi-
rigid joints will be:
8i = y(0) Sj=X/)
e
/ =
_MO)
+/(
o) o,=A/(0 (3.10)
Chapter 3: Theoretical Model 47
in which k\ and k$ are the stiffness of the ends i and; respectively. Equations (3.10) relate
the displacements and rotations of the ends / and j (see Figure 3.1) to the integration
constants in Equation (3.9) and may be shown in a matrix form as:
A = <DC (3.11)
in which A is the displacement vector. Matrix <b contains a number of terms related to the
beam geometry, stiffness and frequency, and matrix C includes the integration factors.
Expanded forms of the matrices used here are given in Appendix A. The end shear forces
and bending moments can be related to the displacement function as:
Vi = V(0) = _i_>""(0) Vj = -V(/) = -_.//'(/)
Mi = -M(0) = -_s_>"(0) Mj = M(r) = EIy'XT) (3.12)
when substituted from equation (3.9) yields:
P = _QC (3.13)
in which P is the force vector. Matrix Q contains a number of terms related to the beam
geometry, stiffness and frequency. Combining Equations (3.11) and (3.13), the stiffness
matrix of a beam with semi-rigid joints can be derived as:
P = __/Q<_r _ (3.14a)
.-. k = ii/Q<D-
1
(3.14b)
&ij, a member of matrix k is a function of modulus of elasticity (E), moment of inertia (i),
length (/), mass per unit length (m), frequency (co) and stiffness of the end springs (fa and/or
fcj). For example, k
33
is (see Figure 3.2):
33 =
EI
I
-A[(l-e
2X
)(cos>.-2aj>,sinA
<
)+sinl(l+e
2x
)j
2e
x
-X(ai+aj)(cosA,+siiiA)+A.e
2X
(ai+aj)(cosA,-sinA)+cOsA,(l+e
2X
)+2aiajA,
2
sinX(l-2e
2X
) ^ ' '
Chapter 3: Theoretical Model
48
in which cti = Ellfal and ctj = El/kjl. The complete stiffness matrix of a beam with semi-
rigid ends is given in Appendix A.
The relative stiffness of the beam with respect to the stiffness of the end springs is
represented by cti or ctj. For a fixed or a hinged support, a equals to zero or infinity,
respectively. The ideal fixed or hinged conditions, i.e. a support which produces y' = 0 or M
= 0, do not exist in reality. Therefore, a in practice has a value greater than zero.
3.2.2 Effect of shear deformations on bending stiffness
When the effect of shear deformations on bending stiffness is included, the differential
equation of displacement of the beam axis will be:
dx
4
Afi dx
1
EI
)y v
in which A
s
is the shear area of the beam section and G is the shear modulus of elasticity.
The modified boundary conditions for the rotational degrees of freedom are:
8,=j<0) &j=y(l)
e, = -^p-+y'(0)-y'
s
(0) o, =+^+/(0-^(0 (3.17)
in which y'
s
refers to the tangent of the shear displacement curve only. y'
s
can be obtained
from the shear force, V, as:
V
T7
_,_,d
3
y men
2
dy.
Rest of the calculations are similar to the case of bending deformation only, described
earlier.
3.2.3 Stiffness determination of a joint
Generally two methods can be used to determine the stiffness or rigidity of a joint via
dynamic measurements:
1) Forced vibration that is measurement of forces and displacements under a forced
vibration and then using equation (3.14a) to calculate the rigidity (Figure 3.4a).
2) Free vibration that is using the property of structures regarding natural
frequencies. The determinant of the stiffness matrix is zero at the resonant conditions.
Natural frequencies are the eigenvalues of the stiffness matrix of a structure and
Chapter 3: Theoretical Model
49
include the interaction of members meeting at a joint. Therefore, by knowing natural
frequencies, rigidity of a joint can be determined using Equation 3.19 below:
K x = 0 or Det[K] = 0 (3.19)
The procedures for determining the rigidity of a T-joint through dynamic testings are
shown in Figure 3.4.
P=P
0
Sin cot
5=5
0
Sin cot
Known parameters: frequency, geometry and material properties
Measured parameters: 5
0
and P
0
Unknown parameter to be calculated: rigidity by using: K x
0
= P
0
s\
(1) Forced vibration
^r
6= Sin ^
Unknown parameters: geometry and material properties
Measured parameter: frequency
Unknown parameter to be calculated: rigidity by using: K x
0
= 0
_____
(2) Free vibration
Figure 3.4. Dynamic methods to determine rigidity of a joint
W h e n using method (1) above, the displacement vector (A) is required in order to
determine the joint stiffness. A n accurate measurement of A is a difficult task in terms of
instrumentation and laboratory equipment. Method (2), shown in Figure 3.4, was used in
this study to evaluate the rigidity of the tubular joints. The advantage of this method is
employing the resonant conditions instead of using the force-displacement relationship of
Equation (3.13 a).
Forced vibration method was used by Springfield and Brunair (1989) on the end flexibility
determination of pre-tensioned cables. The main reason by Springfield and Brunair for
using a dynamic approach was that the real loading on the cables in their study had a
dynamic fashion and caused fatigue problem.
Chapter 3: Theoretical Model
50
There are only few researchers w h o have used dynamic approach in their analyses.
Kohoutek (1985a) introduced the method of free vibration and applied it to the joints made
of open sections. The present study is using the same experimental method as Kohoutek's
with a different theoretical model described in Section 3.2.
By the knowledge of this author, the dynamic experimental method and analytical model
are examined for the first time on the tubular in this study .
Other methods for flexibility analysis of joints employ the static approaches through using
the ratio of moment to rotation concept. A review of such investigations are reported in
Chapter T w o as a part of the literature survey.
3.3 Bending rigidity factor
The model developed in Section 3.2 for rigidity of a joint produces a rotational stiffness,
which theoretically ranges from zero to infinity. However, it is c o m m o n in practice to
express the degree of fixity of a joint by using a non-dimensional coefficient between 0 and
1. This coefficient is called rigidity factor and is 0 or 1 for a hinged or fixed joint,
respectively. The model of a joint in the present study is similar to the model used by
Monforton and W u (1963) in using a spring to express the joint stiffness, but the rigidity
factor defined here is slightly different.
To define a rigidity factor, this author has followed the pattern of moment rotation
relationship of a beam with end springs. Considering the beam member in Figure 3.5, 0j = 1
and rest of the D O F s are restrained.
'
33
5j=9j=5j=0
Figure 3.5. Beam model for definition of rigidity factor (unit length)
The relationship between the bending moment and the rotation of end / is:
Mi = k
33
Qi = k
33
(\) = k
33
(
3
-
2
)
Chapter 3: Theoretical Model 51
k
33
is the diagonal coefficient of the stiffness matrix of a beam with semi-rigid ends
corresponding to the rotational degree of freedom (see Appendix A ) , which is maximum
for a rigid and zero for a hinged end. The variation of k
33
with respect to the end flexibility
cti is shown in Figure 3.6 for different X values. X equals to zero represent the static case.
X = / (m(o
2
/EI)
1M
1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1111111
0 2 4 6 8 10
end flexibility, a j
Figure 3.6. Relationship between rotational stiffness and beam end flexibility
It is seen that a variation of X from 0 to 3.5 (X = 3.5 indicates a high level of dynamic
behaviour) does not have that much effect on the shape of the curves that state the
relationship between k
33
and ctj. A n equation for the rigidity factor v
b
is proposed here,
based on the relationships shown in Figure 3.6, as:
1
u
EI
v
k
= , where a = -rz l + 4a
kl
(3.21)
Equation (3.21) is also shown in Figure 3.6, which is actually a normalised form of k
33
when X = 0.
Chapter 3: Theoretical Model 52
3.4 Stiffness of a joint due to brace axial loading
Rotational flexibility is generally most important in the analysis of framed structures.
However, due to the radial flexibility of pipes in tubular structures (diameter to thickness
ratio effect), axial deformation of brace is also influenced by the flexibility of chord wall.
To consider the axial flexibility of a joint, an axially loaded member with two translational
end springs, as shown in Figure 3.7, is considered.
AAMAr ^
I
*t* *t* *"
i , J
0 / 0
Figure 3.7. Axially loaded member with semi-rigid ends
The differential equation of equilibrium for the axially loaded member of Figure 3.7 under
no external load is:
EA
^-m^ = 0 (3.22)
dx
2
dt
2
in which h(x,i) is the function that describes the axial deformation, E is the modulus of
elasticity and A is the cross sectional area. A solution as h(x,t) = u(x)sm((at) may be
assumed which is a harmonic boundary displacement with a frequency of co. Therefore,
Equation (3.22) can be written as:
EA^- + ma
2
u = 0 (3.23)
dx
2
or,
^
+ (
_ ! _ _ _ 1 )
W =
0 (3.24)
dx
2
" EA
Assuming \j/ = l(m(o
2
/EA)
,/2
the solution of Equation (3.23) will be
u(x) = C5sin(\|/x//) + C.cos(\|/x//) (3.25)
Chapter 3: Theoretical Model
53
For the degrees of freedom shown below, the boundary conditions of an axially loaded
member with semi-rigid ends are given by Equation (3.26).
Ii "Hj
P
J
T l , = - ^ + (0)
1
^=+-^r+(0 (3-26)
fc'i and 'j are the axial stiffness of the ends / and j respectively, P(x) is the internal axial
force of the member. Substituting from Equation (3.25) in (3.26), it can be written in a
matrix form similar to Equation (3.10) that:
A_ = O
fl
C
a
(3.27)
in which A_ is the displacement vector. Matrix <E>_ contains a number of terms related to the
geometry, axial stiffness and frequency, and matrix C
fl
includes the integration factors, Cs
and C_. Expanded forms of the matrices used here are given in Appendix B. The axial
forces at the end of an axially loaded member can be written as:
Pi = -P(0) = -EAu'(0) P
]
= P{l) = EAu\l) (3.28)
when substituted from equation (3.25) in (3.28):
P
fl
= i_Q_C_ (3.29)
in which P_ is the force vector. Matrix Q
fl
contains a number of terms related to the
member geometry, stiffness and frequency. The stiffness matrix of an axially loaded
member with semi-rigid ends can be obtained by combining Equations (3.29) and (3.27),
resulting:
k
a
= EAQ
a
O
a
-
1
(3.30)
fcj, a member of matrix k_ is a function of modulus of elasticity (E), sectional area (A),
length (/), mass per unit length (m), frequency (co) and stiffness of the end springs (k\
and/or kj). For example k
a
is:
Chapter 3: Theoretical Model 54
k
a
=
EA \j/(cosi|/ - Pjsinu/)
/ i|/cos\j/(pi+pj) + sinvi/(l-pipjV|/
2
)
(3.31)
in which pi = EAIk'i and PJ = EA/k)l. The complete stiffness matrix of an axially loaded
member with semi-rigid ends is given in Appendix B. p values represent the relative axial
stiffness of the member and the end springs. P = 0 and P = oo are limit values for rigid and
sliding supports, respectively.
Axial stiffness of a joint can be determined similarly as for the bending case (Figure 3.4)
by conducting forced or free vibration tests. In the axial tests the force P or, in case of free
vibration, the excitations are acting in the axial direction of the brace.
3.4.1 Axial rigidity factor
A rigidity factor can be defined to show the degree of fixity of a joint under axial
deformations. The factor maps the joint stiffness to a range from 0 to 1. In this study, the
relationship between axial force and displacement at the end of an axially loaded member
is used to establish the function that converts the joint stiffness to the rigidity factor. This is
in fact the diagonal coefficient of the stiffness matrix of an axially loaded member. Figure
3.8 shows the variation of &," with respect to pi (= EA/k[l) for a member fixed at nodal
pointy.
v
f
/ = 0
'
( V f l =
l
+
p7
)
F! = k(1) = k?i
\y=l(m
2
/EA}
/2
J K
axial end flexibility, p,
Figure 3.8. Relationship between axial stiffness and axial end flexibility
Chapter 3: Theoretical Model 55
Accordingly, the following equation for the axial rigidity factor is suggested:
v.^. where p-ff (3.32)
3.5 Summary
The dynamic stiffness matrix of a beam with semi-rigid supports was derived in this
chapter considering continuous mass and stiffness (ET) properties. The beam deformations
which were assumed included: 1) only bending properties, and 2) bending plus shear
properties. A coefficient of the stiffness matrix is a function of modulus of elasticity (E),
moment of inertia (7), length (/), mass per unit length (w), frequency (to) and stiffness of
the end springs (fa and/or kj).
The theoretical model used in this thesis for determination of the semi-rigidity of a joint
employs the resonant condition and natural frequencies. This approach is examined for the
first time on the tubular joints in this thesis. A new definition for rigidity factor, which uses
the bending behaviour of a beam, was given to produce a measurement number for rigidity
between 0 and 1.
The dynamic stiffness matrix of an axially loaded member with semi-rigid ends was also
derived in this chapter. The stiffness matrix include both stiffness and continuous mass
properties. For the axial case, a rigidity factor which varies between 0 and 1 is also
suggested.
Chapter 4
EXPERIMENTAL DETERMINATION OF
JOINT STIFFNESS
Chapter 4: Experimental Determination of Joint Stiffness
57
4.1 Introduction
A dynamic method of testing is used in this study for evaluating the rigidity of tubular joints.
The word dynamic is used here because inertia forces are involved in the applied experi-
mental procedures. The method has been developed and previously used by Kohoutek
(1985a) for inplane rigidity of open section joints. In this study, the inplane and out of plane
flexibility of tubular T-joints are investigated experimentally. The theoretical base used to
determine joint stiffness from the experimental results is described in Chapter 3.
The dynamic method of analysis employed in this study to determine the stiffness of tubular
T-joints, is not as cumbersome as the c o m m o n procedures where load and displacements
are measured. Furthermore, the measurements of natural frequencies on steel specimens are
easier and less complicated compared to the static measurements of displacements.
4.2 Test specimens
Eleven tubular T-joints were tested with the geometrical dimensions and non-dimensional
parameters reported in Table 4.1. The specimens were fabricated from steel pipes readily
available in Australia. Subsequently, from a limited choice of diameter and thickness
dimensions available selection was made for the specimens. A survey of the non-dimensional
parameters P (= dID), y (= D/2T) and x (= tIT) was made to ascertain the apphcability of the
test pieces to real structures. The results of this survey are shown in the next section.
Table 4.1. Geometrical dimensions and non-dimensional parameters of tested joints
Joint
Tl
T2
T3
T4
T5
T6
T7
T8
T9
T10
Til
D(mm)
219.1
219.1
219.1
219.1
219.1
219.1
219.1
219.1
219.1
219.1
219.1
T(mm)
8.2
8.2
8.2
8.2
6.4
6.4
6.4
6.4
4.8
4.8
4.8
Um)
1.50
1.50
1.50
1.50
1.50
1.50
1.50
1.50
1.50
1.50
1.50
d(mm)
219.1
114.3
114.3
60.3
219.1
114.3
114.3
60.3
219.1
114.3
60.3
t(mm)
4.0
6.0
4.8
3.9
4.8
6.0
4.8
3.9
4.8
4.8
3.9
Km)
1.23
1.23
1.23
1.23
1.23
1.23
1.23
1.23
1.23
1.23
1.23
P
1.00
0.52
0.52
0.28
1.00
0.52
0.52
0.28
1.00
0.52
0.28
Y
13.36
13.36
13.36
13.36
17.12
17.12
17.12
17.12
22.82
22.82
22.82
I
0.49
0.73
0.59
0.48
0.75
0.94
0.75
0.61
1.00
1.00
0.81
a
13.69
13.69
13.69
13.69
13.69
13.69
13.69
13.69
13.69
13.69
13.69
See Figure 1.5 for the notations used in this table.
Chapter 4: Experimental Determination of Joint Stiffness 58
The selected values for P and y cover the limits of the respective parameters in the offshore
structures plus one value approximately in the mid-range. The chord length parameter, a,
was kept 13.69 in all specimens. Strictly speaking, the length of the chord should be
selected according to the condition of a member in a real structure. However, other studies
show that a has a minor influence on the flexibility of a T-joint, which is also indicated by
the flexibility and SCF parametric formulae, not including any a parameter. The effect of a
on joint flexibility is discussed in Section 6.2. The reader is referred to Chapter 6 and
Appendix G, for flexibility and SCF parametric formulae, respectively.
4.2.1 Survey of non-dimensional parameters of T-joints (P, y and x)
The test specimens in this study were restricted to T-joints and therefore only the survey of
those joints will be shown here. To choose the practical ranges for the non-dimensional
parameters of the joints, two surveys were examined.
First, the survey reported in U E G (1985) for six different structures in the North Sea, which
were believed to include a common ranges of variables. The total number of intersections
were about 2700 for p population, and 1602 for y parameter. In this set, there was
insufficient information available to identify x ratio. The number of the joints corresponding
to P and y ranges are shown in Figure 4.1 and 4.2.
Figure 4.1. Distribution of P parameter for T- Figure 4.2. Distribution ofy parameter for T-joints.
joints. The survey was carried out by UEG (1985) The survey was carried out by UEG (1985) on six
on six offshore structures offshore structures
Second, was the database also referenced in U E G (1985). The results for p, y and x are
shown in Figures 4.3, 4.4, and 4.5. The non-dimensional parameters of the tested joints in
this thesis vary as follows:
0.28 <P< 1.0
13.36 < Y < 22.82 (4-1)
0.49 < x < 1.0
Chapter 4: Experimental Determination of Joint Stiffness
59
150-
140
26 36 46
Y
Figure 4.3. Distribution of y parameter for T-
joints. The survey was carried out on the data-
base in UEG (1985)
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Figure 4.4. Distribution of $ parameter for T-joints. The
survey was carried out on the database in UEG (1985)
25 -
20 ..
1
5.
o
15 4-
10 ..
5 ..
21
6 6
| K S _ | KSSSI
t
KSSS11
14
l ^ t ^ l ^
<0.15 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 >0.95
T
Figure 4.5. Distribution of % parameter for T-joints. The survey was carried out on the database in UEG
(1985)
As can be seen, the most common values of y in offshore structures are in the range of 10 to
20, which indicates the range chosen for the specimens. A range of 0.15 to 1.0 was found
for P with most joints having P values between 0.4 and 0.6. The survey also shows the
following limits for the non-dimensional parameters of T-joints:
0.2 < p < 1.0
5 < Y ^ 35
0.2 < x < 1.0
(4.2)
4.2.2 Material properties and fabrication of test specimens
The steel pipes used for the fabrication of the test specimens were seamed and had the
properties shown in Table 4.2. The pipes were m a d e of mild steel and manufactured in
Australia by B H P Company.
Chapter 4: Experimental Determination of Joint Stiffness
60
Table 4.2. Metallurgical properties of the pipe sections used for test specimens
Element
C
P
Mn
Si
S
Al
CE
Average%
0.14
0.023
0.74
0.12
0.007
0.022
0.32
M i n i m u m %
0.13
0.014
0.65
0.10
0.003
0.014
0.29
Maximum%
0.16
0.029
0.80
0.13
0.012
0.040
0.35
Brace sections were welded to the chords using Complete Joint Penetration Groove Weld.
Details of the weld used for the fabrication of the specimens are shown in Figure 4.6.
1) Weld size was 6mm. There were 3
passes on each joint.
3rd and last pass
2) Preparation on adjoining brace was a
45 bevel, feathered edge with 1.6mm gap
(penetration gap) for full penetration root
run.
3) Weld passes 1 and 2 were ground out
before proceeding.
Ground out edges
Figure 4.6. A typical weld cross section of a T-joint specimen
The Modulus of Elasticity (E) was confirmed experimentally to be between 200 and 210
GPa. Therefore, the common value of E = 200 GPa was used through this study in
calculations.
4.3 Test set-up
The test set-up used to determine joint flexibility is shown diagrammatically in Figure 4.7.
The main equipment for testings was as follows:
-Fourier Analyser, 2630 Tektronix,
-Personal Computer,
-accelerometer, and
-charge amplifier.
Chapter 4: Experimental Determination of Joint Stiffness
61
Chord
fWRffJ
Accelerometer
Brace
_
Charge amplifier
\
Spectrum analyzer
__
Figure 4.7. Diagrammatic test set-up for frequency measurement of T-joints (IPB)
The vibration induced by impacting the cantilever, was picked up by an accelerometer
mounted on the desired spot (usually tip of the cantilever on a joint). The output signal from
the accelerometer was amplified and then processed by the Fourier Analyser.
The Fourier Analyser, Tektronix 2630, digitised and calculated the frequency components
of the input signal. It was geared to a Personal Computer for its software needs. In addition,
the P C was used for recording the measured data generated by the Analyser.
Figure 4.8a shows one of the T-joints supported in place for testing. The support assembly
is shown in Figure 4.8b.
Figure 4.8a) T5, one of the joint specimens on the support assembly
Chapter 4: Experimental Determination of Joint Stiffness
62
Figure 4.8b) Hinged support assembly
4.3.1 Test procedure
The natural frequency of a joint corresponding to the cantilever mode of vibration was
measured in the tests. The joints were excited by a simple impact load through a light
hammer or hand and the frequency spectra were calculated by the spectrum analyser and
recorded on disk. Since the desired frequencies were lower than 500Hz, using hand or
hammer with a rubber tip seemed to be adequate. A typical frequency spectrum of a T-joint
is shown in Figure 4.9 with the peaks as the natural frequencies. It is necessary to verify the
recorded peaks to make sure that they are related to the desired mode shapes. This is
important to avoid any noise and interference from other sources.
-90
dB Volts 5Q/Div
-10
10
/Div
- - - - ) - - - r - - - ( - - - r - - - l - - - T - - - | - - - T
- - -i- - J A - - -i- * - * - M -i- - - * Jl - -1 - - - *
A -_^ _ _ k - _ -.-_-. _ _\_ _ _* . A - - - J
- Hertz-
500
Ch=1/ ASPECa f=95Hz
Figure 4.9. Typical frequency spectrum of a tubular T-joint
Chapter 4: Experimental Determination of Joint Stiffness
63
Three methods were examined to confirm the correspondence between natural frequency of
a joint and the related mode shape. The methods are:
1) Eigenvalue analysis by Finite Element method using one dimensional beam elements.
This would clarify the mode shapes which should be expected from the test samples and
give some general idea about the dynamic behaviour of the joints. Of course, using more
complicated finite elements such as plate and shell or three dimensional element will
produce more realistic results than using one dimensional beam elements. However, for
mode shape identification of a joint, simple methods of analysis still can be sufficient The
application of the F E M to the flexibility analysis of tubular joints will be discussed further in
Chapter 5.
A series of analyses were carried out to identify the mode shapes of the joints using ALGOR
computer program. This program assumes a lumped mass formulation for the inertia forces
in dynamic analysis. A typical joint model is shown in Figure 4.10, where the springs, at the
supports, are a result of the provisions used to simulate hinged conditions and will be
explained in Section 4.5.1.
lumped mass f
|-VWv\_-^-^-A-^^
s
vWv^|
^* S77777 /t///9 ^
Figure 4.10. Lumped mass model used to establish the mode shapes of joints
The first two mode shapes due to the inplane bending mode of the specimens are calculated
by using ALGOR and shown in Table 4.3. The mode shapes of the joints are also reported in
Appendix C. Obviously, the connection of the chord and brace in the ALGOR analyses were
assumed to be rigid.
2) Eigenvalue Analysis using Dynamic Deformation Method. This method of analysis is
similar to method 1 above. The difference is only due to the mass formulation of the beam
elements. Method 2 assumes a distributed mass properties which is an improvement on the
assumption of lumped mass used in ALGOR program in Figure 4.10. First two natural
frequencies of the T-joints calculated by using method 2 for in plane bending mode are
Chapter 4: Experimental Determination of Joint Stiffness 64
shown in Table 4.4. T o make the analyses consistent with method 1, the connection of
chord and brace in this series of analyses were also assumed to be rigid.
Table 4.3. Natural frequencies of the T-joints using lumped mass theory (IPB)
Joint
T l
T2
T3
T4
T5
T6
T7
T8
T9
TIO
Til
hinged support
1st mode
7.74
8.03
8.21
8.73
8.13
8.77
9.01
9.70
8.82
9.98
10.94
2nd mode
140.88
74.96
75.03
37.93
140.20
75.89
75.86
38.23
139.75
76.99
38.67
fixed support
1st mode
7.74
8.03
8.21
8.73
8.13
8.77
9.01
9.70
8.82
9.98
10.94
2nd mode
143.98
75.31
75.32
37.95
144.41
66.33
76.22
38.25
144.86
77.47
38.70
Table 4.4. Natural frequencies of the T-joints using Dynamic Deformation Method (IPB)
Joint
T l
T2
T3
T4
T5
T6
T7
T8
T9
T10
Til
hinged support
1st mode
7.74
8.0
8.22
8.73
8.13
8.76
9.01
9.70
8.82
9.98
10.93
2nd mode
143.09
76.27
76.37
38.62
142.49
77.22
77.20
38.92
141.96
78.35
39.37
fixed support
1st mode
7.74
8.03
8.22
8.73
8.13
8.76
9.01
9.70
8.82
9.98
10.93
2nd mode
146.28
76.63
76.67
38.65
146.82
77.67
77.58
38.94
147.22
78.83
39.40
3) Modal Analysis. This is an experimental procedure used to characterise the dynamic
properties of an elastic structure in terms of its modes of vibration (SMS, 1990). A mode of
vibration is a global property of a structure and is defined by a specific natural frequency
and a mode shape. The measurements for modal analysis are obtained by exciting the
structure and measuring its responses at various points across its surface. The
measurements in this study were acquired by using a Fourier Analyser. The different natural
frequencies and mode shapes were then processed using a program called STAR. Figure
4.11 shows the first two mode shapes of joint T10 calculated through modal analysis.
Chapter 4: Experimental Determination of Joint Stiffness
65
Frequency: 11.59Hz
Frequency: 46.71 Hz
Mode#: 1
Mode#: 2
Figure 4.11. First two mode shapes of T-joint TIO from Dynamic Modal Analysis
Modal analysis can reveal any gross imperfection which may be present in the specimen
assembly and its supporting system. It is more realistic than methods 1 and 2, as it includes
direct measurements from a joint specimen. However the cost of modal analysis is higher
than the other two methods and in some cases it becomes very tedious such as in tubular
structures or generally in shell structures. Hence, method 3 may be used for identification of
the dynamic behaviour of a structure wherever a high accuracy is required or when the
computer modelling techniques do not work accurately. Obviously, only methods 1 and 2
can be used in the design stage before a structure is fabricated.
After the relevant natural frequency to the cantilever mode of bending is determined, rigidity
or flexibility of the joint can be calculated using the theoretical model. Similar procedures
were used for out of plane bending modes.
4.3.2 Test results and discussion
The measured natural frequency of the T-joints due to the cantilever mode of vibration are
reported in Table 4.5. This table also shows the calculated results by Dynamic Deformation
method ( D D M ) and ALGOR analyses. It is seen that there is a distinct difference between
the measured natural frequencies of the joints and those frequencies resulted from the
stiffness method analysis, using one dimensional beam element and rigid joints. It is also
seen that the natural frequencies resulted from lumped mass model (ALGOR) are less than
the values from Dynamic Deformation Method and closer to the measured results. This is
because the resultant inertia forces from lumped masses produce bigger deformations than
the distributed inertia forces in D D M model. This situation is similar to the different
maximum static moments under distributed and concentrated loads in the framed structures.
Chapter 4: Experimental Determination of Joint Stiffness
66
Table 4.5. Natural frequency of the T-joints due to cantilever mode of vibration
Joint
Tl
T2
T3
T4
T5
T6
T7
T8
T9
TIO
Til
Inplane bending
Measured
95
55
56.3
33.8
88.8
50
51.3
31.3
82.5
46.3
28.8
DDMt
143.09
76.27
76.37
38.62
142.49
77.22
77.20
38.92
141.96
78.35
39.37
ALGOR*
140.88
74.96
75.03
37.93
140.20
75.89
75.86
38.23
139.75
76.99
38.67
Out
Measured
46.1
30.8
33.4
27.3
36.8
24.4
26.5
23.1
33.3
20.1
17.9
of plane bending
DDMt
104.95
66.52
68.06
37.07
94.61
65.41
67.09
37.00
87.33
65.70
36.90
ALGOR*
103.72
65.72
67.23
36.60
93.54
64.63
66.28
36.54
86.51
64.92
36.44
fDDM: Dynamic Deformation Method (bending only).
XALGOR: Finite Element Package (lumped mass model).
Figure 4.12 shows the relationship between natural frequency of the joint specimen obtained
from measurements and Dynamic Deformation analyses. The difference seen between the
measured and calculated natural frequencies is the basis for determination of the rigidity of a
joint. In theory, a joint may be assumed rigid when its calculated natural frequency, using
Dynamic Deformation method, is equal to its measured natural frequency. Obviously, due to
the flexibility of tube walls there is no such joint in reality.
150-
140-
130-
_ - -
g 110-
S loo-
s' 90-
^ 80-
70-
3 60-
36
50-
40-
30-
20
measurement
- - - calculation, Dynamic
Deformation method
-D- -^13.36
-_- r="12
-0 fll.il
a f=13M
it f\l.\l *
0- 1*22.82
0
i 1 1 1 1 | 1 1 1 1 1 | i | i |
0_> 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Diameter ratio, P
Figure 4.12. Relationship between natural frequencies of specimens and diameter ratio
It is noted in Figure 4.12 that the joints which have large diameter ratios (fJ) or small chord
to thickness ratios (y) show the highest frequencies. This indicates the higher stiffness of
these types of joints which will be also confirmed in Chapter 6 when the stiffness of the
joints are calculated.
Chapter 4: Experimental Determination of Joint Stiffness
67
4.4 Calculation of rigidity factor and joint stiffness
Determination of the rigidity factor (v_) for a joint is, in principle, carried out using the
following equation:
K x = 0 (4.3)
in which K is the stiffness matrix of the joint and x is the displacement vector. Equation
(4.3) represents an eigenvalue problem, which can be used to detennine the natural frequen-
cies of a structure. The non-trivial solution for the displacement vector x, is obtained when
the determinant of the stiffness matrix K becomes zero. The natural frequencies are the
roots of the resultant equation.
The method explained in Chapter 3 is used here for the calculation of rigidity factor (v*,) of
a joint. This method uses a continuous mass property to calculate the dynamic stiffness
matrix.
Following the concept of zero determinant for the stiffness matrix, rigidity factor can be
determined when the natural frequency is known. Elements of the stiffness matrix K in
Equation (4.3) are functions of E (modulus of elasticity), / (moment of inertia), / Qength),
v
ft
(rigidity factor), m (mass per unit length) and co (frequency). After measuring the natural
frequency, Equation (4.3) can be solved for v
b
. Furthermore, the spring stiffness of the
brace connection to the chord in a T-joint can be calculated using Equation (4.4) shown
below:
1 , , 42*/ v,
v = T-FT, a n d * =
v
" , AEI
1+
kl
L l-v
(4.4)
The flow chart for the calculation of rigidity factor and finally stiffness of a T-joint is shown
in Figure 4.13.
Natural frequency
measurement
(Fourier Analyser)
Kx = 0
*: =
4EI v
b
/ l-v
Figure 4.13. Stiffness calculation of a T-joint using measured natural frequency
Chapter 4: Experimental Determination of Joint Stiffness
68
The analytical model which suggests the flow chart of Figure 4.13 is a joint constructed
with one dimensional beam elements (line elements) as shown in Figure 4.14. The
connection between brace and chord is assumed to be semi-rigid with the spring stiffness k.
The supports were designed to simulate the ideal hinge condition for inplane bending as
closely as possible, but in the analytical model a semi-rigid connection can also be assumed,
provided the rigidities are known. However, as it will be shown in the next section, chord
supports can be safely assumed as hinged for inplane bending mode.
A similar model was used for out of plane bending mode. The supports, though, were
assumed to be rotationally fixed.
_
77
y
r^
/7777
Figure 4.14. Analytical model used for rigidity calculation of tested T-joints (IPB)
Calculated rigidity factors (v_), stiffness and related natural frequencies of the tested joints
for inplane and out of plane bendings are reported in Table 4.6. Rigidity and joint stiffness
have been calculated according to the flow chart in Figure 4.13.
Table 4.6. Measured na
Joint
Tl
T2
T3
T4
T5
T6
T7
T8
T9
TIO
Til
tural frequencies (<>), rigidity factor (v^), and stiffhe
IPB
CO (Hz)
95
55
56.3
33.8
88.8
50
51.3
31.3
82.5
46.3
28.8
v
b
0.389
0.504
0.530
0.763
0.318
0.399
0.426
0.642
0.256
0.329
0.526
*(kNm)
6454
1984
1825
576
5620
1298
1197
323
4163
789
200
ss (k) of the tested T-joints
OPB
co(Hz)
46.13
30.75
33.38
27.25
36.75
24.38
26.50
23.13
33.25
20.13
17.88
v
b
0.117
0.191
0.226
0.540
0.069
0.117
0.141
0.390
0.061
0.077
0.235
*(kNm)
1346
462
470
211
895
258
266
115
787
134
55
Chapter 4: Experimental Determination of Joint Stiffness
69
4.5 Effect of joint support conditions
Inplane bending
The effect of chord support conditions on the natural frequency was examined for inplane
bending m o d e using the following two approaches:
1) T w o models of the joints with the limit states for their boundaries, i.e. hinged and fixed,
were analysed for the first two natural frequencies using the Dynamic Deformation method.
In these analyses the joints were assumed to be rigid. The variation of the natural
frequencies of a joint when its support conditions vary from hinged to fixed will indicate
h o w much the structure is sensitive to the boundary conditions. The results of the analyses
are reported in Tables 4.3 and 4.4. It is noted that the effect of support fixity of a chord can
be as high as 3.7% on the IPB natural frequency of the joint when compared to a hinged
support.
2) Using results of the tests on the joints that have the hinged supports described in Section
4.5.1, the rigidity factors and stiffness of the joints were calculated based on two analytical
models described above (one supported on hinged and the other on fixed supports). The
results reported in Table 4.7, show that the maximum modelling error can be 4 % . This
indicates that the change of the support conditions of the analytical model from one extreme
to the other, has effectively no influence on the calculated stiffness of the joint
Table 4.7. Joint stiffness calculated for hinged and fixed analytical models using test results of
specimens that have hinged supports
Joint
Tl
T2
T3
T4
T5
T6
T7
T8
T9
TIO
Til
v
(supports
0.389
0.504
0.530
0.763
0.318
0.470
0.399
0.642
0.256
0.329
0.526
fc(kNm)
are hinged)
6454
1984
1825
576
5620
1298
1197
323
4163
789
200
v_
(supports
0.381
0.502
0.528
0.763
0.310
0.398
0.424
0.642
0.250
0.327
0.525
fc(kNm)
are fixed)
6248
1968
1808
576
5428
1288
1186
323
4029
783
199
Khinged/Kfixed
1.03
1.01
1.01
1.00
1.04
1.01
1.01
1.00
1.03
1.01
1.00
According to the results of the analyses presented in this section, chord support conditions
have a negligible effect on the calculated IPB rigidity. This will be also shown in Section
5.5.1 when Finite Element analysis of the tested joints are described.
Chapter 4: Experimental Determination of Joint Stiffness
70
Out of Plane bending
In out of plane bending mode the chord of a joint is subjected to torsional moments.
Therefore, a fixed conditions at the supports were required for out of plane bending mode,
otherwise instability problems would have occurred.
4.5.1 Provisions of end supports for test specimens
Inplane Bending
Although the boundary conditions of a chord have a very small influence on the calculated
rigidity and stiffness of a T-joint for inplane bending mode, the best arrangement was made
to simulate a hinged condition for the supports of the test specimens. For inplane bending,
the support arrangement provided for the joints is illustrated in Figure 4.15.
The support conditions shown in Figure 4.15 were used for both ends of a chord. The
reeds, which were made from thin spring steel, could induce effectively no bending
resistance to the joint specimen. The reeds were glued in place to reduce the introduction of
any vibrational noise. The stirrups around the chord held it firmly along the perimeter
causing a uniform radial pressure on the chord wall. Because all the components in the
support assembly were held in contact with each other, vibrational noise was minimal and a
clear frequency spectrum could be obtained.
Figure 4.15. Support arrangement for inplane bending test
Chapter 4: Experimental Determination of Joint Stiffness
71
Out of plane bending
A fixed support condition was tried to be made in out of plane bending tests. The
arrangement for the supports is shown in Figure 4.16. There is a solid stirrup at each end of
the chord holding it in place. The reaction forces at the supports are torsional moments and
are provided by the frictional forces between the stirrups and the chord wall. If there was
any looseness between the stirrups and chord walls, it would have produced instability in the
test specimen and could be easily detected. Furthermore, the tests were conducted with a
low amplitude of vibration which did not produce large forces at the supports, so the
support conditions could be assumed to remain fixed when a joint specimen was vibrating.
The whole support assembly was bolted to the floor using high tensile strength bolts.
It will be seen in Chapter 5, though, that the integration which bolting has created
apparently has not been sufficient to produce vibrational fixity for the joint specimens with
large P ratios. This problem requires further investigation that can be pursued as the future
research work of the thesis.
Figure 4.16. Support arrangement for out of plane bending test
Chapter 4: Experimental Determination of Joint Stiffness
72
4.6 S u m m a r y
The experimental part of this study includes testing of eleven T-joints. The geometrical
specifications of the specimens are selected to cover the range of parameters commonly
used in offshore structures. To determine the range of parameters, a survey was carried out
on the non-dimensional parameters of a number of offshore structures. The database for the
survey was taken from U E G (1985). The survey indicated the highest population of joint
parameters: P, y, and x to be 0.5, 15, and 0.5 respectively.
The pipes used for the fabrication of the test specimens were seamed and made of mild
steel. The primary measurements in the tests were natural frequencies of the specimens
which was determined by a Fourier Analyser. A free vibration procedure as described in
Section 3.2.3 was employed in the experiments. The excitations were applied by hand or
through a hammer with a rubber tip. T w o modes of deformation i.e. inplane bending and
out of plane bending were considered. Computer modelling and modal analysis were used
to ascertain the correspondence between a measured natural frequency and the related
mode shape. A clear difference was seen between the measured natural frequencies and
those obtained from numerical modelling assuming a rigid joint. For example joint T9 had
a free vibration with a frequency of 83Hz for in plane bending whereas numerical analysis
showed 140Hz. This difference between measured and computed frequencies was the basis
for the calculation of joint stiffness.
The stiffness values of the tested joints were calculated based on the method presented in
Sections 3.3 and 4.4. It was noted that the joints with large diameter ratios or small chord
to thickness ratios possess higher stiffness values compared to other ranges of parameters,
both for inplane and out of plane bending modes. This is also discussed in Chapter 6 when
the stiffness of T-joints are calculated using F E Method.
Effect of support conditions on the natural frequencies of a joint was investigated for
inplane bending mode. The results show a maximum difference of 4 % in the calculated
stiffness of a T-joint when the assumed conditions for its supports are changed from hinged
to fixed which shows the insignificance of support fixity for IPB mode. Despite the small
effect of support conditions on the inplane bending behaviour of T-joints, a special support
arrangement was made to simulate the true hinged conditions.
Out of plane bending requires fixed supports as the chord of a joint is under torsional
loading. However, it is indeed difficult to build rotationally fixed support conditions,
especially when the structure experiences dynamic forces. This is clearly observed in the
difference between the F E analysis and the experimental results of the joint specimens in
IPB and O P B modes and presented in Section 5.6.1.
Chapter 5
DETERMINATION OF JOINT RIGIDITY
USING FINITE ELEMENT METHOD
Chapter 5: Determination of Joint Rigidity Using Finite Element Method
74
5.1 Introduction
The Finite Element Method (FEM) is the only analytical procedure that is widely used to
analyse the structural behaviour of tubular joints. The ability of the F E M to model arbitrary
boundary conditions is a significant advantage over other methods used to solve different
types of structural problems. However, it is still an approximate solution and requires to be
verified by other methods that are mainly experimental.
This chapter gives a brief introduction on the F E method. Further, the results of the FE
analyses of the test specimens of Chapter 4 is reported and discussed.
5.2 A review of the Finite Element method
The behaviour of most structural systems may be formulated by using the differential
equation that expresses the interaction between the system components. The basic
differential equation governing the behaviour of a three dimensional solid structure made of
an elastic material can be determined by imposing the equihbrium conditions on a cube with
the sides dx, dy, dz, yielding the following set of equations:
_____.
+
__!_L
+
__!__.
+
p__l
u
dx By dz dt
*-_ Cri o _ ,, ~
d<T
v
dx dx
vx
d
2
u
v
_,
n
,_ ..
*r
+
lf
+
lf
+p
#
+r=0 (51)
^ _
+
_ I _
+
_^_.
+
p|^
+
Z = 0
dz dx dy dt
2
o, o
y
,..., %xz are the stress components due to the body and surface loads. The body forces
per unit volume are denoted by X, Y, Z in Equation (5.1). If the surface forces per unit area
are assumed as X\ Y and Z', the equilibrium conditions at the surface can be represented by
Equations (5.2) below:
X' = Gxl + Xxym+ Xxtfl
Y' = Gym + Tyzn+ Xxy/ (5.2)
Z' = o_n + X_+ Tyz/n
Chapter 5: Determination of Joint Rigidity Using Finite Element Method
75
I, m and n are the direction cosines of the external normal vectors at the point under
consideration. The six components of strain at each point can be calculated by using the
three displacement functions u, v and w. Hence, the components of strain can not be taken
arbitrarily as functions of x, y and z and they are interrelated. These interrelations are known
as the equations of compatibility or Saint-Venant's compatibility equations. The stress
solution of an elastic system must satisfy all conditions of internal and external equilibrium
plus the compatibility conditions.
The Finite Element Method employs the above conditions in terms of displacement
components, aiming to obtain the best approximation to the solution of u, v and w. This is
achieved by dividing the elastic body into smaller segments (finite elements) and
approximating the displacement fields in each element with an arbitrary function. The finite
elements are interconnected via nodal points. There are many numerical methods to
approximate a function, but the c o m m o n practice is assuming polynomial functions, called
shape functions or interpolation functions, with unknown coefficients. The equilibrium
conditions described by Equations (5.1) are imposed on each element, allowing the
evaluation of the unknown coefficients introduced by the shape functions. Generally, the
method is formulated so that the unknown coefficients are displacements or rotations at the
nodal points for a given load. After determination of the displacement functions for each
element, strains are calculated and then stress components can be obtained using Hook's
law.
The F E M is more sophisticated than as explained above with various numerical problems
ranging from computational stability of finite mathematics to the convergence of the
solutions. However, the method is generally very applicable to the stress analysis of
structural systems.
Another method for stress analysis of structures is the Finite Difference Method. This
method is not as popular as the F E M but is still used in practice. The difference between
Finite Element and Finite Difference method in the solution of a differential equation is that
the latter approximates the desired function as a whole, while the former attempts to satisfy
the differential equation for each individual element, while mamtaining the continuity
conditions.
5.2.1 Galerkin Method (a finite element formulation)
In this section a version of the Finite Element formulation, which is called the Galerkin
method, is described. Assuming a differential equation as:
L(M)+P = 0 (5-3)
with L ( M ) as a differential operator, a solution u can be assumed such as:
Chapter 5: Determination of Joint Rigidity Using Finite Element Method
76
M
" (5.4)
where TV, are linearly independent trial functions or interpolation functions and a, are
M
multipliers to be determined in the solution. u
b
satisfies the boundary conditions and X .A.
=i
is zero at the boundaries. The unknowns a, can be determined by using Nj as weighting
functions, so that
jf/y,<L()+p)_/i_) = 0 (5.5)
where ii is the domain of u. Equation (5.5) provides adequate number of equations to
calculate a, values. It results in to the following equations:
r
M
J Nj{L(u
b
) + X L(M) + p]dQ = 0, and (5.6)
a =i
fNJL(u
b
)dQ + X OifNUNdda + fp NjdQ = 0 (5.7)
a t=i __ 'a
Equation (5.7) can be restated in a matrix form to be used for computer programming
which suggests:
K a = f (5.8)
where,
kji = - / NjLiNddQ, and f
}
= - / NJUu
b
)dQ - J N
jP
dCl (5.9)
in which kji, at and fi are the elements of k, a and f, called the stiffness matrix, the
displacement vector and the load vector, respectively.
Three main criteria should be observed by the interpolation functions in order to obtain
convergence upon the exact solution. The criteria are: 1) rigid body displacement, 2) a
constant strain condition, and 3) continuity. The rigid body condition ensures that no strain
will be developed when a constant displacement is imposed on an element The second
criterion, i.e. constant strain, is required upon convergence to the exact solution when the
finite elements are very small which demands effectively constant stress distribution in each
element The third condition is the continuity of displacements or certain derivatives of
displacements. The mathematical proof of these criteria is not in the scope of this thesis and
is available in the mathematics references.
Chapter 5: Determination of Joint Rigidity Using Finite Element Method
77
5.2.2 Shell Finite Element
The F E M becomes complicated for certain types of structures such as plates and shells. A
review of the shell finite element literature reveals that shell and plate modelling have not
been solved as completely as two or three dimensional solid elements. Continuity and rigid
body modes are the properties that contribute most to the difficulty of a shell structure
modelling. There are basically three categories of finite elements that can be employed to
model a shell structure. These are:
1) plate elements, which are also called facet elements,
2) curved shell elements based on a shell theory, and
3) degenerated isoparametric elements.
Initially flat plate elements employing Kirchhoff theory were used for the finite element
analysis of shells. In this formulation shear deformations are neglected which makes it
difficult to satisfy interelement continuity on displacements and edge rotations. The
problems with using plate elements are the spurious bending moment they generate at the
intersection of the elements, and also uncoupled membrane forces and bending moments.
Huang (1989) states that in spite of certain shortcomings in such an approach, facet
elements are very efficient for the approximate analysis of many shell elements.
Despite the popularity of using shell elements which are formulated based on shell theories,
there are still some limitations concerning the true behaviour of a shell element. One of the
difficulties is with finding appropriate deformation idealisations which allow truly strain free
rigid body movements.
The third group of shell element formulation is based on Mindlin theory that includes the
effects of shear deformations. With this theory, the displacements and rotations of the mid-
surface normals are independent and the interelement continuity conditions on these
quantities can be satisfied as in the analysis of continua. This type of shell element which is
also called a degenerated shell element has become very popular. A well known type of a
degenerated shell element which is very efficient and simple, is the Semi Loof element a
family of A h m a d elements.
Finite Element analyses in this study were carried out using a computer package called
ALGOR. The shell element implemented in ALGOR is a flat plate element and will be further
explained in the next section. Although the results of F E analyses were up to 7 % different
from the experimental results for inplane bending, a comparison is also made in Section
5.3.3 between the performance of the flat plate element in ALGOR and other types of plate
and shell elements.
Chapter 5: Determination of Joint Rigidity Using Finite Element Method
78
5.3 Description of the Finite Element Package: A L G O R
ALGOR is a Finite Element Package developed by a company with the same name, Algor.
The linear analysis processor of this program, which was used in this study, is similar to the
program SAPIV; developed at the University of California, Berkeley (Bathe et al., 1974).
The major differences between SAPIV and ALGOR, in addition to the fact that ALGOR runs
on PC, are due to the powerful and useful pre- and post-processors accompany the latter
program. ALGOR consists of different programs or modules performing different tasks. A
list of the modules which were used in different steps of a T-joint analysis are given in
Appendix D.
5.3.1 Shell Element in ALGOR
The plate and shell element in ALGOR is a general quadrilateral element formed as an
assemblage of four triangular elements called L C C T 9 , developed by Clough and Felippa
(1968). This element has been counted as one of the most efficient mesh units for both plate
and thin shell applications by Clough and Felippa. Figure 5.1 shows the quadrilateral
element used in ALGOR, which is composed of four constant curvature triangular L C C T 9
elements.
Figure 5.1. Thin plate and shell element in ALGOR
Each LCCT9 element is in turn a condensed form of three triangular elements as shown in
Figure 5.2. The use of three triangular elements is to provide compatibility in the L C C T 9
element The cubic interpolation functions used in L C C T 9 can not produce fully compatible
triangular plate elements, therefore L C C T 9 has been divided into three sub-elements. The
central nodal point of the quadrilateral element is located at the average of the coordinates
of the four corner nodes. Generally six degrees of freedom can be associated with each
nodal point which gives a total of 24 degrees of freedom for a quadrilateral element in a
three dimensional position. In the case of a flat plate or a shallow shell, the rotation about
the normal to the plate should be eliminated. Further information on the plate and shell
element of ALGOR is available in the paper by Clough and Felippa (1968).
Chapter 5: Determination of Joint Rigidity Using Finite Element Method
79
Figure 5.2. LCCT9 element assembly
5.3.2 A Bench Mark on ALGOR
To evaluate the accuracy of the thin shell element in ALGOR, a simple example of a free
cylinder under internal pressure, shown in Figure 5.3, is analysed and its results are
compared with the results of other two types of plate and shell elements. This example was
originally reported in Knowles et al. (1976) to examine a flat plate and a curved shell
element, called T B 3 and G C S 8 respectively.
Overall geometry
D
Mesh for GCS8 andQ12
D
Coarse mesh for TBC3
Fine mesh for TBC3
Figure 5.3. Geometry and meshes used for cylinder underpressure
Chapter 5: Determination of Joint Rigidity Using Finite Element Method
80
W e call the shell element in ALGOR, Q12. The performance of the different shell elements,
TB3, G C S 8 and Q12, are shown in Table 5.1, indicating that Q 1 2 is suitable for
20 <_9/r<100, which is the applicable range in the offshore structures. It is also seen that
Q 1 2 performs reasonably satisfactory compared to G C S 8 element with a lower number of
nodal points. The plate element Q 1 2 does not produce a good result for D/T = 3.33, but this
is well outside the ranges of the tubular joint parameters in offshore towers. G C S 8 is a
semi-loof element and T B C 3 is a triangular element with six degrees of freedom at each
node. Descriptions of the finite elements, G C S 8 and T B C 3 , are reported in Appendix E.
Table 5.1. Results of different types of shell elements - cylinder under pressure
D/T
3.33
10
20
100
TB3
Coarse mesh Fine mesh
(NDOF* = 240)
(a) Displacements,
Error of -7.5% to + 5 %
(b)Membrane strains,
Correct
(a) Displacements,
Error of 3 3 %
(b) Membrane stresses,
Correct
(b) Displacements,
Garbage
(b) Membrane stresses,
Correct
(a) Displacements,
Garbage
(b) Membrane stresses
Correct
(NDOF* = 480)
(a) Error of 1 %
(b) Correct
(a) Error of 8 %
(b) Correct
(a) Error of -27% to +30%
(b) Correct
(a) Garbage
(b) Correct
GCS8
GCS8 mesh
(NDOF*=155)
(a) Correct
(b) Correct
(a) Correct
(b) Correct
(a) Correct
(b) Correct
Q12
GCS8 mesh
(NDOF* = 90)
(a) Error of-38%
(b) Error of - 8 %
(a) Error of+6%
(b) Error of -3%
(a) Error of-2%
(b) Error of-2%
(a) Error of-2%
(b) Error of - 2 %
* Number of Degrees of Freedom
5.3.3 Comparison of shell and three dimensional element performances
In order to verify the performance of the shell elements for F E modelling of the joints, two
analyses in IPB and O P B modes were carried out on joint T5, using shell element and three
dimensional solid element The results of the analyses are compared in Table 5.2. The shell
element has performed slightly better than the 3 D element for IPB mode, and its
performance in O P B mode is clearly better than 3 D element Both F E models had the same
mesh patterns as shown in Figure 5.4.
Chapter 5: Determination of Joint Rigidity Using Finite Element Method
81
Table 5.2. Comparison of shell and three dimensional solid elements
Shell finite element
3D finite element
Measurement
Natural frequency (Hz)
IPB
94.1
98.01
88.8
OPB
58.3
63.00
36.8
Finite Element
Error% =
Measuremnet
IPB O P B
+6 +58
+10 +71
5.4 M e s h generation procedure used to model T-joints
A typical mesh used to model the T-joints is shown in Figure 5.4. The modes of
deformation considered were inplane, out of plane bending and axial deformation of brace.
These modes are resulted in to one plane of symmetry/anti-symmetry along the longitudinal
axis of the chord, and one plane of symmetry/anti-symmetry across the branch section.
Therefore a model of one quarter of a T-joint with the appropriate boundary condition was
sufficient for analysis. The general layout for the data generation of the T-joints is shown in
Figure 5.5. Mesh generation of the T-joints were carried out using Superdraw (SD2), the
computer aided design module in ALGOR. Intersection coordinates of a chord and its
branch was produced by using a separate Fortran program written by the author and applied
as input to Superdraw. The chord and branch meshes were generated separately and then
connected to each other by the program Substruc (one of the modules in ALGOR).
Boundary conditions were applied to the F E models in Superdraw.
a) Three dimensional solid element
b) Plate and shell element
Figure 5.4. Typical mesh patterns used for the analysis of T-joints
Chapter 5: Determination of Joint Rigidity Using Finite Element Method
82
Calculation of the nodal co-
ordinates on intersection
brace
I
Construction of the brace
mesh in a planar co-
ordinate system (ALGOR)
I
Applying boundary
conditions and material
properties (ALGOR)
I
Decoding the graphic model
into ASCII form (ALGOR)
I
Folding the planar mesh
into a 3D co-ordinate system
I
j.
chord
Construction of the chord
mesh in a planar co-
ordinate system (ALGOR)
I
Applying boundary
conditions and material
properties (ALGOR)
I
Decoding the graphic model
into ASCII form (ALGOR)
I
Folding the planar mesh
into a 3D co-ordinate system
I
Joining chord mesh to
brace mesh (ALGOR)
Figure 5.5. Data generation procedure for tubular T-joints
5.5 Boundary conditions for the F E models
There are two categories of boundary conditions which can be applied to a T-joint or any
other structure, called here the internal and external boundary conditions. Internal boundary
conditions are those associated with the planes of symmetry or anti-symmetry; and external
boundary conditions are those assigned to the interface of a structure and its supports.
Applying the internal boundary conditions to an F E model will reduce the size of the model,
therefore less computer memory is required and also a higher accuracy for the solution can
be attained. However, it is essential that the internal boundary conditions produce exactly
the same deformations at the respective boundaries as when the structure was continuous at
those boundaries.
The T-joints had the conditions of symmetry and anti-symmetry, therefore only one quarter
of a joint was modelled. The boundary conditions applied to the T-joints are shown in
Figure 5.6.
Chapter 5: Determination of Joint Rigidity Using Finite Element Method
83
U
0
*
\
i
I
i
I
61
60 f
1000N
1B
17
\
ir
15
16
Loading and node numbering on
T10 wfth reference to Table 5.2


-
-Q


-
IPB
OPB
Plane
1-1, symmetry
2-2, anti-symmetry
chord ends
1-1, anti-symmetry
2-2, symmetry
chord ends
Sx
free
fixed
fixed
fixed
free
fixed
Boundary Conditions
6y
fixed
fixed
fixed
free
free
fixed
8
Z
free
free
free
fixed
fixed
fixed
e*
fixed
free
free
free
fixed
free
e-
free
free
free
fixed
fixed
free
e
z
fixed
fixed
free
free
free
free
1-1, symmetry
axial 2-2, symmetry
chord ends
free
free
fixed
fixed
free
fixed
free
fixed
free
fixed
fixed
free
free
fixed
free
fixed
free
free
Figure 5.6. Boundary conditions applied to FE models of the T-joints
The chord walls at the supports were restrained from translation in x and y directions to
maintain the roundness of the chord section. This is because the joint is defined as a part of
the structure where the beam theory is no longer valid and must be replaced by the shell
theory. Therefore, the end cross sections of a joint behave as rigid planes with no
ovalization. The internal boundary conditions, assumed in the analyses of a quarter of a
joint, were examined by analysing two different F E models of the same joint, TIO, under a
static point load at the top of the cantilever. One was a full model of TIO and the other was
one quarter of it. Results of the analyses are reported in Table 5.3, indicating that the
assumed internal boundary conditions for the models, at symmetry and anti-symmetry zones
are acceptable.
Chapter 5: Determination of Joint Rigidity Using Finite Element Method
84
Table 5.3. Application of internal boundary conditions to the FE model of T-joint TIP (IPB) (cont.)
Node No.*
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
Complete model
6>_)
6<rad)
-1.7163E-12
-2.5008E-08
1.2821E-11
2.1821 E-03
6.4962E-12
7.1041 E-03
-8.4091 E-12
8.6872E-03
2.4173E-12
3.6357E-02
1.6454E-11
6.281 OE-03
-1.7312E-11
-5.3693E-02
-5.7971 E-11
-6.4790E-02
-3.1568E-11
-6.4209E-02
-7.4752E-12
-4.5478E-02
0.0000E+00
-2.1811E-02
0.0000E+00
-2.5357E-03
0.0000E+00
5.5674E-03
-1.1231E-12
3.3087E-03
-2.6760E-12
4.3068E-05
-1.0867E-12
7.0691 E-09
2.5844E-11
6.8927E-03
2.9054E-11
-1.7660E-04
2.9122E-11
2.2591 E-03
2.6045E-11
4.2533E-03
2.5998E-11
4.7094E-03
2.5995E-11
3.2768E-03
2.5833E-11
1.1407E-03
2.5418E-11
-3.8399E-04
2.5050E-11
-1.2925 E-03
2.4773E-11
-1.7899E-03
5y(m)
6y(rad)
-8.3464E-12
-7.8849E-02
-9.0210E-12
-8.0667E-02
-9.6672E-12
-8.6299E-02
-1.4899E-11
-9.2282E-02
-1.4501 E-11
-8.0755E-02
-1.2650E-11
-7.6181E-02
-3.4945E-11
-7.6540E-02
-6.5757E-11
-2.1633E-02
-3.6042E-11
1.4029E-02
0.0000E+00
2.9298E-02
1.3603E-11
1.8154E-02
7.4700E-12
-1.0221 E-03
3.3415E-12
-3.2668E-03
5.2698E-12
7.9826E-04
6.4871 E-12
2.2450E-03
5.9530E-12
1.9156E-03
-1.9544E-11
-1.1519E-01
-2.6657E-11
-8.7149E-02
-3.5038E-11
-9.2651 E-02
-3.7981 E-11
-9.8018E-02
-4.6567E-11
-9.9044E-02
-6.7108E-11
-1.0624E-01
-8.7552E-11
-1.0928E-01
-1.0748E-10
-1.1293E-01
-1.2651 E-10
-1.1650E-01
-1.4469E-10
-1.1995E-01
8z(m)
e
z
(rad)
1.3214E-05
-1.0487E-07
1.3574E-05
-2.8538E-08
1.1962E-05
8.1788E-08
9.6660E-06
3.0550E-08
5.1620E-06
-1.1088E-07
1.1240E-06
5.7145E-08
-6.3856E-07
3.761 OE-07
-3.4163E-06
3.3281 E-08
-6.4916 E-06
-1.7717E-07
-7.4030 E-06
-7.8433E-08
-6.9535 E-06
-2.3341 E-10
-5.4822E-06
1.4185E-08
-3.8531 E-06
1.6397E-09
-2.5970 E-06
-4.8369E-09
-1.6044E-06
6.4709E-10
-1.2053E-06
8.5135E-10
2.3400E-05
-2.3321 E-08
5.1465E-05
-3.3055 E-08
7.9433E-05
-1.5242E-08
1.0824E-04
-1.2445 E-08
1.3787E-04
-3.8042 E-08
1.9031 E-04
-4.1206E-08
2.4471 E-04
-4.0137E-08
3.0098E-04
-3.8842 E-08
3.5904E-04
-3.6682E-08
4.1885E-04
-3.5298 E-08
One quarter model
8x(m)
8(rad)
0
0
0
2.1827E-03
0
7.1067E-03
0
8.6901 E-03
0
3.6381 E-02
0
6.2655E-03
0
-5.3699E-02
0
-6.4787E-02
0
-6.4209E-02
0
-4.5481 E-02
0
-2.1813E-02
0
-2.5360E-03
0
5.5681 E-03
0
3.3091 E-03
0
4.2995E-05
0
0
0
6.8920E-03
0
-1.8862E-04
0
2.2592E-03
0
4.2564E-03
0
4.7036E-03
0
3.2767E-03
0
1.1422E-03
0
-3.8608E-04
0
-1.2920E-03
0
-1.7903E-03
5y(m)
0
y
(rad)
0
-7.8851 E-02
0
-8.0670E-02
0
-8.6302E-02
0
-9.2286 E-02
0
-8.0752E-02
0
-7.6190E-02
0
-7.6542E-02
0
-2.1629E-02
0
1.4030E-02
0
2.9298E-02
0
1.8155E-02
0
-1.0215E-03
0
-3.2672 E-03
0
7.9813E-04
0
2.2451 E-03
0
1.9156E-03
0
-1.1531 E-01
0
-8.6915E-02
0
-9.2708 E-02
0
-9.8007E-02
0
-9.9008 E-02
0
-1.0629E-01
0
-1.0929E-01
0
-1.1294E-01
0
-1.1652E-01
0
-1.1997E-01
5z(m)
0z(rad)
1.3216E-05
0
1.3575E-05
0
1.1963E-05
0
9.6674E-06
0
5.1632E-06
0
1.1253E-06
0
-6.3742 E-07
0
-3.4155E-06
0
-6.4912E-06
0
-7.4031 E-06
0
-6.9538E-06
0
-5.4824E-06
0
-3.8532E-06
0
-2.5971 E-06
0
-1.6044E-06
0
-1.2054E-06
0
2.3402E-05
0
5.1468E-05
0
7.9436E-05
0
1.0824E-04
0
1.3788E-04
0
1.9032E-04
0
2.4472E-04
0
3.0099E-04
0
3.5905E-04
0
4.1887E-04
0
Chapter 5: Determination of Joint Rigidity Using Finite Element Method
85
Table 5.3. Application of internal boundary conditions to the FE model of T-joint TIP, IPB (cont.)
Node No.*
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
Complete model
8x(m)
O^rad)
2.4592E-11
-1.9413E-03
2.4509E-11
-1.9517E-03
2.4473E-11
-1.8840E-03
2.4489E-11
-1.7925E-03
2.4527E-11
-1.7367E-03
2.4553E-11
-1.6482 E-03
2.4594E-11
-1.5825 E-03
2.4616E-11
-1.5399E-03
2.4638E-11
-1.4625E-03
2.4645E-11
-1.4153E-03
2.4632E-11
-1.3509E-03
2.4630E-11
-1.2947E-03
2.4611 E-11
-1.2532 E-03
2.4591 E-11
-1.1799E-03
2.4566E-11
-1.1256 E-03
2.4535E-11
-1.0774E-03
2.4515E-11
-1.0142E-03
2.4483E-11
-9.6329E-04
2.4458E-11
-8.9900E-04
2.4432E-11
-8.4961 E-04
2.4404E-11
-7.9028 E-04
2.4388E-11
-7.3456E-04
2.4361 E-11
-6.8374E-04
2.4345E-11
-6.1142E-04
2.4331 E-11
-5.6129E-04
2.4318E-11
-5.0563E-04
8y(m)
O^rad)
-1.6231 E-10
-1.2326E-01
-1.7972E-10
-1.2667E-01
-1.9760E-10
-1.2996E-01
-2.1497E-10
-1.3311 E-01
-2.3251 E-10
-1.3625E-01
-2.5013E-10
-1.3904E-01
-2.6792E-10
-1.4206E-01
-2.8578E-10
-1.4469E-01
-3.0374E-10
-1.4738E-01
-3.2176E-10
-1.4997E-01
-3.3977E-10
-1.5233E-01
-3.5782 E-10
-1.5477E-01
-3.7587E-10
-1.5697E-01
-3.9390E-10
-1.5910E-01
-4.1191 E-10
-1.6122E-01
-4.2988 E-10
-1.6309E-01
-4.4782E-10
-1.6500E-01
-4.6573E-10
-1.6672E-01
-4.8360E-10
-1.6837E-01
-5.0143E-10
-1.6994E-01
-5.1922E-10
-1.7136E-01
-5.3701 E-10
-1.7275E-01
-5.5536E-10
-1.7401E-01
-5.7307E-10
-1.7517E-01
-5.9077E-10
-1.7623E-01
-6.0844E-10
-1.7716E-01
8
z
(m)
6,(rad)
4.8044E-04
-3.4530 E-08
5.4369E-04
-3.4222E-08
6.1090E-04
-3.4163E-08
6.7750E-04
-3.4501 E-08
7.4567E-04
-3.4726E-08
8.1537E-04
-3.4974E-08
8.8652E-04
-3.5248 E-08
9.5909E-04
-3.5369E-08
1.0330E-03
-3.5581 E-08
1.1083E-03
-3.5605E-08
1.1848E-03
-3.5623E-08
1.2625E-03
-3.5680 E-08
1.3414E-03
-3.5647E-08
1.4214E-03
-3.5620E-08
1.5025E-03
-3.5542E-08
1.5846E-03
-3.5469E-08
1.6676E-03
-3.5434E-08
1.7516E-03
-3.5335E-08
1.8364E-03
-3.5285 E-08
1.9220E-03
-3.5192E-08
2.0084E-03
-3.5142E-08
2.0955E-03
-3.5109E-08
2.1863E-03
-3.5014E-08
2.2747E-03
-3.4982E-08
2.3636E-03
-3.4935E-08
2.4530E-03
-3.4896 E-08
One quarter model
8x(m)
6_<rad)
0
-1.9421 E-03
0
-1.9519E-03
0
-1.8847E-03
0
-1.7927E-03
0
-1.7370E-03
0
-1.6487E-03
0
-1.5829E-03
0
-1.5401 E-03
0
-1.4630 E-03
0
-1.4155E-03
0
-1.3514E-03
0
-1.2949E-03
0
-1.2537E-03
0
-1.1802E-03
0
-1.1259E-03
0
-1.0777E-03
0
-1.0145 E-03
0
-9.6360E-04
0
-8.9924E-04
0
-8.4980E-04
0
-7.9053E-04
0
-7.3475 E-04
0
-6.8394E-04
0
-6.1160E-04
0
-5.6142E-04
0
-5.0579E-04
8y(m)
0y(rad)
0
-1.2328E-01
0
-1.2669E-01
0
-1.2998E-01
0
-1.3313E-01
0
-1.3627E-01
0
-1.3906E-01
0
-1.4209E-01
0
-1.4471 E-01
0
-1.4740E-01
0
-1.4998E-01
0
-1.5235E-01
0
-1.5479E-01
0
-1.5698E-01
0
-1.5911 E-01
0
-1.6123E-01
0
-1.6311 E-01
0
-1.6502E-01
0
-1.6673E-01
0
-1.6839E-01
0
-1.6995E-01
0
-1.7138 E-01
0
-1.7277E-01
0
-1.7402E-01
0
-1.7518E-01
0
-1.7624E-01
0
-1.7718E-01
8_(m)
6_(rad)
4.8045E-04
0
5.4371 E-04
0
6.1092E-04
0
6.7753E-04
0
7.4570E-04
0
8.1540E-04
0
8.8656E-04
0
9.5914E-04
0
1.0331 E-03
0
1.1083E-03
0
1.1849E-03
0
1.2626E-03
0
1.3415E-03
0
1.4215E-03
0
1.5026E-03
0
1.5847E-03
0
1.6677E-03
0
1.7517E-03
0
1.8365E-03
0
1.9221 E-03
0
2.0085E-03
0
2.0956E-03
0
2.1864E-03
0
2.2748E-03
0
2.3637E-03
0
2.4532E-03
0
Chapter 5: Determination of Joint Rigidity Using Finite Element Method
86
Table 5.3. Application of internal boundary conditions to the FE model of T-joint TIP, IPB (cont.)
Node No.*
53
54
55
56
57
58
59
60
61
Natural freque
Complete model
8x(m)
0x(rad)
2.4315E-11
-4.4219E-04
2.4316E-11
-3.8249E-04
2.4339E-11
-3.2054E-04
2.4392E-11
-2.6894E-04
2.4508E-11
-2.2389E-04
2.4693E-11
-2.5251 E-04
2.4990E-11
-2.5280E-04.
2.5289E-11
-1.6593E-04
2.5449E-11
-1.6276E-03
8><m)
0>(rad)
-6.2609E-10
-1.7802E-01
-6.4373E-10
-1.7876E-01
-6.6139E-10
-1.7941 E-01
-6.7914E-10
-1.7995E-01
-6.9715E-10
-1.8041 E-01
-7.1571 E-10
-1.8091 E-01
-7.3523E-10
-1.8061 E-01
-7.5620E-10
-1.8113E-01
-7.7948E-10
-1.8209E-01
ncies: co i *= 40.60 Hz a>2*=
5_(m)
0_(rad)
2.5429E-03
-3.4866E-08
2.6332E-03
-3.4862E-08
2.7239E-03
-3.4963E-08
2.8148E-03
-3.5283E-08
2.9060E-03
-3.6084E-08
2.9975E-03
-3.7457E-08
3.0890E-03
-4.0049E-08
3.1807E-03
-4.3600E-08
3.2725E-03
-4.9592E-08
319.27 Hz
One quarter model
8x(m)
0x(rad)
0
-4.4232E-04
0
-3.8261 E-04
0
-3.2063E-04
0
-2.6901 E-04
0
-2.2401 E-04
0
-2.5249E-04
0
-2.5280E-04
0
-1.6606E-04
0
-1.6281 E-03
a>i*= 40.69 Hz
dy(m)
O^rad)
0
-1.7803E-01
0
-1.7877E-01
0
-1.7943E-01
0
-1.7997E-01
0
-1.8042E-01
0
-1.8093E-01
0
-1.8062E-01
0
-1.8115E-01
0
-1.8211 E-01
*
2 =
8_(m)
0_(rad)
2.5431 E-03
0
2.6334E-03
0
2.7241 E-03
0
2.8150E-03
0
2.9062E-03
0
2.9976E-03
0
3.0892E-03
0
3.1809E-03
0
3.2727E-03
0
319.63 Hz
*Nodal point locations are shown in Figure 5.6.
*Natural frequencies are only due to inplane bending mode of brace.
It is seen in Table 5.3 that the deformation values, including displacements and rotations,
have improved by adopting the correct internal boundary conditions. This is due to the less
mathematical operations carried out by the computer for the smaller model of the structure.
Therefore, a less round off error will occur.
In the example above, however, the improvement in the results have not been significant.
The values of 8, o"
y
, and 0
Z
that should be theoretically zero due to the symmetry are in the
orders of 10"
11
, 10"
11
, and 10"
8
respectively and have reduced to 0 in the one quarter model.
The maximum variations between the two models for 8
Z
, Q
x
and 0,, are 0.18%, 6 % and
0.27% at nodal points 7, 18, and 18, respectively. Furthermore, the computing time for the
full model was 29.9 minutes, whereas it was 1.6 minutes for the one quarter model
indicating a 18.7 times difference. Using a quarter of a joint mainly saves computational
time and the modelling effort. It should be stated that three analyses are required on a small
model in order to obtain the similar results of the analysis on a full model.
5.5.1 Effect of joint support conditions
The influence of chord support fixity on dynamic behaviour of a T-joint should be evaluated
in order to obtain a reliable estimate of the joint flexibility. T w o different support conditions
i.e. fixed and hinged were examined on the shell Finite Element models of the joint
specimens; similar to that discussed in Section 4.5 for the models constructed from one
Chapter 5: Determination of Joint Rigidity Using Finite Element Method
87
dimensional beam elements. The results of the analyses, reported in Table 5.4, confirm the
conclusion made in Section 4.5, that the fixity of the end supports has very little influence
on the natural frequencies associated with the inplane bending. The results of the ID beam
element analyses are also reported in Table 5.4. It should be noted that these results are due
to the joints supported on hinges, with no transverse springs (the model shown in Figure
5.6). Therefore, there is a difference between the results presented in Table 5.4 and those in
Table 5.5 or Table 4.6.
Considering the higher accuracy of the Finite Element method, it can be concluded that
there is a very little influence from the chord support conditions on the dynamic behaviour
of a T-joint vibrating in inplane bending mode. It is seen in Table 5.4 that the variation of
boundary conditions from hinged to fixed has its biggest effect on the joints with large
diameter ratios (Tl, T 5 and T9). Hence, the joints with |_ = 1 or generally the joints with
large diameter ratios may be more sensitive to the boundary conditions. This property of
stiff joints will be also shown in Section 5.6.1 when comparing the results of F E analyses
and experiments.
Table 5.4. Calculated results for the FE models of the T-joints with different support conditions (IPB)
Joint
Tl (shell element)
(ID beam)
Tl
T3
T4
T5
T6
T7
T8
T9
TIO
Til
Natural free
Fixed supports
G)f(Hz) Q)f(Hz)
92.05 547.66
129.98 772.80
48.47 331.11
69.62 421.53
51.22 339.91
70.56 427.48
31.77 202.02
36.81 223.74
81.45 530.78
125.44 744.16
43.64 321.73
69.37 419.58
46.55 330.42
70.35 425.83
29.74 194.77
36.80 223.65
73.71 514.65
121.84 723.29
40.85 320.57
70.01 423.06
25.80 185.67
36.78 223.49
uency (IPB)
Hinged supports
0)f(Hz) 0)f(Hz)
91.01 539.54
127.08 743.72
48.34 330.08
69.28 419.07
51.09 338.93
70.27 425.36
31.76 201.93
36.80 223.64
80.35 522.36
121.61 715.48
43.51 320.82
68.95 416.64
46.43 329.56
69.99 423.28
29.73 194.68
36.78 223.51
72.46 501.20
117.28 695.85
40.60 319.27
69.56 419.89
25.78 185.58
36.75 223.32
Frequency ratio
cof/co? cof/cof
1.01 1.02
1.02 1.04
1.00 1.00
1.00 1.01
1.00 1.00
1.00 1.00
1.00 1.00
1.00 1.00
1.01 1.02
1.03 1.04
1.00 1.00
1.01 1.01
1.00 1.00
1.01 1.01
1.00 1.00
1.00 1.00
1.02 1.03
1.04 1.04
1.01 1.00
1.01 1.01
1.00 1.00
1.00 1.00
Chapter 5: Determination of Joint Rigidity Using Finite Element Method 8
5.6 Finite Element analysis of the tested T-joints
Chapter 6 of this thesis reports the results of a parametric Finite Element study carried out
to establish a relationship between non-dimensional parameters and stiffness of the T-joints.
There are 270 analyses reported in Chapter 6 and it was required to experimentally verify
the F E results. T o do so, a series of Finite Element analyses were performed to obtain the
natural frequencies of the tested joints. One quarter of a joint with hinged and fixed
boundary conditions was modelled for inplane bending and out of plane bending modes,
respectively. The first three mode shapes, calculated for the F E model of the joint T 2 are
shown in Figure 5.7, as examples. The second natural frequencies of the joint specimens
were measured experimentally and compared with those obtained from F E analysis.
Mode 1 Mode 2 Mode 3
Figure 5.7. T-joint T2, first three mode shapes in IPB mode
Chapter 5: Determination of Joint Rigidity Using Finite Element Method 89_
5.6.1 Comparison of the Finite Element analysis and test results
The Finite Element analysis and experimental results for the IPB and O P B natural
frequencies of the T-joint specimens are reported in Table 5.5. The column labelled error
gives the relative difference between the results of the Finite Element analyses and those
from experiments.
Table 5.5. Results ofFE analyses and experiments
Natural frequency (Hz)
Joint
Tl
T2
T3
T4
T5
T6
T7
T8
T9
TIO
Til
Finite Element
IPB OPB
101.7 67.6
53.0 33.5
55.1 36.0
32.9 27.1
94.1 58.3
48.8 27.5
51.1 29.7
31.2 23.9
88.1 53.2
46.1 23.7
27.8 18.6
Measurement
IPB O P B
95.0 46.1
55.0 30.8
56.3 33.4
33.8 27.3
88.8 36.8
50.0 24.4
51.3 26.5
31.3 23.1
82.5 33.3
46.3 20.1
28.8 17.9
_, Finite Element
Error% =
Measurement
IPB O P B
+7 +47
-4 +9
-2 +8
-3 0
+6 +58
-2 +13
0 +12
0 +3
+7 +60
0 +17
-3 +4
It is seen in Table 5.5 that the results of the F E analyses are close to those obtained from
experiments, especially for inplane bending mode.
For the IPB natural frequencies the bigger discrepancies are seen for joints Tl, T5, and T9.
The diameter ratio (p) is noted to be one for all these three joints. Apparently this type of
tubular joints (p = 1) has always been problematic. Nearly all parametric formulae for the
estimation of stress concentration factors and also the D N V formulae for the stiffness
determination of T-joints do not cover the joints with p = 1. One reason for this, also stated
by Smedley and Fisher (1991), is the separation of the weld toes at the saddle points which
is smaller than the nominal diameter size of the brace, as shown in Figure 5.8. A formula is
recommended by Lloyd's Register as p' = l-(-sin
065
\|0 to be used instead of P for the joints
Y
with p = 1. \|f is the degree of weld cut-back as shown in Figure 5.8. Roesset and Pan
(1988) studied the behaviour of X-joints under axial loading and extensively used F E
method. They have observed that the elastic stiffness obtained experimentally for large
diameter tubes are always smaller than those obtained by computer modelling. This was also
Chapter 5: Determination of Joint Rigidity Using Finite Element Method
90
observed on the joints with p = 1 tested in this study. However, the difference between
calculated and experimental results was caused by the reasons explained below.
10 30
7
ideal standard extreme
Figure 5.8. Weld cut-back at the saddle on joints with p = 1 (Smedley and Fisher, 1991)
Out of plane bending mode had the worst agreement with the experimental results. The
calculated natural frequencies of Tl, T 5 and T 9 were 4 7 % , 5 8 % , and 6 0 % higher than the
measured values respectively. This may be attributed to the inadequate fixity of the joint
support assembly to the ground. The support assembly, as shown in Figures 4.8 and 4.16,
was clamped to the floor with high strength bolts. It is seen in Table 5.5 that the most
flexible joints such as T4, T8, and Tl 1 show the best agreement between the calculation and
experiment either in IPB or O P B mode. The remaining joints which have a P equal to 0.52
and therefore possess a medium flexibility compared to the joints with P = 0.28 and P = 1
show a moderate agreement which is not as good as the joints with P = 0.28 but better than
the joints with P = 1. This variation of agreement between calculation and experiment
suggests that a different attachment system should be tried for connection of the support
assembly to the floor.
The correlation between the results of the shell F E analyses and experiments is shown in
Figure 5.9. These results indicate that, for engineering purposes, experiment can be replaced
by Finite Element analysis to predict the dynamic behaviour of a single T-joint. However,
this can be only used if the structure is not deficient such as the situation right after the
fabrication of a specimen which was the case for the T-joints in this study.
Chapter 5: Determination of Joint Rigidity Using Finite Element Method
91
100-
N 80
I
3
S 60
L.
LJ
40-
20-
0-^
1
40
2.
* * * * * Inplane bending
D a a a Out of plane bending
T1
T5
T9
a
/
xT3
T3 '
T2 /
*T4
T10^/*T6
/
/
or
T11
s
T8 T4
T6DrtK_
,w
/
/
/
i_.
100
T1
/
T5
T9
/
* /
* /
* /
/
-100
-80
-60
-40
T 1 1 1 1 1 r -
20 40 60 80
Measured Frequency (Hz)
r-
100
-20
Figure 5.9. Correlation between natural frequencies obtained from FE analysis and experiment
5.7 Summary
In this chapter a brief look at the Finite Element Method was made and results of the Finite
Element analyses of the tested T-joints were reported and discussed..
It was pointed out that the Finite Element method becomes complicated for shell and plate
structures due to satisfying the conditions of continuity and rigid body modes. The different
types of shell formulation were also explained to be 1) plate elements, 2) curved shell
elements, and 3) degenerated shell elements. The Finite Element computer program,
ALGOR, and its plate element used in this study were described. This type of element has
been proved to be efficient for the analysis of many shell structures. This was also shown by
comparing the results of a bench mark study on the plate element in ALGOR with those
from a curved thin shell element and a triangular flat plate element The results of such a
comparison indicate that a good performance can be obtained by using the plate element of
ALGOR in the analysis of a cylindrical structure. A comparison of the results of two
analyses on a T-joint using plate elements and three dimensional solid elements with
Chapter 5: Determination of Joint Rigidity Using Finite Element Method
92
experimental results showed that plate elements can produce a more accurate result than 3 D
elements.
The mesh generation of the tubular joints and the boundary conditions adopted for the joint
models were described. It was shown that modelling a quarter of a joint decreases the
computing time by at least 6 fold with respect to a full model analysis time. Based on the FE
analyses, it was shown that the effect of boundary conditions on the inplane mode of
vibration is very little with the maximum of 4 % for the joints with P = 1 when the boundary
condition vary from hinged to fixed.
The F E analysis results had a good relationship with the test results, indicating that for a T-
joint with no noticeable deficiency Finite Element analysis works to an acceptable
engineering accuracy. For IPB mode, the mean error of the absolute differences between
Finite Element and experimental results were 3.1% with the maximum difference being for
the joints with P = 1. Similar error for O P B was 2 1 % when all eleven joint specimens were
considered and 8.3% when the joints with P=l were excluded.
Chapter 6
RIGIDITY AS FUNCTION OF JOINT
GEOMETRICAL PARAMETERS
Chapter 6: Rigidity as Function of Joint Geometrical Parameters
94
6.1 Introduction
Formulation of rigidity in terms of joint non-dimensional parameters is possible and
necessary in order to apply stiffness of the joints as a routine to the structural analysis
computer programs. Such formulae as discussed in Chapter 2 have been proposed by others
such as D N V (1977), Efthymiou (1985), Fessler (1986) and Ueda (1990) based on static
analyses. A set of formulae for different deflection modes of a T-joint (i.e. inplane bending,
out of plane bending and brace axial deformation) is proposed in this chapter, based on a
total of 270 dynamic Finite Element analyses. A beam member with rotational and axial end
springs, as described in Chapter 3, is employed as the theoretical model.
6.2 Finite Element analysis results
The flexibility of tubular T-joints are determined using resonant frequencies. A series of
Finite Element analyses were carried out to obtain the natural frequencies of the T-joints.
The deformation modes of brace include inplane bending, out of plane bending, and axial
deflection of brace. The following ranges for non-dimensional parameters of the joints are
considered:
0.28 <P< 1.0
13.36 < v < 28.83 (6.1)
0.2 < x < 1.0
in which p is the diameter ratio of the brace to the chord (dID), y is the chord diameter to
thickness ratio (DI2T) and x is the thickness ratio of the brace to the chord wall (tIT). When
these ranges are compared with those of the formulae proposed by D N V (1977), Efthymiou
(1985) and Fessler (1986), they cover a wider range of p values. Table 6.1 reports the
different non-dimensional parameters used in the Finite Element analysis of the T-joints in
this study. It is seen that there are totally 90 analyses required to cover the range of
parameters stated in Table 6.1. Since only one quarter of a joint was modelled and each
model was suitable only for one mode of deformation, therefore three analyses were carried
out for each parameter. This made the total number of analyses to be 270.
Chapter 6: Rigidity as Function of Joint Geometrical Parameters
95
Table 6.1. Non-dimensional parameters used for FE models of T-joints
parameter value
V(d/D)
y(DI2T)
x(tlT)
a (2UD)
0.28
13.36
0.20
13.69
0.41
15.22
0.50
0.52
17.12
1.0
0.64
22.82
0.75
28.83
1.0
The geometry of the Finite Element models is shown in Figure 6.1. The dimensions adopted
are similar to the test specimens described in Chapter 4. A chord length of 1500mm and a
diameter size of 219.1mm were assumed for all models which resulted in an a value of
13.69.
t
variable
*J-
variable
All dimensions
in millimeters
M
1=1230
z
*
A
L=1500
A-A
Figure 6.1. Geometry of Finite Element models used for flexibility analysis of T-joints
Few findings exist in the literature regarding the effects of a ratio on the measurements
carried on a tubular joint. This parameter represents the chord length significance. In
practice more than 6 0 % of tubular joints have a values in excess of 20 and, for 3 5 % of
joints a exceeds 40 (Smedley and Fisher 1991). Efthymiou (1985) has shown that for IPB
and O P B modes, a chord length of minimum 6D yields stiffness coefficients that are within
1 % to 4 % of the true stiffness. A n idealised analytical solution for decay length has been
reported by Efthymiou (1985) which agrees with the minimum length of 6Z>. This solution is
presented in Appendix F. Roesset and Pan (1988) conducted a study on the effects of chord
length and boundary conditions on the behaviour of tubular X-joints. They analysed six joint
geometries under tensile and compressive loads, using non-linear Finite Element method. It
is concluded by Roesset and Pan that for a chord length of two chord diameters or less, the
end conditions substantially affect the analysis results in relation to ultimate load and
Chapter 6: Rigidity as Function of Joint Geometrical Parameters
96
stiffness of the joints. They added that for a chord length of four diameters or more the
effect of end conditions is relatively small.
In this study a = 13.69 (L/D = 6.85) was selected for the F E models as a representative for
the purpose of practical design.
The natural frequencies and mode shapes of a structure can be determined by calculating the
eigenvalues and eigenvectors of its stiffness matrix. This is called dynamic modal analysis in
ALGOR computer package and carried out by the program module SSAP1 in the package.
The identification names and the specification of the individual F E models are given in Table
6.2. There are 90 joints in Table 6.2, which are basically three batches of 30 joints. Each
batch has a different x value. The joint names with suffixes 'a', 'b' and 'c' have x values
equal to 0.2,0.5 and 1.0, respectively. T 2 and T6 have the same geometry as T3 and T7.
Table 6.2. Identification names and individual joint parameters of FE models
joint
Tla,b,cT
T2a,b,c
T3a,b,c
T4a,b,c
T5a,b,c
T6a,b,c
T7a,b,c
T8a,b,c
T9a,b,c
T10a,b,c
Tlla,b,c
P
1.0
0.52
0.52
0.28
1.0
0.52
0.52
0.28
1.0
0.52
0.28
Y
13.36
13.36
13.36
13.36
17.12
17.12
17.12
17.12
22.82
22.82
22.82
joint
T12a,b,c
T13a,b,c
T14a,b,c
T15a,b,c
T16a,b,c
T17a,b,c
T18a,b,c
T19a,b,c
T20a,b,c
T21a,b,c
T22a,b,c
P
0.41
0.64
0.75
0.41
0.64
0.75
0.41
0.64
0.75
1.0
0.75
Y
13.36
13.36
13.36
17.12
17.12
17.12
22.82
22.82
22.82
15.22
15.22
joint
T23a,b,c
T24a,b,c
T25a,b,c
T26a,b,c
T27a,b,c
T28a,b,c
T29a,b,c
T30a,b,c
T31a,b,c
T32a,b,c
P
0.64
0.52
0.41
0.28
1.0
0.75
0.64
0.52
0.41
0.28
Y
15.22
15.22
15.22
15.22
28.83
28.83
28.83
28.83
28.83
28.83
t Joint names with suffix 'a' have T values = 0.2, joint names with suffix 'b' have i values = 0.5, and
joint names with suffix 'c' have x values = 1.0.
Chord length parameter (a) is 13.69 for all models.
The theoretical model in this study employs the first natural frequency relevant to the
desired deformation mode. However, it is generally possible to include the higher natural
frequencies in the derivation of joint stiffness. The natural frequencies and stiffness of the T-
joints, analysed by the Finite Element method, are reported and discussed in Sections 6.2.1
to 6.2.3 for different modes of deformation.
6.2.1 Inplane bending mode
Tables 6.3 report the results of the analyses for inplane bending mode. Also, Figures 6.2 to
6.4 illustrate these results as relationships between p, y, x and joint stiffness. The
relationships suggest that an equation can be established to cover the calculated stiffness
values.
Chapter 6: Rigidity as Function of Joint Geometrical Parameters
97
Table 6.3a. Calculated natural frequencies and stiffness of T-joint FE models, IPB (t =
joint
Tla
T2a
T3a
T4a
T5a
T6a
T7a
T8a
T9a
TlOa
Tlla
NF*(Hz) *(Nm)
106.87 .645E7
60.57 .138E7
34.61 .508E6
105.25 .463E7
59.17 ~mE6~'
34.77 .265E6
103.31 .308E7
57.64 591E6
32.60 .150E6
joint
T12a
T13a
T14a
T15a
T16a
T17a
T18a i
T19a
T20a
T21a
T22a
N F ( H z )
48.34
73.17
85.02
48.02
72.38
84.07
46.98
70.77
82.32
106.21
84.72
*(Nm)
.794E6
.231E7
.349E7
.547E6
.160E7
.243E7
.339E6
.106E7
161E7
.547E7
.284E7
joint
T23a
T24a
T25a
T26a
T27a
T28a
T29a
T30a
T31a
T32a
NF*(Hz)
72.96
60.88
48.39
35.19
98.57
80.81
69.32
57.87
45.92
32.48
P.2)
Jt(Nm)
.188E7
.119E7
.644E6
.348E6
.195E7
.117E7
.740E6
.439E6
235E6
.971E5
*NF: Natural Frequency
Table 6.3b. Calculated natural frequencies and stiffness of T-joint i
joint
Tib
T2b
T3b
T4b
T5b
T6b
T7b
T8b
T9b
flOb
Tllb
NF*(Hz)
101.74
52.65
31.56
87.30

50.64
30.50
86.30
48.32
28.29
Jfc(Nm)
8.58E6
1.62E6
4.74E5
6.10E5
1.07E6
2.98E5
4.18E6
6.81E5
1.53E5
joint
T12b
T13b
T14b
T15b
T16b
T17b
T18b
T19b
T20b
T21b
T22b
NF*(Hz)
42.76
62.82
72.49
41.62
61.13
70.65
39.81
58.67
68.05
89.54
71.68
*(Nm)
9.47E6
2.77E6
4.28E6
6.30E5
1.91E6
2.93E6
3.79E5
1.24E6
1.90E6
7.18E6
" 3.55E6
VEmode
joint
T23b
T24b
T25b
T26b
T27b
T28b
T29b
T30b
T31b
T32b
Is, IPB (x =
NF*(Hz)
62.10
52.31
42.34
30.93
84.03
65.94
56.63
47.54
38.18
27.74
P.5)
Jt(Nra)
2.31E6
1.35E6
7.62E5
3.59E5
3.05E6
1.35E6
8.45E5
4.97E5
2.60E5
1.24E5
*NF: Natural Frequency
Table 6.3c. Calculated natural frequencies and stiffness of T-joint i
joint
Tic
T2c
T3c
T4c
T5c
T6c
T7c
T8c
T9c
TlOc
Tile
NF*(Hz)
76.44
45.19

28.86
74.51
42.83

28.43
72.46
40.69
24.72
*(Nm)
.106E8
.185E7

.556E6
.763E7
.123E7

.414E6
.509E7
.773E6
.180E6
joint
T12c
T13c
T14c
T15c
T16c
T17c
T18c
T19c
T20c
T21c
T22c
NF*(Hz)
37.97
53.83
60.63
36.22
51.66
58.42
33.96
48.89
56.53
75.69
59.48
*(Nm)
.115E7
.316E7
.478E7
.746E6
.217E7
y- -327E7
.450E6
.140E7
.223E7
.898E7
.394E7
rE models, IPB (i =
joint NF*(Hz)
T23c 52.82
T24c 44.98
T25c 37.18
T26c | 28.75
T27c 70.75
T28c 54.33
T29c 46.72
T30c 39.41
T31c 32.05
T32c 24.10
1.0)
*(Nm)
.263E7
.160E7
.897E6
.474E6
.379E7
.156E7
.982E6
.557E6
.294E6
.129E6
*NF: Natural Frequency
98
s
ad
LUN
vd
(Bdl) X
2
.
E
6
-
0
.

-
-
0
.
2

II
*">
II
vo /
en/
en
^H
JL
Tl
I
1 1
a/
JL
" i i
es /
H
_-
1
so/
JL
1 1 !
m
oo
oo
^ ri>
\ / $-
i
WN (Bdl) >l
oo
d
vo ca
d
V O V O V O V O N O N O N O V O VO
es
6
oo

NO C_L

i 1 1 1 1 1 1
VO vo NO vo vo NO vo
UJ UJ UJ w UJ u uJ
r - ^ v o i r i ^ t en es ^
WN (adi) a
e s
<=>
II
%
<

c
_
_
ft
.s.
C
2
_3
t _
OS
<N
vd
s
.60
WN (Bdl) X
m -f es -< oo
t- vfiin^es
S S tS S tS
i -f rn es
WN (Bdl)
w-> ^<S-oc
o t- vowi^
-
M
C__LC__UCCLCCLCE^CUi
S 23 23 S i8 $ S
r~ d <n ^ en es
i"N (Bdl) X
Chapter 6: Rigidity as Function of Joint Geometrical Parameters
100
12E6-
10E6-
J 8.E6-
f 6.E6
*4.E6
2.E6
0.
- " " ,
'
0. 0.2 6.4 6.6
X
d.8
1.0
(a)
p=i.o
8.E6- P=0.75
7.E6-
-.6.E6-
Z
5.E6-I
m
04.E6
*3.E6
2.E6
1.E6
0.
0. 0.2
i 1
0.4 0.6
t
0.8
i
1.0
(c)
10E6-
8.E6
m
n.
6.E6-
--4.E6
2.E6
0.
p=1.0
P=0-75
y=15.22 M - 6 4
r
B=0-52
B=0.41
P=0-28
i
0.2
i 1
0.4 0.6
x
(b)
6.E6-, P=10
1
P=0-75
5.E6
4.E6
03
_=3.E6-
2.E6
1.E6-
0.
J J g Z : r_ Y=22.82
P_041
P=0-28
i
0.8
-1
1.0
0.
I
0.2
0.6
(d)
0.8
-1
1.0
4.E6
3.E6-
m
2.E6
1.E6-
0.
P=1.0 -
P=0.75-
P=0-64-
p=0-52"
P=0-41-
P=0-28-
Y=28.83
0.
~02 04 06 0?8 LO
(e)
Figure 6.4. Relationship between . and IPB stiffness. a)y= 13.36, b)y= 15.22, c)y= 17.12,. d)y = 22.82,
e)y= 28.83
Chapter 6: Rigidity as Function of Joint Geometrical Parameters
101
It is seen in Figures 6.2 and 6.3 that inplane bending stiffness of T-joints increases as P or y
ratios increase, or the joints with large diameter braces possess larger stiffness than those
with small diameter braces. It is also noted that thicker chords result in stiffer joints. These
effects should not be mistaken with the overall effect of a stiff chord on the moment
distribution in a structure. A thick chord causes high stiffness as regard to the global
behaviour of a structure, however it includes a high local stiffness which enhance the
stiffness of the joint.
Furthermore, the curves showing the relationships between &
IPB
and p for different y ratios,
in Figure 6.2, are almost equally spaced, whereas they correspond to the values of y =
13.36, 15.22, 17.12, 22.82, and 28.83 (Figure 6.2c). This indicates that IPB stiffness is
most sensitive to y ratio in the lower range of this parameter. In other words, the stiffness
of the T-joints that have thick chord walls has a large gradient with respect to y ratio.
The highest variation of inplane bending stiffness with respect to p ratio also occurs for the
joints with large diameter braces. This is seen in Figure 6.3c where the distance between
the stiffness curves, suddenly increases at p = 0.75. However, comparing Figure 6.2c and
6.3c, it is noted that k
im
is more sensitive to the variation of y ratio than p.
B y the knowledge of this author, the effect of x ratio on the stiffness of tubular joints has
not been considered until now. The diagrams in Figure 6.4 show that there is a rather
significant effect from x ratio on the inplane bending stiffness of T-joints. It is not,
however, as much as the effects of p and y ratios. The average sum of the ratios of the
stiffness values for the joints with x=l and x=0.2 based on the results reported in Tables
6.3c and Table 6.3a is 1.40. This indicates an average difference of 4 0 % in IPB stiffness of
a T-joint when x ratio is varied from 0.2 to 1.
Chapter 6: Rigidity as Function of Joint Geometrical Parameters
102
6.2.2 Out of plane bending mode
Tables 6.4a to 6.4c report the results of the Finite Element analyses for out of plane bending
mode. Figures 6.5 to 6.7 also illustrate the same results as relationships between a joint non-
dimensional parameter and its stiffness.
Table 6.4a. Calculated natural frequencies and stiffness of the T-jo
joint
Tla
T2a
T3a
T4a

T5a
T6a
T7a
T8a
T9a
TlOa
Tlla
NP(Hz)
86.66
48.53
31.64
83.58
44.31 ]
30.34
f
81.60
41.10
27.11
*(Nm)
.316E7
484E6
.200E6
.212E7
.274E6
.107E6
".148E7
164E6
.516E5
joint NF*(Hz)
T12a
_.__^
T14a
T15a
40.60
56.61
65.30
38.05
T16a j 52.65
T17a
T18a
T19a
T20a
T21a
T22a
61.33
34.32
48.02
56.73
85.14
63.34
*(Nm)
.304E6
.740E6
.119E7
.174E6
.452E6
.753E6
.935E5
.256E6
.446E6
.259E7
.951E6
int FE models, OPB (x = P.2)
joint
T23a
24a
T25a
T26a
T27a
T28a
T29a
T30a
T31a
T32a
NF*(Hz)
54.64
46.89
39.54
31.65
79.12
53.35 "
44.61
37.45
31.31
25.16
Jfc(Nm)
.593E6
.362E6
.227E6
.146E6
.102E7
.298E6
.171E6
' .973E5
.563E5
.290E5
*NF: Natural Frequency
Table 6.4b. Calculated natural frequencies and stiffness of the T-joint FE models, OPB (x = 0.5)
joint
Tib
T2b
T3b
T4b
T5b
T6b
T7b
T8b
T9b
TlOb
Tllb
NF*(Hz)
67.61

37.71
26.86
65.11
29.66
24.86
65.34
27.86
21.25
Jfc(Nm)
4.11E6

5.56E5
2.12E5
2.87E6
2.27E5
1.20E5
2.13E6
1.51E5
5.51E4
joint
T12b
T13b
T14b
T15b
T16b
T17b
T18b
T19b
T20b
T21b
T22b
NF*(Hz)
32.47
43.42
50.13
29.35
39.45
46.16
25.62
35.18
41.83
67.63
48.11
*(Nm)
3.38E5
8.65E5
1.45E6
1.99E5
5.20E5 __
8.67E5 ~~
1.05E5
2.98E5
5.26E5
3.72E6
1.13E6
joint
T23b
T24b
T25b
T26b
T27b
T28b
T29b
T30b
T31b
T32b
NF*(Hz)
41.38
JKNm)
6.76E6
35.77 4.21E5
30.97 2.63E5
25.66 1.54E5
61.36 1.45E6
~ 38.76 " 3.47E5
32.17 1.95E5
27.02
22.84
19.09
1.14E5
6.37E4
3.26E4
*NF: Natural Frequency
Table 6.4c. Calculated natural frequencies and stiffness of the T-joint FE models, OPB(x
joint
Tic
T2c
T3c
T4c
T5c
T6c
T7c
T8c
T9c
TlOc
Tile
NF*(Hz)
56.47
30.53

23.30
53.98
26.93
22.58
53.16
23.65
17.54
*(Nm)
.591E7
.634E6

.245E6
.398E7
.366E6
.175E6
.270E7
.197E6
.646E5
joint
T12c
T13c
T14c
T15c
T16c
T17c
T18c
T19c
T20c
T21c
T22c
NF*(Hz)
27.15
35.07
39.58
23.99
31.42
35.92
20.58
27.63
32.78
55.23
37.61
*(Nm)_
.422E6
.103E7
.162E7
.232E6
.640E6
.101E7
.122E6
.356E6
.592E6
.473E7
.119E7
joint NF*(Hz)
T23c 33.16
T24c i 28.92
T25c J 25.56
T26c 22.83
T27c 50.08
T28c i 30.10
T29c ' 25.03
T30c 21.16
T31c 18.12
T32c 15.74
= 1.0)
JKNm)
.818E6
499E6
.319E6
.201E6
.184E7
.391E6
.212E6
.128E6
.718E5
.387E5
*NF: Natural Frequency
103
WN '(BdO) X
ll
II
es
ll
LUN '(BdO) X
53
CO.
e
_
_ _
as
vd
3
.00
s $ s s
<* en H
UJN '(BdO) X
O t > voire* es
G__L^__LC0^CCL^__L0__L
*3
o
H
-oh-n
oooo
H#
o
J"
!?
1 1"
III
< '..
:4
l~
"es
-*
es
_,
' ' 1
/ / III
IS PS iS S S
d
vd <ri * en es -J
WN '(BdO) X
i :
r -t es oo
O t- v o m * es
"if
<
?
ddd<
?
CQ.aO>CQLCQ.aCLCQ.
rr
rS
-t-
es
:/;
//I
/ iit
iri
S S tS s
^r en <S <
"JN '(BdO) X
S iS tS s
u"N '(Bd6j X
Chapter 6: Rigidity as Function of Joint Geometrical Parameters
105
(C)
2.5E6-, 0=1.0
2.E6-
m
a.
1.5E6-
1.E6-
0.
J=0.75 - -
J=0.64
5=0.52
5=0.41 -
5=0.28
5.E6-1
4.E6-
E
z
^ 3 . E 6 -
m
o
W
2.E6-
__
1.E6-I
0 -
0
0=1.0
P=0-75
P = 0 . 6 4 - Y=15.22
P=0.52
B=0.41 ^ ^
0=0.28 - ^
L _ ~ -
-
0.2 0.4 0.6 0.8
i
1.0
X
(b)
1
P=0-75
.2.E6-
m
o.
o
1.E6
0=0.64 v=22.82
0=0.52
0=0.41
0=0.28
0. 0.2 0 4 06 Oe? T.O
(d)
7=28.83
0. 0.2
~0A
i
0.6 0.8
i
1.0
(e)
Figure 6.7. Relationship between x and OPB stiffness. a)y= 13.36, b)y= 15.22, c)y= 17.12,. d) y =
22.82, e)y = 28.83
Chapter 6: Rigidity as Function of Joint Geometrical Parameters 106
The relationship between out of plane bending stiffness and non-dimensional parameters of
T-joints are similar to the IPB mode. The O P B stiffness has an ascending trend with respect
to p and x ratios and a descending trend with y ratio. The T-joints with large brace
diameters, or low chord diameter to thickness ratios have a relatively high stiffness.
It is noted in Figures 6.5 that out of plane bending stiffness becomes most sensitive to y
ratio in the lower range of this parameter. It is also seen in Figure 6.6 that O P B stiffness of
T-joints becomes significandy sensitive to P for the joints with high (J ratios. In comparison
with IPB mode, the O P B stiffness varies more drastically for large ps (comparing Figure
6.3c and 6.6d).
It can be said that the O P B stiffness of the T-joints with large brace diameters, or the
stiffness of those which have small chord diameter to thickness ratios should be determined
more carefully.
There is also a variation observed for O P B stiffness with respect to x ratio. It is seen in
Figures 6.7 that the effect of x parameter should be necessarily considered on the O P B
stiffness. From the results shown in Tables 6.4a and 6.4c, the average sum of the ratios of
the stiffness values for the joints with x=l and x=0.2 is 1.44. This indicates an average
difference of 4 4 % in O P B stiffness of a T-joint when x ratio is varied from 0.2 to 1.
6.2.3 Axial deformation of brace
The results of Finite Element analyses due to the axial deformation of brace are reported in
Tables 6.5a to 6.5c. The graphical presentation of the same results are given in Figures 6.8
to 6.10.
Table 6.5a. Calculated natural frequencies and stiffness of the T-joint FE models, axial deformation
ft = P.2)
joint
Tl
T2
T3
T4
T5
T6
. _T7 ,
T8
T9
TIO
Til
NF*(Hz)
213.84
225.18
235.08
213.99
226.48
236.17
215.67
227.26
237.33
fc(N/m)
3.31E8
8.19E7
3.87E7
__ 2.38E8
5.71E7
3.08E7
1.80E8
4.19E7
1.62E7
joint
T12
T13
T14
T15
T16
T17
JL
18
T19
T20
T21
T22
NF*(Hz)
229.37
221.50
217.95
230.58
222.74
219.21
231.55
223.62
220.46
214.30
218.62
*(N/m)
5.31E7
1.13E8
1.51E8
3.74E7
7.91E7
1.06E8
2.60E7
5.29E7
7.67E7
2.79E8
1.29E8
joint
T23
T24 "
T25
T26
T27
T28
T29
T30
T31
T32
T31
NF*(Hz)
221.53
225.59
230.00
236.17
215.58
220.93
" 224.29
227.79
231.91
237.91
*(N/m)
9.02E7
6.95E7
4.54E7
3.28E7
1.03E8
" 5.62E7
4.04E7
2.79E7
1.90E7
1.33E7
*NF: Natural Frequency
Chapter 6: Rigidity as Function of Joint Geometrical Parameters
107
Table 6.5b. Calculated natural frequencies and stiffness of the T-joint FE models, axial deformation
<x-0.5)
joint
Tl
T2
T3
T4
T5
T6
T7
T8
T9
TIO
Til
NF(Hz)
179.67
201.91
220.24
180.07

202.14
220.15
180.27
202.49
220.63
*(N/m)
1.22E9
2.05E8
1.05E8
6.82E8
1.264E8
5.41E7
4.78E8
8.41E7
3.56E7
joint
T12
T13
T14
T15
T16
T17
T18
T19
T20
T21
T22
NF*(Hz)
209.82
195.79
190.02
209.86
196.15
190.59
209.36
196.12
190.90
180.22
190.29
*(N/m)
1.41E8
3.06E8
4.35E8
8.96E7
1.90E8
1 2.65E8
5.46E7
1.12E8
1.63E8
9.40E8
3.35E8
joint
T23
T24
T26
T27
T28
T29
T30
T31
T32
NF*(Hz)
195.96
202.34
209.80
220.83
181.93
190.76
195.64
201.31
208.25
219.70
*(N/m)
2.40E8
1.67E8
1.14E8
8.88E7
.307E9
110E9
.752E8
.513E8
.345E8
.255E8
*NF: Natural Frequency
Table 6.5c. Calculated natural frequencies and stiffness of the T-joint FE models, axial deformation
(x = 1.0)
joint
Tl
T2
T3
T4
T5
T6
T7
T8
T9
T10
Til
NF*(Hz)
150.56
176.14
198.46
150.27
174.26
197.69
150.21
172.08
195.89
*(N/m)
1.99E9
3.42E8

1 27E8
1.09E9
1.52E8
| 7.12E7
6.62E8
8.99E7
4.34E7
joint
T12
T13
T14
T15
T16
T17
T18
T19
T20
T21
T22
NF*(Hz)
184.71
167.15
161.24
183.26
166.65
161.11
180.82
165.59
159.95
150.67
161.22
fc(N/m)
2.07E8
4.62E8
6.57E8
1.20E8
2.62E8
3.76E8
6.70E7
1.44E8
2.15E8
1.46E9
4.85E8
joint
T23
T24
T25
T26
T27
T28
T29
T30
T31
T32
T31
NF*(Hz)
166.86
174.46
183.90
198.43
150.83
159.26
164.24
170.16
177-85
192.00
J.(N/m)
3.34E8
2.29E8
1.54E8
1.03E8
4.03E8
1.42E8
8.92E7
6.15E7
4.22E7
2.83E7
*NF: Natural Frequency
o
c
.2
c

>
(0
CD
CD
id
o
<d
-\ h
vo es <s es en
tn es -H oo oo
en id r^ es oo
M H H N nl
JLiLJLjL?.
T
00
d
*> en.
o
a I I a a s a a s a
N ^ ^ - . * ei
w/N *(|D1*) X
es

voesesesm
m cs < oo oo
rn <ri i~ es oo
^ * M es es
JLJLJLiLiL
r ~r
ON os o
PJ w u
* es _;
"5 s s a"
OO VO ^ <S
W / N '(IDIXD)
oo
d
vo __
d
en
X ii- JL X ^
r -r
a 3
IT) <*
en
a a
"1 es
es
oo
UJ
n
oo
-J
S
"J/N '(IDIXD)
oo
d
io oa.
d
es
o oooo
m * es oo
o r-v in * es
l?
de
_
<
_
dd
Ov vo m
UJ/N'(|01XD)>|
o
en
es
-t
e-j
es >-
a s a s s s s o
es _;
UU/N*(|DIXD)
' 1
. 1 .
1 1 1
: 11
m^e>l-o<
r~vom-r c
*rr'T? *i?-if'W f
^3-^_LCQ-CQ-C__LGC
CS
1
/
1 1 f
/ ' 1
\ / >
i /
:
i'h
/ , * * i
/ .
f
/
/ - '' '
>' < ll
r / ' if
.' / / II
i i
-
-
-es
-es
es?_
n "1 es "I - id
m
m
cs
^/N'(|0!XD)
Chapter 6: Rigidity as Function of Joint Geometrical Parameters
110
2.E9-I
g 16E9
O 1.2E9
X
53

8.E8
4.E8-
0.
.0 0.2
r -
0.4
0J
x (t/T)
(a)
6 0.8
i
1.0
1.5E9
1.2E9
C?9.E8
03
r-C
X
10

6.E8
3.E8
P=1.0
P=0-75
P=0-64
0=0-52
B=0-41
P=0-28
.0 0.2
0.4 0.6
i (VT)
(b)
0.8 1.0
1.4E9-
1.2E9-
s
^l.E9.
f6.ES.
e_
W 4.E8
2.E8
0.
P=1.0
P=0-75
P=0.64
P=0-52
P=0-41
P=0-28
0.2
0.4 0.6
x (VT)
(C)
0.8 1.0
7.5E&H
6.E8-
*4.5E8
<0
X 3.E8-]
CO
1.5E8-I
0.
p=i.o
B=0-75
P=0.64 T=22.82
P=0.52
P=0.41
P=0.28
0.2
0.4 0.6
x (VT)
(d)
0.8 1.0
5.E81
S
4
E8
^ 3 . E 8 H
X
oa

2.E8
1.E8-
B=1.0
P=0-75
0=0.64 Y=28.83
p=0-52
P=0.41
P=0.28
i 1 1 r-
0 2 0.4 0.6 0.8
T (t/T)
i
1.0
(e)
Figure 6.1P. Relationship between x and axial stiffness, a)y= 13.36, b)y= 15.22, c) y = 17.12,. d)y-
22.82, e)y= 28.83
Chapter 6: Rigidity as Function of Joint Geometrical Parameters
111
The main observation from the results shown above in Figures 6.8 to 6.10, is the effect of x
ratio on the axial stiffness of a T-joint. The average sum of the ratios of the stiffness values
shown in Tables 6.5a and 6.5c for the joints with x=l and x=0.2 is 3.32, which indicates a
mean change of 2 3 2 % in the axial stiffness of a joint when the x ratio varies from 0.2 to 1. It
is also seen in Figures 6.9 and 6.10 that axial stiffness of T-joints becomes more sensitive to
P in the higher range of this parameter.
Further, it is noted that a large brace diameter or a small chord diameter to thickness ratio
result in a high axial stiffness. Figure 6.8c shows that for the differences between y ratios
equal to 1.86, 1.9, 5.7, and 6.01, the curves of relationships between axial stiffness and P
are almost equally spaced. This indicates that axial stiffness becomes significantly sensitive
to y in the lower range of this parameter.
As a result, x ratio should be definitely considered when the axial stiffness of a T-joint is
calculated. Furthermore, more attention should be paid to determine the axial stiffness of
the joints which have large P or small y ratios .
6.3 Derivation of stiffness equations
The joint stiffness values which were calculated by using the results of the Finite Element
analyses, indicate a clear relationship with non-dimensional geometrical parameters of the
joints. Therefore, a formula may be established for each m o d e of deformation. A power
regression method is used in this study to derive the stiffness formulae. The results of such
an analysis yields Equations (6.2), (6.3), and (6.4) for IPB, O P B , and axial brace
deformation respectively, as reported in Table 6.6. This table shows also the correlation
factor of each equation (R
2
). R
2
values are very close to one indicating that there is a very
good correlation between the data values used to establish the formulae.
Table 6.6. T-joint stiffness formulae and the corresponding correlation factors
k = o 169i_D
3
B
(Yt/122+155)
Y
(p/2I7
"
182)
x
025
/?
2
= 0.996 (6.2)
k
OPB
= 0.088__D
3
p
(ir/86 4+0 31)
Y
(187P
'
3 27)
x
212 R2 =
"
997
k = 2 712__D3
(
"
Y / 2 7 3 + 0 - 6 t )
Y
( 1 3 8 M )
x
0 9 1
/?
2
= 0.985 (6.4)
The proposed formulae in this study are valid in the following ranges of non-dimensional
parameters:
0.3<P<0.9, 13<Y<30, and 0.2<x<1.0 (6.5)
Chapter 6: Rigidity as Function of Joint Geometrical Parameters
112
The predicted stiffness values by the developed equations are shown in Figure 6.11 against
the data from F E results to check the performance of the curve fittings. T o save space only
the comparison is made for the joints that have a x ratio equal to 0.5. It is seen that the
developed equations produce a good estimation for the stiffness values, especially for IPB
and O P B modes. The maximum curve fitting errors for IPB, O P B , and axial modes are
24.9%, 18.3% and 28.6% respectively. These are the maximum difference between the
predicted and calculated stiffness results for each deformation mode.
9.E6-,
8.E6
p7.E6,
Z
6.E6
CO
__ 5.E6
^ 4 . E 6 -
3.E6-
2.E6
1.E6
0.
T = 0.5
measured -y==15.22
-- predicted
y=13.36
7=28.83
I r I 1 I I I I TT T T I I I I I I | 1 TT I T T T T I ' I I I I T T I T I I |
0.2 0.4 0.6 0.8 1.0
P
a) Inplane bending mode
5..E6
n
4.E6
3.E6-
m
a.
o
2.E6
1.E6-
b) Out of plane bending mode
1.4E91
1.2E9
. 1.E9
O 8.E8-
o
x
O 6.E8
4.E8
2.E8
0.
0.2 0.4
0.6
P
0.8
7=28.83
- i
1.0
c) Axial deformation of brace
Figure 6.11. Predicted values against data used for curve fitting
To have a better picture of the relationships between joint stiffness and non-dimensional
parameters, the three dimensional plots of the stiffness equations adopted herein are made
and shown in Figures 6.12 to 6.14. These plots include only the results of the joints with x =
0.5. The absolute errors of the curve fittings are also reported in these diagrams. It is seen
Chapter 6: Rigidity as Function of Joint Geometrical Parameters 773
that the stiffness of the T-joints under IPB, OPB and axial loading have increasing trends
with P ratio, but they decrease w h e n Y ratio is going up.
Figure 6.12. Model adopted for IPB joint stiffness (eq (6.2)} with the absolute errors of curve fitting
Figure 6.13. Model adopted for OPB joint stiffness (eq (6.3)) with the absolute errors of curve fitting
Chapter 6: Rigidity as Function of Joint Geometrical Parameters
114
Figure 6.14. Model adopted for joint axial stiffness {eq (6.4)} with the absolute errors of curve fitting
6.4 Comparison with other stiffness equations
The main objective of establishing the parametric formulae is to be able to predict the
stiffness of a joint without having to carry out detailed analysis, either model testing or F E
analysis. It is usually preferred in practice to use simplified analysis and design methods. A
parametric formula can be employed in a structural analysis computer program to
incorporate the joint stiffness automatically in the analysis. Nevertheless, it is required to
conduct detailed studies in some special cases. The stiffness consideration of joints is
especially important in the reappraisal of existing offshore structures.
To examine the accuracy of the equations developed herein, a comparison is made between
the predicted results of the equations in this study and those reported by other researchers.
It will be shown that the method used in this study can produce compatible results with
other investigations. However, this study can be distinguished among other works on the
stiffness of tubular joints, as the dynamic approach was employed here.
Chapter 6: Rigidity as Function of Joint Geometrical Parameters 775
The literature survey carried out in this study shows seven sets of formulae for the
calculation of joint stiffness or flexibility using non-dimensional geometrical parameters.
The formulae are by:
DNV (1977),
U E G (1985),
Efthymiou (1985),
Fessler (1986),
Ueda etal. (1990),
Chen B.etal. (1990), and
Kohoutek and Hoshyari (1992).
Each set of formulae is described in the following sections.
6.4.1 DNV (1977)
D N V (Norwegian Rules for the Design, Construction and Inspection of Offshore Structures,
1977), has recommended two formulae for the stiffness of T-joints, as follows:
k
m
= 0.43__
3
(i-.01)
(235
-
,5fJ)
(6.6)
Y
k
OPB
= 0.0016_i_?
3
(215-P)(--.02)
(245
-
,6p)
(6.7)
y
The range of validity for Equations (6.4) and (6.5) are given as:
0.33<p<0.8, and 10<y<30. (6.8)
DNV Formulae are probably the first set of equations proposed for the stiffness of tubular
joints.
6.4.2 UEG (1985)
The Underwater Engineering Group (UEG, 1985) has surveyed the experimental data to
the date of publishing and recommended a set of formulae which has been taken from
Professor Fessler at Nottingham University. There have been only two sets of formulae
available to U E G in 1985, DNV's and Fessler's. The recommended formulae by U E G are:
Chapter 6: Rigidity as Function of Joint Geometrical Parameters
116
k^M = E D y
2
'
3
exp(3.3p) sin"
2
G/2.3 (6.9)
kjp
B
= E D
3
y -
h65
exp(4.6p) sin^O/H 1 (6.10)
k
OPB
= E D
3
y
2
'
5
exp(3.7P) sin"
2
e/48.1 (6.11)
No validity range is given for the UEG formulae.
6.4.3 Efthymiou (1985)
A set of equations were proposed by Efthymiou (1985) for inplane and axial stiffness of T
and Y-joints. The work is based on Finite Element analysis and then calculation of joint
stiffness using equations (6.12).
M P
k
iPB=~jj-
md
^ _ = y (6-12)
in which M and P are the bending moment and the axial force of the brace, respectively. G
and 8 are the local rotation and axial deformation of the joint. The recommended equations
by Efthymiou (1985) are:
k
IPB
= 0.163 ED
3
y
lM
p("5
+
yi25)
sm
-<P+o.4)
e
(6.13)
koPB = 0.288 E D
3
Y
(07(0.55-P)2-2.20) 02.12 ^-(0+1.3)0
(6 14)
with a range of validity imposed as:
0.3<p<0.8, and 10<Y<30. (6.15)
6.4.4 Fessler (1986)
Fessler has derived a set of formulae for the stiffness of T and Y-joints, besides the ones
published by U E G (1985), based on testing a number of araldite models. The formulae are:
kaxua = ED y
215
(1-P)-
13
sin-
219
e/1.95 (6.16)
k
WB
= ED
3
Y'
1
'
73
exp(4.52P) sin
122
9/134 (6.17)
k
OPB
= ED
3
Y"
2
'
20
exp(3.85P) sin-
216
9/85.5 (6.18)
A range of validity is imposed on the formulae by Fessler as:
0.3<P<0.8, and 10<Y<20.
(6.19)
Chapter 6: Rigidity as Function of Joint Geometrical Parameters
117
6.4.5 Ueda et al. (1990)
The paper by Ueda et al. was discussed earlier in Chapter 2. The equations developed by
Ueda using Finite Element analysis are:
k^ = 3.195 ED y
2
-
3
P
12
sin-
2
9 (6.20)
k
IPB
= 0.237 ED
3
j-
1
-
7
P
22
sin^e (6.21)
No validity range is given for the Ueda's formulae.
6.4.6 Chen B. et al. (1990)
The formulae developed by Chen B. et al. are originally for a symmetrical K-joint. However,
it is recommended by Chen that the formulae can also be used for axial and inplane bending
of T and Y-joints. The Chen's formulae are very similar to those recommended by U E G .
The formulae by Chen are:
_ = ED Y'
2
V'
25
" sin"
202
6 / 4.71 (6.22)
k^B = ED
3
y-
168
_
4
'
58(i
sin"
1
-
25
e/169 (6.23)
The validity range of the above equations are given as:
0.3 < p < 0.8,7.5 < Y < 35, and 30 < 9 < 90 (6.24)
6.4.7 Kohoutek and Hoshyari (1992)
A formula has been suggested for IPB rigidity factor of tubular T-joints by Kohoutek
(1992). The formula is actually based on the experimental results of this study described in
Chapter 4. The proposed equation is:
T = 1.307 - 0.819 p - 0.0130 Y+ 0.488 P
2
(6.25)
T varies between 0 and 1. A hinged or a fixed support is assumed to have a T = 0 or 1,
respectively.
6.4.8 Graphical Comparison of different stiffness formulae
Figures 6.15, 6.16 and 6.17 show the comparisons between the performance of the
equations of joint stiffness developed by other researchers and those developed herein for
axial, IPB and O P B modes. The significant difference between the formulae developed in
this study and other equations is the inclusion of thickness ratio or x parameter. Therefore,
Chapter 6: Rigidity as Function of Joint Geometrical Parameters
118
using other equations gives the same results for different x values, while the developed
equations in this study predict different stiffness values with respect to x. This leads to three
samples on a line in Figures 6.15 to 6.17 corresponding to x =0.2, 0.5 and 1 which makes
the comparisons inevitably inaccurate to some extent
A discrepancy ratio, R, is defined as the ratio between the results of other equations and the
formulae of this study. The geometrical parameters, stiffness and R values of the joints
which are used in Figures 6.15 to 6.17 are reported in Tables 6.7a, 6.7b and 6.7c. The
following findings can be obtained from the comparisons of different formulae:
Table 6.7a. Geometrical parameters and stiffness of the joints of Figures 6.15 to 6.17 calculated using the
formulae by others
Equation P IPB tf OPB fit axial /?*
Fessler 0.33
0.8
0.33
0.8
13.36
13.36
20
20
.787E6
.659E7
.392E6
.328E7
1.18
1.36
1.18
1.22
.292E6
.179E7
.120E6
.735E6
1.06
1.01
1.17
0.82
f
.144e9
.692E9
.603E8
.290E9
135
1.43
1.21
1.22
DNV
Efthymiou
Ueda
0.33
0.8
0.33
0.8
0.33
0.8
0.33
0.8
0.33
0.8
0.33
0.8
13.36
13.36
20
20
13.36
13.36
20
20
13.36
13.36
20
20
.707E6
.486E7
.289E6
.279E7
.600E6
.483E7
.316E6
.267E7
.530E6
.372E7
.267E6
.187E7
1.06
1.00
0.87
1.04
0.90
1.00
0.95
0.99
0.79
0.77
0.81
0.70
.271E6
.151E7
.848E5
.744E6
.210E6
.141E7
.875E5
.590E6




0.98
0.85
0.82
0.84
0.76
0.79
0.85
0.67












.953E8
.276E9
.377E8
.109E9








.89
.57
.75
.46
R : discrepancy ratio = ifcothers/fahis study
Table 6.7b. Geometrical parameters and stiffness of the joints of Figures 6.15 to 6.17 measured by others
Equation P Y IPB # O P B R* axial R*
Tebbett
McDermott
0.33
0.54
0.59
0.92
0.53
20
20
32
20
24
.395E6
.673E6
.491E6


1.19
0.68
0.85


* R: discrepancy ratio = fcxheis/tthis study
.081E6
.095E6
.180E6
0.79
0.76
.487E8
.762E8
.285E9
0.97
0.79
0.77
Chapter 6: Rigidity as Function of Joint Geometrical Parameters
119
Table 6.7c. Geometrical parameters and stiffness of the joints of Figures 6.15 to 6.17 calculated by the
formulae developed in this study
&
0.33
0.8
0.33
0.8
0.54
0.59
0.92
0.53
Y
13.36
13.36
20
20
20
32
20
24
X
.2
.5
.97
.2
.5
.97
.2
.5
.97
.2
.5
.97
.2
.5
.97
.2
.5
.97
.2
.5
.97
.2
.5
.97
IPB
.551E6
.669E6
.745E6
.389E7
.485E7
.566E7
.278E6
.331E6
.359E6
.216E7
.269E7
.312E7
.810E6
.988E6
.111E7
.477E6
.576E6
.637E6
.318E7
.398E7
.467E7
.579E6
.702E6
.781E6
OPB
.227E6
.276E6
.317E6
.147E7
.178E7
.205E7
.848E5
.103E6
.119E6
.731E6
.887E6
.102E7
.211E6
.256E6
.294E6
.103E6
.125E6
.144E6
.133E7
.161E7
.185E7
.137E6
.166E6
.191E6
axial
.565E8
.107E9
.142E9
.219E9
.484E9
.831E9
.265E8
.500E8
.668E8
.108E9
.238E9
.408E9
.467E8
.962E8
.148E9
.245E8
.514E8
.809E9
.162E9
.368E9
.656E9
.329E8
.675E8
.106E9
Axial rigidity of brace
The predictions of the axial rigidity equation adopted here are within 1 0 0 % of the values
calculated by Fessler and Ueda's equations and also some experimental results by Tebbett
and McDermott. It should be noted that the 1 0 0 % difference in case of Ueda's equation can
not be counted very reliable because Ueda's equation is based on only 7 samples.
The following points are observed in Figure 6.15:
1) A good agreement is found when the axial stiffness is compared with the experimental
results of Tebbett (1982). There are only three samples from Tebbett, which their x ratios
are not quoted in Tebbett's paper. The average difference with Tebbett's equation obtained
for the joint parameters examined here is 16%.
2) The results of the equation of this study and Ueda's equation have a close relationship for
the joints with low P ratios.
Chapter 6: Rigidity as Function of Joint Geometrical Parameters 120
3) Fessler's results show a good relationship with the results of the formula developed here,
except for the joint with P = 0.92; perhaps because Fessler's equation is not valid in this
range.
4) The variation of axial stiffness with respect to x has been obtained in this study to follow
the equation: pH'
27
-
3
*
0
-
6
*)^-
91
- j ^ _ jmpijes
m a t m e
j^uits
0
f the other stiffness equations
which do not include x ratio are susceptible to a large variations as shown in Figure 6.15.
joint parameters for data points
column p y
1
2
3
4
5
6
0.33
0.54
0.33
0.8
0.92
0.8
20
20
13.36
20
20
13.36
***** Fessler (1986)
00000 Ueda (1990)
Tebbett, experimantal (1982)
1C? 10"
Kaxiai , developed equation
Figure 6.15. Comparison of different equations for axial joint stiffness, the arrows specify the range of
axial stiffness when X varies from P.2 to P.97
Inplane bending mode (IPB)
There is generally a good agreement between the results of all IPB formulae compared here,
including the developed equation. It is assumed that the IPB formula established here,
produces reasonably accurate stiffness values. The reason for this assumption is the good
correspondence obtained between the test and F E analysis results, reported in Section 5.6.1.
Furthermore, the curve fittings used in constructing the IPB formula showed small errors.
The following points can be deduced from Figure 6.16 about the performance of different
equations:
1) Fessler's equation is more appropriate for the joints with large x values. Fessler's equation
generally predicts high values for the stiffness of a T-joint
2) Ueda's equation is more appropriate for the joints with small x values. Ueda's equation
generally predicts low values for T-joint stiffness and shows the poorest consistency among
the equations compared here. It should be noted that Ueda et al. have considered only 11
joints to derive their equation for IPB stiffness of T-joints (Ueda, 1990).
Chapter 6: Rigidity as Function of Joint Geometrical Parameters
121
3) The results of D N V and Efthymiou's equations are in a very good agreement with those
from the new equation of this study, especially for T-joints with x values around 0.5. DNV
and Efthymiou's equations produce higher stiffness results for small x ratios than the actual
values and lower results for large x values.
}
P=0-8
y=20
}
p=0-8
Y=13.36
***** Fessler (1986)
+++++DNV (1986)
00000 Efthymiou (1986)
DD-IDD Ueda (1987)
O O
0 0 0
Tebbett, experimental (1982)-Table 6.7c
KIPB x10 , developed equation
Figure 6.16. Comparison of different equations for IPB joint stiffness
4) The variation of IPB stiffness with respect to x ratio is seen as p^^+i-ss^o.* _ j ^ ^
make a difference of approximately 40% in the joint stiffness of a joint which has a x = 1
compared to when x = 0.2.
5) The Tebbett's results shown in Figure 6.16 are taken from experiment and are also
reported in Table 6.7b. It is seen that the experimental results may be well predicted for the
practical design purposes. However, a realistic comparison is not possible since the x ratios
used in Tebbett's work are not reported.
6) The calculated stiffness values should be converted to rigidity factors in order to have a
comparison with Kohoutek's equation. The rigidity factors of the joints in Table 6.7a are
calculated using Equation (3.20) suggested in Chapter 3 and Kohoutek's equation. The
results are shown in Table 6.8.
Chapter 6: Rigidity as Function of Joint Geometrical Parameters
122
Table 6.8. Comparison of the results of proposed equation with Kohoutek's equation (IPB)
p
0.33
0.8
0.33
0.8
Y
13.36
13.36
20
20
*IPB
.787E6
.659E7
.392E6
.328E7
v
b
0.67
0.48
0.58
0.43
r
0.92
0.79
0.83
0.70
*t
1.37
1.65
1.43
1.63
R : discrepancy ratio = Vb/T
As the theoretical basis of Kohoutek's equation is different from the joint model adopted in
this thesis (stated in Chapter 2), the comparison of the two equations may not be very
realistic. The maximum difference of rigidity factors calculated by the two formulae is 6 5 %
for the batch of parameters selected here. However, the similar trend is observed between
the results of the two equations, indicating that a closer agreement can be obtained by
adopting a different conversion rule to convert the joint stiffness to rigidity factor.
Out of plane bending mode (OPB)
The comparison of the out of plane bending stiffness values does not include Ueda's
equations since they do not cover this mode of deformation.
The proposed O P B equation herein predicts higher stiffness values than the experiments of
this study for the joints with large p ratios. The poor agreement of F E and experimental
results, as discussed in Section 5.6.1, is due to the poor fixity at the support assembly of the
joints that have large brace diameters. The F E method, however, is believed to produce
acceptable results as it predicted the IPB stiffness with a high accuracy. The following
conclusions can be made by comparing the different formulae of O P B stiffness:
1) The correlation among different formulae of OPB stiffness compared here is generally
good. However, Fessler's equation is more in agreement with the new formula developed in
this study, compared to D N V and Efthymiou's. Efthymiou's equation shows the poorest
comparison. Fessler's equation is seen to be more appropriate for the joints with x ratios
around 0.5, whereas D N V and Efthymiou's equations are appropriate for the joints with
small x values.
2) It is seen that the predicted stiffness values for the joints with p = 0.8 are also in a close
relationship to the values resulted from other equations. This supports the accuracy of the
F E analyses of the joints with large p ratios, since the developed equations herein are based
on the F E analysis results.
3) The results of the established equations here are in a good agreement with the
experimental results of Tebbett (1982) and McDermott (1979). The maximum difference is
2 4 % as shown in Table 6.7b.
Chapter 6: Rigidity as Function of Joint Geometrical Parameters
123
4) The variation of O P B stiffness with respect to x is up to 4 0 % ; the joints with higher x
values produce larger stiffness.
M
c
o
? :
o
ID
o
x

o
__:
0.1-
parameters shown
in Table 6.7c
iN.8
X
Y=20
} P=0-
1*13.36
*****
I l i t I-
00000
mnm
ooooo
Fessler (1986)
DNV (1986)
Efthymiou (1986)
Tebbett, experimental (1982)
McDermott, experimental (1979)
T 1I I I |
S 7
1
KQPB
X 1
, developed equation
Figure 6.17. Comparison of different equations for OPB joint stiffness
6.5 Summary
The results of a parametric study carried out based on the FE analyses to establish a set of
equations for prediction of the stiffness of tubular T-joints were reported in this chapter.
There are 90 different geometrical parameters which were covered by the F E models. A
chord diameter to thickness ratio of 6.85 was adopted for all models. The non-dimensional
parameters of the F E models considered in the analyses were diameter ratio (P), chord
diameter to thickness ratio (v), and thickness ratio (x). Because only one quarter of a joint
was modelled, three analyses on each model with various boundary conditions were carried
out.
Finite Element results had generally a good correlation with the verification tests, reported
in Chapter 5. Therefore, it is believed that the parametric study based on the F E analyses
produces realistic results. The IPB, O P B and axial stiffnesses calculated for T-joints show
that the stiffer joints are most sensitive to the variation of non-dimensional geometrical
parameters. For the three modes of deformations considered here, the joints with high P and
Chapter 6: Rigidity as Function of Joint Geometrical Parameters
124
low Y ratios are relatively stiff. Consequently, it is recommended that accurate methods to
be used to determine the rigidity of the stiff joints.
Three equations for IPB, O P B and axial stiffness of T-joints were established based on the
F E analysis results. A power regression method was used for curve fittings of the calculated
stiffness values. It was shown that the maximum prediction errors corresponding to IPB,
O P B and axial stiffness values were 24.9%, 18.3%, and 28.6 compared to the F E results,
respectively.
The developed formulae include thickness ratio (x), which to the knowledge of this author,
has not been covered by other formulae reported before. This parameter makes a difference
of up to approximately 4 0 % for O P B and IPB stiffness, and 2 3 0 % for axial deformation of
brace. The significance of x ratio in the stiffness of T-joints, especially for axial loading,
indicates that this parameter should be considered in the prediction of the stiffness of a
tubular joints.
The predicted results by the developed equations in this study were compared with other
equations for a number of joint parameters, and good agreements were generally obtained.
The average of the absolute differences between the results of other formulae and the
equations reported here are 15.4%, 15.2% and 27.4% for IPB, O P B and axial loading
modes, respectively. The variation of joint stiffness due to the change of thickness ratio, x,
was clearly observed in those comparisons. However, x ratio is not usually reported in the
research works which makes the comparison of the different stiffness formulae to some
extent inaccurate. It was shown that the stiffness formulae which do not include the
thickness ratio are consequendy suitable for certain joints with specific x values. Despite
that P=l is highly used in tubular joints of offshore structures, most parametric equations do
not cover the joints with large P ratios.
Chapter 7
DYNAMIC STRESS
CONCENTRATION FACTOR
Chapter 7: Dynamic Stress Concentration Factor
126
7.1 Introduction
Fatigue design of tubular joints for offshore structures by the S-N method requires the
calculation of stress values to be as accurate as possible at the area surrounding the weld.
The complicated geometry of the tubular joints makes the stress distribution spatially
dependent and highly nonlinear, particularly near the weld toe, resulting in the need for an
extensive research into the solution of the stress problem at the intersection of tubes. A
typical stress distribution for a simple tubular T-joint, when the brace is axially loaded, is
shown in Figure 7.1.
Normal stress distribution
away from itersection
Maldistribution of
nominal stress
Saddle Chord
Figure 7.1. Stress distribution at the intersection of a simple T-joint (UEG, 1985)
Methods for investigating the local behaviour of tubular joints are time consuming and
expensive, as referred to in Section 2.3, whereas, in practice, the use of relatively simple
procedures are preferred for analysis and design of structural systems. Because of this
complexity a coefficient which is called S C F (Stress Concentration Factor) is employed for
the analysis of tubular joints. S C F is generally defined as the ratio between the stress at the
intersection area and the nominal stress in the brace, from which S C F may be interpreted as
the normalised stress with respect to the brace nominal stress.
Fatigue design of tubular joints requires the knowledge of the maximum stress range.
M a x i m u m stress occurs at a location called the hot-spot, and there are different definitions
for the hot-spot SCF. S o m e investigators have employed the m a x i m u m principal stress at
the hot-spots in their works. Others have referred to the m a x i m u m stress perpendicular to
the weld line as the hot-spot stress ( U E G , 1985), and there are other methods with different
Brace
Crown
\
tx- -
III! Ill
Chapter 7: Dynamic Stress Concentration Factor
127
definitions (Tebbett, 1984 and Lalani, 1986). The proximity of the hot spot to the weld toe
is a matter of concern too, i.e. h o w far from the weld toe the strain measurements should be
made.
All these factors have imposed an uncertainty on the SCFs determined by different methods.
A parameter which has not been taken into account (to the knowledge of this author) in
regard to the uncertainty of the calculated S C F is dynamic strain reading. As the results of
the experiment and computer analysis showed in this study, the way loads are applied to a
joint in order to detect the strains makes a substantial difference to the calculated values of
stress or strain concentration factors.
This chapter reports the results of a series of dynamic strain reading tests carried out on
tubular T-joints and also the analytical solutions compatible with the experiments.
The stress field at a tubular joint under service load can be described as follows (UEG,
1985):
1) Action of the joints under general loading on the structure causes nominal stresses.
These stresses are due to external loading and are calculated through a global analysis of the
structure. This analysis, which is generally carried out on computer using the stiffness
method, leads to the recovery of internal forces and moments in different structural
members. Nominal stresses can then be calculated using the internal forces and the formulae
from strength of materials. In this regard, the global modelling of a platform structure plays
an important role in the determination of nominal stresses.
2) The geometry of a joint at the intersection changes the internal forces in the members by
developing deformation stresses. The chord wall deforms locally to maintain the
consistency of deformation between the brace end and the chord body, thus introducing
deformation stresses. Variable flexural stiffness between the crown and the saddle results in
a large stress variation at a joint. The deformation stresses are two types, a membrane type
which is constant through the thickness of the wall and a bending type which is assumed to
vary linearly.
3) Notch stresses are developed because of the sudden thickness change introduced by weld
and consequently an additional stiffness that is created at the intersection. This effect comes
as a part of the deformation stress. For example, if the weld is not modelled in the F E
analysis, the result does not contain the notch stresses, and will be called gross deformation
stresses (UEG, 1985). The stress system due to the geometry of a joint affected by the weld
is called local deformation stresses. A n example of the latter type of stress is the
measurements obtained through strain readings by electrical strain gauges on test
specimens. Figure 7.2 shows a typical stress distribution close to the weld toe, in which the
linearity of the bending stress is disturbed (Abel, 1982).
Chapter 7: Dynamic Stress Concentration Factor
128
strain concentration
at weld toe
Chord wall
Figure 7.2. Stress distribution near the weld toe
1
1
4) Residual stresses which are c o m m o n in steel structures also exist in tubular joints. They
are generally created via two sources; one is welding and the other is a lack of fit in
members. The former sometimes causes stresses up to yield point. Residual stresses are
tensile at the position of the weld passes.
7.2 Application of stress concentration factor in fatigue analysis
The following example shows how the stress ranges in tubular joints are calculated using
SCFs. It is taken from U E G (1985) and summarised here. The same procedure is used in
Chapter 8 of this study to predict the fatigue life of tubular joints.
Example
A T-joint is considered for the evaluation of the fatigue damage ratio (nIN).
Figure 7.3 shows the joint geometry and the forces obtained from a computer
analysis.
OPB
IPB
Axial
Chord 457x9.5mm
L=5484mm
(6 far side) 8
Brace 457x6.3mm
2 (4 far side)
t
1 (5 far side)
Figure 7.3. The geometry and forces of the T-joint considered in Example 1 (UEG, 1985)
Chapter 7: Dynamic Stress Concentration Factor
129
Using a set of S C F parametric formulae, the stress concentration factors for the
different deformation modes and the different locations around the intersection
of the brace and the chord can be calculated. The results of such calculations are
shown in Table 7.1, below.
Table 7.1. Hot-spot Stress Concentration Factors for the joint of Example 1
Stress Concentration Factor
Location
1
2
3
4
5
6
7
8
Axial
Chord
7.64
8.62
9.61
8.62
7.64
8.62
9.61
8.62
Brace
5.81
6.43
7.05
6.43
5.81
6.43
7.05
6.43
IPB
Chord
0.00
2.30
3.26
2.30
0.00
2.30
3.26
2.30
Brace
0.00
2.16
3.05
2.16
0.00
2.16
3.05
2.16
OPB
Chord
9.55
6.75
0.00
6.75
9.55
6.75
0.00
6.75
Brace
7.02
4.96
0.00
4.96
7.02
4.96
0.00
4.96
Points 2, 4, 6 and 8 are located on a 45 degree angle with respect to the crown
points, as shown in Figure 7.3. From Table 7.1 the hot-spot stress range can be
calculated for points one to eight using Equation (7.1):
A/s = A/_ SCF
a
+ AfwB SCFIPB + AfoPB SCFOPB (7.1)
in which A/_, AfipB and AfoPB are the nominal brace stress ranges due to axial,
inplane and out of plane loadings, respectively. A better estimation of the hot
spot stress under combined loadings on a joint can be obtained by using the
formulae or methods which consider the interaction between the loads.
The nominal stress ranges are usually obtained from a computer analysis of the
platform structure under wave loading, wind loading, operational loads, etc.. To
assess the fatigue damage to the joint, the stress ranges corresponding to a
specific load case is calculated using a linear damage summation as shown in
Equation (7.2). It is usually assumed that fatigue failure occurs when under total
effects of individual loads the damage summation I,(n/N) becomes equal or
greater than one.
_ ; ^ A ( = I . O )
(7.2)
Chapter 7: Dynamic Stress Concentration Factor
130
Considering a wave distribution as shown in Table 7.2:
Table 7.2. No. of occurrences for different wave heights
used in the fatigue example
Wave height
(m)
1.5
4.5
7.5
10.9
17.1
Occurrence
(Direction from south)
1026467
252347
13621
1211
601
The following nominal stress ranges are obtained from a structural analysis,
when the third wave with the height of 7.5m is considered:
Table 7.3. Nominal stress ranges in brace member for
different deformation modes
Load type Nominal stress range
Axial 32.16
IPB 23.26
OPB 15
;
39
Using Equation (7.1) with the stress range values of from the above table and
SCFs from Table 7.1, hot-spot stress ranges for the eight points around the
intersection can be determined as shown in Table 7.4 below :
Table 7.4. Hot spot stress ranges for the T-joint in fatigue example
Location Hot-spot stress range (N/mm
2
)
1
2
3
4
5
6
7
8
Chord
393
435
385
435
395
435
385
435
Brace
295
333
298
333
294
333
298
334
It is seen in table 7.4 that the hot-spot locations are points 2, 4, 6 and 8 with the
m a x i m u m stress values of 435 N/mm
2
. Using a design S-N curve, N value is
assumed to be 29888, which is the number of cycles for fatigue life
Chapter 7: Dynamic Stress Concentration Factor
131
corresponding to the constant stress range of 435 N/mm
2
. The cumulative
damage then would be:
i _ ! _ _ _ o . 4 5 6 (7.3)
N 29888
V
'
The value 0.456 shows the degree of fatigue damage to the joint, produced by
the 7.5m high waves. The same calculations should be carried out for the other
wave heights and loadings, and the damage summation then compared with
unity. W h e n a factor of safety (FS) is assumed the maximum permissible
damage ratio will be reduced to 1/FS.
This simple example shows how crucial S C F is in the fatigue design of tubular joints. It is
the key parameter for converting a nominal stress, calculated in a structural analysis, to the
hot spot stress used for fatigue calculation. This study is trying to make improvements on
the calculation of both nominal stress values and stress concentration factors. The stiffness
consideration of the joints in the structural analysis increases the accuracy of the calculated
nominal stresses. It is shown in this chapter that the dynamic aspect of the loading on
tubular joints should also be considered in order to have better predictions of the SCFs.
The SCFs, in the above example, were calculated by using parametric formulae which is the
method commonly exercised in practice. However more comprehensive methods, which will
be explained in the next section, are required to establish the S C F parametric formulae.
7.3 Methods of SCF determination
The methods used to study the local behaviour of tubular joints are also applied to the
investigation of SCFs. Referring to Section 2.3, S C F may be evaluated via the following
methods:
1) analytical methods,
2) experimental methods, and
3) Finite Element method.
The major study on tubular joints involves stress distribution and stress concentration at the
joints. In this regard, experimental and Finite Element methods are most successful and
extensively used by the researchers. Using any method other than experiment is an attempt
to by-pass direct measurement and inevitably introduces some approximations. However,
when a non empirical model is confirmed by a systematic experimental approach, a formula
may be established from the results and used instead. The parametric formulae for S C F or
Chapter 7: Dynamic Stress Concentration Factor
132
stiffness of tubular joints are the examples. In this section a brief introduction is given to
strain gauging for strain and stress recovery in tubular joints. There will be also a general
review on some Finite Element studies.
7.3.1 Experimental methods
A m o n g the experimental methods of stress and strain identification, applying strain gauges
is the most popular and perhaps the oldest procedure. This method can be applied to steel
models as well as the models made of synthetic materials such as araldite. The most popular
strain gauges are foil gauges that were made in 1932 for the first time in the U S A . The foil
gauges are produced by a printed circuit process from selected alloys, which have been
rolled to a thin foil. The foil is placed on a backing material and encapsulated in an
insulating matrix. The surface area of the foil conductor per its cross-sectional area is very
large which provides good heat transfer between the grid and the specimen. The significant
economic measure in strain gauging is the total cost of the complete installation and the cost
of strain gauge itself is not ordinarily a prime consideration. There are of course some
applications in the industry that require very expensive strain gauges, too.
Strain measurements on tubular joints are generally carried out under static loads. The
specimens used for strain gauge tests require elaborate test rigs in order to be supported
properly, especially for ultimate strength tests in which large loads are applied, resulting in
expensive procedures. Steel tubular joints must be fabricated according to the standard
offshore procedures. Furthermore, models should simulate the real joints as closely as
possible in terms of geometry, material, fabrication procedures, loading and restraint
conditions. Other testing precautions are also necessary such as longitudinal seam welds on
the pipes which may affect the readings and should not coincide with the expected hot
spots, or weld size that could also be a problem for scaled down specimens.
The first strain measurements on tubular joints were conducted in the 1960s. Toprac and
Beale (1967) established the first parametric formulae using a limited database from steel
joints. Since that time a large number of experiments have been carried out on strain
measurements of tubular joints. U K O S R P (United Kingdom Offshore Steels Research
Program) has established several projects on the comparison of different methods of
studying the stress distribution of tubular joints. Based mainly on U K O S R P results, U E G
(1985) recommends experimental methods for the cases where,
1) high accuracy is required, or
2) the problem is such that the numerical methods are either not cost effective
or not readily interpreted and/or accepted (because of accuracy).
Strain gauges, like any other measurement device, have some particular deficiencies which
preclude their absolute accuracy. These are:
Chapter 7: Dynamic Stress Concentration Factor
133
1) gauge length that produces an average strain reading.
2) They are very difficult for determination of strain gradient, especially when
there is a high stress gradient, physical dimensions of strain gauges become a
problem.
3) Strain gauge implementation is impossible or sometimes very difficult on the
inside surface of tubes.
Apart from the above problems interpretation of strain readings for fatigue design purposes
requires a rational approach in order to produce consistent results from different tests. This
will be discussed in the following section.
7.3.2 Interpretation of stress concentration factor from measurements
There is debate on h o w strain measurements can be incorporated into S C F determination.
Several methods have been proposed, which generally try to avoid picking up notch
stresses, since their accurate measurement is not possible. This is due to the rapid three
dimensional variation of the notch stresses within a few millimetres of the weld toe and also
because the notch stresses depend on the weld geometry and vary from joint to joint
Therefore, the S-N curves established for fatigue life prediction of tubular joints are specific
to particular weld profiles. For the same reason a general S-N curve e.g. an as-welded S-N
curve becomes very conservative. The European Community of Steel and Coal (ECSC)
recommends a method for strain measurement to determine SCF. This method, which is
similar to that of the U K Department of Energy (DEn) Guidance Notes is shown in Figure
7.4.
There are two issues involved in the E C S C or similar methods. One is the location of strain
gauges relative to the weld toe and the other is the form of extrapolation on the strain
readings to obtain the hot spot strain or stress. Gurney (1979) recommends a distance of
approximately 0.35 times the wall thickness from the weld toe for applying strain gauges.
This has been obtained based on Finite Element analyses of simple fillet-welded joints in flat
plates. H e has not suggested any distance for the second strain gauge or guidance on how
to extrapolate to the toe position.
The following methods are generally applicable to the S C F evaluation of tubular joints,
when stress or strain extrapolation are required (Lalani et al., 1986):
1) linear extrapolation of maximum principal stresses,
2) curvi-linear extrapolation of maximum principal stresses,
3) curvi-linear extrapolation of maximum principal strains,
4) individual curvi-linear extrapolation of strains parallel, perpendicular and at
45 orientations to the weld toe and combining them to obtain the maximum
principal stress, and
5) curvi-linear extrapolation of strains perpendicular to the weld toe.
Chapter 7: Dynamic Stress Concentration Factor
134
0.6S(rt)
a = 0.2(d) , but not smaller than 4 m m
Chord
Figure 7.4. Hot-spot definition for T and DT-joints recommended by ECSC
In a paper in 1986, Lalani et al. describe the third method above as being increasingly
favoured by investigators. In that article, however, it is recognised that the linear
extrapolation of maximum principal stresses is appropriate for vertical braces, and the curvi-
linear extrapolation is for joints with inclined braces. It is emphasised by Lalani that the
resulting hot-spot stress should be compatible in terms of extrapolation technique with an
appropriate S-N curve.
7.3.3 The Finite Element Method ( F E M )
Numerical methods have performed extremely well in connection with structural analysis
since the advent of computers. The Finite Element method has become very popular as a
numerical tool to solve the deformation problem of continuum media, and most of the time
performs with sufficient accuracy. Greste (1970) analysed tubular K-joints using F E M . H e
used a mesh generating routine for definition of joint geometry which produced a well
designed mesh. Kuang et al. (1975) derived a set of S C F parametric formulae via Finite
Element analysis. Several other S C F studies have been conducted using F E M , e.g. Gibstein
(1978), Buitrago et al. (1984), Efthymiou et al. (1985) and Hellier et al. (1990), resulting to
the parametric S C F formulae. Some of the analyses included weld modelling. Similar to
strain measurement, the location of maximum stress is a controversial matter in the Finite
Element analysis of tubular joints. Most investigators have selected the brace mid-wall to
Chapter 7: Dynamic Stress Concentration Factor
135
calculate the hot-spot SCF, which differs from that which is practiced on test specimens. It
has been recommended (UEG, 1985) that a specific distance from the weld toe similar to
the electrical strain gauges be used for stress determination around the tubular intersections.
Apart from the problems regarding shell element in the Finite Element method discussed
briefly in Chapter 5, there are other deficiencies associated with the method. For example
residual stresses can be hardly considered in F E M . The weld is not usually included in the
Finite Element model of a joint, producing false bending moments for the areas near the
intersection. Mesh generating is another task that requires sufficient attention. Progressively
reducing the size of elements provides convergence to the exact solution, but becomes
costly due to the small mesh size. The actual mesh chosen then is a compromise between
acceptable accuracy and cost of analysis.
With all descriptions given above, one has to be cautious when modelling a tubular joint by
the Finite Element method to recover strain or stress values. In order to verify the F E
results, convergence studies or comparisons with test results are essential.
7.4 Reliability approach for stress concentration factor
In 1984, Tebbett and Lalani introduced a reliability approach to the stress concentration
factor concept used in design of tubular joints. The subject of their paper is the probability
that the stress concentration factor (SCF) used in calculating the hot spot stress range for a
defined input loading, will underestimate the actual hot spot stress range. The paper
compares the test results of steel tubular joints with the S C F prediction of parametric
formulae, considering variability of the S C F estimation for similar conditions.
Variations in evaluated SCFs stem from the following factors:
1) general form of parametric equation,
2) the method used for generating the original SCFs (e.g. F E M or model testing),
3) out of roundness tolerances on tubes,
4) wall thickness tolerances on tubes,
5) accuracy of fabrication,
6) weld size and profile,
7) accuracy of load application and measurement,
8) accuracy of gauge position and position measurement, and
9) accuracy of stress/strain measurement or conversion.
In the follow-up paper, Lalani et al. (1986) have continued the reliability approach they
introduced in 1984 and showed the variations in SCFs when different parametric formulae
are used. Figure 7.5a shows the distribution of the ratios of the SCFs obtained from
compression tests to the SFCs of the same joints from tension tests. It is seen that for 42
Chapter 7: Dynamic Stress Concentration Factor
136
specimens, approximately 37 S C F ratios (90%) are between 0.9 and 1.1. This indicates that
an S C F which is determined in a compression test can be 1 0 % different from when it is
calculated in a tension test. Further change in S C F is expected for the remaining 1 0 % of the
results. Also, nominally identical joints showed 1 0 % change in measured SCFs. This is
shown in Figure 7.5b where 29 SCFs out of 33 have a variation of 1 0 % with respect to the
average SCF.
Whether a test is performed dynamically or statically can influence the results as much as
those 1) to 9) above. Offshore rigs operate under dynamic loading that contributes to
fatigue in the critical areas such as joints. The different methods referred to, in Section 2.3
are generally performed statically and make no distinction when applied to a structure under
dynamic or static loads. Of course, dynamic loading will not entirely change the internal
forces of the structural members in comparison with static loading. It is important to
consider all parameters that influence the calculation of SCF, in order to establish a
reliability approach for SCFs. This will provide a higher accuracy in the structural analysis
Sample size = 33
JHL_
.9 1. 1.1 1.2 1.3 1.4
Measured SCF
Average SCF
ariations of SCF for nominally
s (after Lalani, 1986)
7.5 Dynamic strain reading
Electrical strain gauges can be used on the structures for dynamic strain measurement,
however more expensive and sophisticated instruments are required. A series of strain
measurements were carried out on eleven T-joints in this study for inplane and out of plane
bending modes. The geometry and non-dimensional parameters of the joints are reported in
Table 4.1. A typical diagram of the strain gauge locations is shown in Figure 7.6. Strain
gauges D and E are located at the proximity of the hot spots for inplane and out of plane
modes, respectively. Strain gauges C and B measure the nominal strain in the brace. The hot
and a higher safety in the structural design.
1.1 1.2 1.3 1.4
Compression SCF
Tension SCF
Figure 7.5a. Variations of SCF for direction of axial
loading (after Lalani, 1986)
Figure 7.5b. Vt
identical joint.
Chapter 7: Dynamic Stress Concentration Factor
137
spot strain gauges were located 5 m m , and the brace strain gauges were positioned at least
2d from the weld toes, where d is the brace diameter.
gauge length = 6 mm
Figure 7.6. A typical strain gauge location diagram of a test specimen
The purpose of the tests was to compare the SCFs of a joint under sinusoidal loading when
the load frequency was varying. It is realised that the real loading on the offshore structures
has a random distribution. However, the generation of inertia forces is the most important
difference between static and dynamic loads which is also simulated by sinusoidal loading.
Strain comparison serves the same purpose as S C F comparison because of the linear
relationship generally assumed between stress and strain in steel offshore structures.
The outcome of each test was a series of strain readings for two points on a joint, for
example D and Zc at points D and C in Figure 7.6. Hence, every two strain readings
correspond to a certain load frequency. The load was also measured to check the quality of
the input excitation to the joint specimen. In static tests the strain readings are single
numbers whereas for dynamic loads they are distributed quantities varying with time and
may be shown by a curve as in Figure 7.8. After the strain readings were recorded, the ratio
of the average readings for each frequency (
D(
_
v
/ec,av.) was calculated and used for
comparison with the ratios from other frequencies.
The strain gauges used had the following specifications :
Gauge length = 6 m m ,
Electrical resistance = 120 Ohms,
Gauge factor = 2.04.
Chapter 7: Dynamic Stress Concentration Factor 138
7.5.1 Test set-up
T w o series of tests were carried out to study the relationship between strain concentration
factor (SNCF) and load frequency for inplane and out of plane bending modes. A
diagrammatic test set-up for the dynamic strain readings of a tubular T-joint under inplane
bending is shown in Figure 7.7. A similar set-up was used for out of plane bending with the
load acting perpendicular to the specimen plane defined by the chord and brace. The load
was measured through the force transducer installed between the shaker and the joint. The
alternating analog signal (voltage) which was produced by the strain gauges under dynamic
deformation of the system, was converted to a digital signal. Diagrams of load and strain for
the same time base could be read directly on the monitor of the PC. Therefore, any noise or
interference from other sources would show on the screen and if necessary the test could be
repeated.
M
=n r
_____
v
-v
1500mm
1- Shaker
4- Strain gauge amplifier
8- Force transducer
2- Shaker amplifier
5- AD converter
9-Strain gauge
3-Charge amplifier
6 and 7- PC
Figure 7.7. Diagrammatic test set-up for the dynamic strain measurement
Chapter 7: Dynamic Stress Concentration Factor 739
Only sinusoidal loads were applied to the joints, with the frequencies from 0.5Hz to 55Hz
and 5Hz increments. N o provision was made for temperature compensation of strain gauges
since the rapid change of strain did not allow any temperature effects on the leadwires.
Typical strain readings for IPB mode, taken at the crown point and on the brace is shown in
Figure 7.8.
20m/Div inplane T10/3
200
Micro
Strain
50
/Div
-200
0 - seconds - 200m
Figure 7.8. Dynamic strain readings for joint TIP
In the following section the results of inplane and out of plane strain measurements will be
presented and discussed.
7.5.2 Test results
Inplane bending mode
Table 7.2 presents the results of the measurements carried out for the inplane bending case.
There are two strain measurements reported for each frequency taken at points C and D
(see figure 7.8), on the brace and hot spot, respectively. The ratio of the two strains is also
reported.
Strain ratio and load frequency relationships are also shown in Figures 7.9a to 7.9c, where
an increasing variation with respect to load frequency is generally observed, however in
some cases the variation is not very pronounced.
Chapter 7: Dynamic Stress Concentration Factor
140
Table 7.2. Dynamic strain readings and SNCFs for IPB mode (mm/mmxlP
6
)
Jo
int
1
2
3
4
5
6
7
8
9
10
11
strain
location
EC
ep/ec
EC
E
D
ED/EC
Ec
E
D
ED/EC
Ec
ED
ED/EC
EC
E
D
ED/EC
EC
E
D
ED/EC
EC
ED
ED/EC
Ec
E
D
ED/EC
EC
E
D
ED/EC
EC
E
D
ED/EC
EC
E
D
ED/EC
Frequency (Hz)
0.5 5 10 15 20 25 30 35 40 45 50 55
3.035 4.894 4.95 4.538 4227 4.442 4.608 4.794 4.92 5.12 5.471 5.945
3.586 5.445 5.512 5.08 4.752 5.045 5.338 5.625 5.895 6.284 6.893 7.718
1.182 1.113 1.114 1.119 1.124 1.136 1.158 1.173 1.198 1.227 1260 1298
6.632 11.86 10.49 10.43 10.47 12.48 14.03 17.17 23.85 44.31 63.87 23.79
12.33 21.07 18.68 18.73 18.99 22.99 26.04 33.01 46.88 89.59 133.5 51.56
1.859 1.777 1.781 1.796 1.814 1.842 1.856 1.923 1.966 2.022 2.09 2.167
7.595 13.16 12.74 12.51 12.51 14.9 16.09 19.31 25.61 41.64 74.42 37.1
1124 18.62 18.05 17.85 18 21.75 23.88 292 39.66 66.16 122 62.8
1.48 1.415 1.417 1.427 1.439 1.46 1.484 1.512 1.549 1.589 1.639 1.693
39.03 6728 69.48 72.61 70.51 113.3 1242 8321 50.91 31.96 22.34 16.54
24.94 41.89 43.65 46.44 462 76.59 87.42 61.66 39.86 26.68 20.07 16.18
.639 .623 .628 .640 .655 .676 .704 .741 .783 .835 .898 .978
1.697 2.851 2.601 2.448 2.416 2.62 2.645 2.801 2.961 3253 3.675 4.341
6.085 9.748 8.894 8.387 8249 8.957 9.043 9.486 1029 1128 12.78 15.09
3.586 3.419 3.419 3.426 3.414 3.419 3.419 3.387 3.475 3.468 3.478 3.476
8.869 13.94 12.87 12..9 13.64 15.9 19.67 27.42 49.09 33.67 24.67 13.57
17.17 28.06 25.95 26.32 2822 3321 4226 60.51 112.7 80.01 61.13 34.64
1.936 2.013 2.016 2.04 2.068 2.089 2.148 2207 2296 2.376 2.478 2.553
8279 14.07 13.63 1424 14.19 16.95 18.17 23.14 3725 68.12 35.86 18.15
16.58 26.5 25.74 27.14 27.44 32.92 36.5 47.88 79.16 149.8 81.83 43.34
2.003 1.883 1.888 1.906 1.934 1.942 2.009 2.069 2.125 2.199 2282 2.388
37.11 64.39 65.51 68.72 8925 1032 87.33 56.74 31.91 24.45 16.8 12
46.7 78.34 80.6 86.31 115.4 138.8 122.7 84.85 51.08 42.38 31.92 25,45
1258 1217 123 1256 1293 1.345 1.405 1.495 1.601 1.733 1.9 2.121
2.453 4257 4.008 3.896 3.534 4.02 4.143 4.347 4.618 5.116 5.91 7.603
5.69 9.563 9.052 8.88 8.091 9248 9.785 10.54 11.58 13.32 16.07 21.85
2.32 2246 2258 2.279 2289 2.3 2.362 2.425 2.508 2.604 2.719 2.874
8.379 14.35 14.13 14.35 15.39 18.47 2425 36.64 54.51 28.58 14.72 8.949
29.76 47.89 47.41 48.77 53.4 65.34 88.71 139.1 216.6 119.5 65.31 42.62
3.552 3.337 3.355 3.399 3.47 3.538 3.658 3.796 3.974 4.181 4.437 4.763
43.3 66.75 68.37 71.31 92.44 84.83 61.95 4023 26.1 17.12 11.95 8.431
77.18 112.8 1172 125.5 168.9 162.9 126.5 88.42 62.74 45.86 36.45 30.19
1.782 1.69 1.714 1.76 1.827 1.92 2.042 2.198 2.404 2.679 3.05 3.581
E
D
is measured at hot spot (crown).
_c is measured on brace.
141
%/ 3 'OI'JBJ UTBJ'JS
C4
I
f
" o ^.
:*(g
3
0

e
n
c
y
,

a
i S"
_.
:o t_
1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 r 1 1 1 1 1 1 1 1 1 1 1 1
O
% / 3 'Ol^.I UIBJ^S
rrf-O
o
s
a
B
1
S
c
<_
_0
a
00
<N
II
c__
"->
C5
II
ao.
3"
c?
"-:
II
CQ.
"a
I
2
5 c
K
!
S
a
.
s.
_.
_
s
?_<
"5
<s
4
I-
1

II
>-
u>
1-
\
\
= M \
* = \
So \
V 8 o \
.. -B \
CVJ _ \
<s
oo
8
II
>-
of
H
I i i i r i i i i i i i i i i i n
:
f
1
1
NO
tn
ro
* *
II
>-
-_r
r-
q
-
ctf
a.
\
':
7
:
:
i-
z
i
-
r
:
i i I i i i i i i i r i I i i i i i i i i i -
S
o g
J
8"
% / ^ 'OI'JBJ UIBJ^S
a

= !
is *
? ^
U (.J
S *
<_ 5
I *
OS _j
* i
a .2
_o *-
s:
Chapter 7: Dynamic Stress Concentration Factor
142
The diagrams in Figure 7.9 have a clear indication that the strain ratios calculated from the
readings at the foot of the welds and brace members have a variation with respect to the
load frequency. Apart from joint T5 in Figure 7.9a, it is seen that the strain ratios of the
joints with large y ratios have the biggest and the steepest variations with respect to the load
frequency. The maximum changes of the strain ratios calculated in a frequency span of
0.5Hz to 55Hz are 2 3 % , 3 4 % , and 1 0 0 % for T9, TIO, and Til respectively. It is also noted
that the variations of strain ratios with respect to load frequency increase for small P ratios.
This means that there is a more possibility of S C F variation due to dynamic loading in the
joints with thin chord walls or small brace diameters. According to the findings in Chapter
5, these types of joints possess a low stiffness.
The real loading on offshore structures has a frequency of less than 1Hz which means a
period of greater than 1 second. Although the frequencies of the loadings used in this study
span well beyond 1Hz but the results show that the strain ratio or S C F is frequency
dependent. The range of frequency used for loadings in the tests herein would only magnify
the observed phenomenon.
Out of plane bending mode
Out of plane bending measurements produced similar results to the inplane bending case.
Measurements of the hot spot (saddle point) strains and the nominal brace strains are
reported in Table 7.3. Points E and B correspond to the hot spot and brace, respectively.
Figures 7.10a to 7.10c shows the relationships of strain ratio with respect to load frequency.
Although the strains in O P B tests were measured up to 55Hz, but due to the very large
ratios obtained for T9, TIO, and Tl 1 after 40 Hz, only the ratios related to 0.5Hz to 40Hz
were plotted in Figure 7.10.
It is seen in Figure 7.10 that the change of strain ratios with respect to load frequency in
O P B mode is bigger for the joints with large y ratios compared to other joints. For this
range of y, the strain ratio shows also the steepest variation especially at high frequencies.
Each diagram in Figure 7.10 is for a different P ratio. The maximum variations of the strain
ratios with respect to 0.5Hz frequency are 3%, 7 5 % , and 1 2 0 % for T9, T10, and Til,
respectively. It is seen that, similar to IPB mode, the joints with low p values show the
greatest variations in strain ratio. The findings here indicate that out of plane bending SCF
has the largest variations with respect to load frequency in the joints with small P or large y
ratios.
Comparing the results of IPB and O P B modes in Figures 7.9 and 7.10, it is seen that higher
strain ratios have been resulted for O P B mode. The increase in strain ratios is noted to
occur drastically for high load frequencies in the O P B tests.
Chapter 7: Dynamic Stress Concentration Factor
143
Table 7.3. Dynamic strain readings and SNCFs for OPB mode
Jo
int
1
2
3
4
5
6
7
8
9
10
11
strain
location
E
B
EE
E
E
/E
B
EB
E
E
E
E
/EB
EB
EE
Efi/Efl
EB
E
E
E
E
/E
B
E
B
E
E
_/e
B
EB
E
E
E
E
/EB
E
B
E
E
E
E
/ E
B
E
B
E
E
E
E
/E
B
EB
E
E
E
E
/E
B
E
B
E
E
_
E
/e_
E
B
E
E
E
E
/EB
Frequency
0.5
1.777
3.889
2.19
8.93
46.26
5.18
8.475
38.49
4.542
91.39
191.3
2.093
1.621
11.72
7.230
8.37
65.97
7.882
9.502
42.36
4.458
43.72
130.3
2.980
1.532
26.11
17.04
8.96
82.99
9.262
41.91
244.8
5.841
5
3.069
6.701
2.18
14.82
72.5
4.892
15.23
64.55
4.238
69.48
136.4
1.963
2.94
20.27
6.895
14.45
105.7
7.315
16.19
67.24
4.153
68.93
190.1
2.758
2.479
40.19
16.21
15.53
129.7
8.352
66.44
355.1
5.345
10
2.952
6.431
2.18
15.19
74.64
4.914
15.11
64.65
4.279
71.89
143.3
1.993
2.735
18.85
6.892
16.29
121.7
7.471
16.51
69.64
4.218
69.95
197
2.816
2.643
42.54
16.10
17.63
150.7
8.548
63.09
352
5.579
15
2.956
6.415
2.17
17
84.4
4.965
16.16
70.38
4.355
76.72
157.3
2.050
3.026
20.89
6.904
19.83
153.8
7.756
18.14
78.95
4.352
74.23
217.7
2.933
2.922
47.2
16.15
23.02
209.1
9.083
57.53
344.3
5.985
20
3.074
6.65
2.16
22.16
111.8
5.045
18.88
84.59
4.480
86.04
183.6
2.134
3.463
23.96
6.919
30.21
246.8
8.169
25.7
116.9
4.549
76.16
236.6
3.107
3.607
58.89
16.33
29.26
287.1
9.812
39.69
267
6.727
25
3.373
7.274
2.16
35.54
183.4
5.16
25.87
120.4
4.654
85.42
192.2
2.250
4.19
29.05
6.933
36.01
316
8.775
41.94
203.2
4.845
57.69
194.3
3.368
4.916
79.72
16.22
13.68
150.3
10.99
23.89
188.7
7.899
30
3.967
8.498
2.14
47.17
249.9
5.298
45.03
220.4
4.895
59.94
144.3
2.407
6.558
45.62
6.956
13.8
132.8
9.623
20.9
108.8
5.206
33.01
131.3
3.978
10.45
170
16.27
6.006
77.46
12.90
13.69
127.9
9.343
35
5.197
11.05
2.13
21.74
118
5.428
29.16
151
5.178
36.37
94.24
2.591
11.66
81.94
7.027
6.428
70.05
10.90
9.577
54.81
5.723
19.97
83.66
4.189
6.058
101.9
16.82
2.946
47.66
16.18
6.758
87.79
12.99
40
8.577
18.13
2.11
10.71
60.04
5.606
13.03
71.89
5.517
22.84
65.42
2.864
6.536
45.85
7.015
3.423
43.91
12.83
5.136
33.24
6.474
12.26
60.7
4.951
3.059
50.4
16.48
1.372
35.19
25.65
2.997
66.44
22.17
45
18.54
38.69
2.09
6.751
39.46
5.845
7.419
44.32
5.974
14.76
47.68
3.230
3.318
23.28
7.016
1.997
30.98
15.51
3.014
23.25
7.714
7.408
45.65
6.162
1.788
28.96
16.20
.6697
22.51
33.61
.5956
52.72
88.52
50
9.174
18.91
2.06
4.297
28.06
5.630
4.756
31.17
6.541
10.4
38.84
3.735
2.097
14.71
7.015
1.081
23.76
21.98
1.812
16.48
9.095
4.441
36.92
8.313
1.245
19.39
15.57
.2274
17.66
77.66
1.169
43.55
37.25
55
4.725
9.628
2.04
3.407
21.75
6.384
3.201
23.32
7.285
7.197
32.2
4.472
1.419
9.931
6.999
.495
18.08
36.52
1.048
13.06
12.46
2.366
31.06
13.13
.9432
14.11
14.96
.3658
13.98
38.22
2.37
36.57
15.43
E
E
is measured
EB is measured
at hot spot (saddle).
on brace.
144
*h/ ^3 'OT\VJ IIIBJ^S
3/ ^3 'oryBJ UIBJ^S
53
3
C
f
s
oo
c_
II
CO.
c_>
II
sn.
S"
c>
i
II
__.
"a
1
I
<_
I
CO.
a
s
>_
_5
<_
s
"8
OO
M
S
II
>-
P
O
II
ca
OQ
Q.
O
/
II
/
o
r ">
o
o
CM
i | l i l l l l l l
O
CO
V
a
S"
fa
.
3/ ^3 OI-JJ: urexiS
a
.5-
S
su
3
C_<
c _
2 S.
1 g
<=> *
L_ **
a .a
s:
Chapter 7: Dynamic Stress Concentration Factor 145
7.6 Dynamic strain calculation using stiffness method
In this section the performance of the stiffness method in predicting the strain ratios at
tubular joints under dynamic loading is examined and the results are compared with
experiment In the model shown in Figure 7.11 a continuous mass system is assumed for
calculation of the inertia forces. W e call this model a beam model as distinct from a test
model (experiment). Since the stress or strain concentration can not be considered by the
stiffness method, the strains at half way from the intersection to the chord support are
calculated (point D in Figure 7.11) instead of a location close to the weld. Therefore the
experimental data which are used in conjunction with the results of the stiffness method is
different from those previously showed in this chapter. The difference is the location of the
strain gauges on the chords. Only inplane bending is considered in this section since out of
plane bending produced very small strains at point D and therefore a reliable strain
measurement was not possible. The loading was a sinusoidal force at the top of the
cantilever equal to 1000N. The strains were calculated at points D and C. The strain gauges
were applied to the test models at the same locations as on the beam models (points D and
C in Figure 7.11) to obtain comparable results.
The strain ratios obtained for the beam models and the test models are shown in Figure 7.12
for a number of joints. The joints with a diameter ratio of 1 are compared in Figure 7.12a,
which are three joints with different y ratios. It is seen that the results of the test and the
stiffness method are approximately 1 0 % different for Tl which can be counted as a good
agreement. However, for T 5 and T 9 the difference between test and theory becomes much
bigger than Tl. Furthermore, the variation of the strain ratios for the test models show a
bigger slope than the beam models, especially in high frequencies.
The same trend is observed for the test and analytical results in Figures 6.7b and 6.7c as in
Figure 6.7a. The stiffness method and experiment on the joints with low diameter to
thickness ratios (T2 and T4) show a good agreement. It is again seen that for the joints with
large y ratios, there is a bigger difference between the results of the test and stiffness
method. Although, the beam model has not predicted the strain ratios of tubular joints as
accurate as experiment, however it shows a dependence of strain ratios on the load
frequency, as observed on the calculated values.
The comparison of stiffness method and test results shows that:
1) Stiff T-joints are less prone to the variation of strain concentration factor due to
load frequency.
2) Stiffness method does no produce reliable results for the variation of strain
concentration factor under dynamic loading.
Chapter 7: Dynamic Stress Concentration Factor 146
P = P
e
Sln(ot
1230
h-375 -H
z.
c-
D
All dim. in m m
T
h
___.
1500
h=3S0 tor T2,14,16,18,110, and 111
h=420forl1,l5,andl9
T9,T= 22-82
T5,7=17.12
T1,T= 13.36
m i TI 1111 n 11 n r TI p i ii n n 11 I I H i n i tji 111111
20 30 40
Frequency (Hz)
60
Figure 7.11. Beam model to study dynamic
effects on strain ratio
a) Joints with $ = 1.0
0.35
0.05"
* * * * * Test Model
o o o o o Beam Model
T10
T6,T= 1712
T2,T= 13-36
0.0 111111111111 11111111111111111111111111111
0 10 20 30 40
Frequency (Hz)
50
60
0.096-,
0.080
J"
0.064-
h
0.048
d
M 0.032
0.016
0.0
__*-_ Test Model
o o o o o Beam Model
IPB, P=0.28
T11.T = 22.82
T8,T= 1712
T4.7= 13.36
T11
11 m i n i | m i |iin |IIIIIIIII| iiii|illllinl|
0 10 20 30 40 50 60
Frequency (Hz)
b) Joints with p = 0.52
c) Joints with p =0.28
Figure 7.12. Strain ratios resulted from beam models and test models for joints with a)$ = LP, b)$ = P.52,
and $ = 0.28
Chapter 7: Dynamic Stress Concentration Factor
147
7.7 Dynamic strain reading and S C F / S N C F reliability
The variation of strain concentration factor under dynamic loading is an indication of
different response behaviour of a structure to dynamic loads compared to static loads.
Stress or strain concentration factors are currently calculated based on static tests or static
Finite Element analyses. Therefore, the variation of stress concentration factor with respect
to load frequency is not considered in the analysis procedures. However, offshore structures
operate under dynamic loadings and therefore their behaviour should be studied and
predicted under such conditions. Dynamic loading on a structure produces inertia forces
acting all along the members, which is described well by the dynamic deformation
formulation used in this study. These distributed forces vary with load frequency and create
a variation of the nominal forces or stresses in the structural members. Therefore, S C F or
S N C F become variable with respect to load frequency.
Tebbett and Lalani (1984) suggest four primary parameters to be considered in the
calculation of the probability of failure by fatigue. These variables are:
1) loading spectrum,
2) accuracy of SCF,
3) scatter in fatigue life for known 'hot spot' stress range, and
4) relationship of stresses at the point of consideration to calculated SCF.
Dynamic strain reading may be placed in the second category of the variables suggested by
Tebbett. It can improve the accuracy of the measurements on physically tested models, and
therefore reflects one of the reasons for the reliability approach to the calculation of SCF or
SNCF.
The best solution to take the load frequency into account is to employ a frequency
dependent factor in the parametric S C F formulae. The determination of such a factor
requires further analytical and experimental studies.
Another solution to consider the variation of S C F with respect to load frequency is also
possible by using a reliability approach. Since the environmental loads impose typically low
frequencies on the offshore structures, it seems to be possible to introduce a constant factor
to give the appropriate level of confidence of SCFs. This approach comes with the concept
of design S C F to be used in the design equations. Design S C F is given by Equation (7.4)
below (Tebbett and Lalani, 1984):
SCF
D S C F ________ (7.4)
r
in which DSCF is the design stress concentration factor and T is the partial factor selected
to give the appropriate level of confidence. Depending on the approach taken for an
Chapter 7: Dynamic Stress Concentration Factor
148
individual design, T could range between 0.6 and 1.5. The T values lower than one will give
increased confidence in design, Le. less chance of a higher actual SCF.
7.8 Summary
It has been shown in this chapter that the static methods of strain/stress determination do
not produce necessarily conservative strain/stress concentration factors when a structure
experiences dynamic loads. The variation of S C F due to the load frequency becomes
important in fatigue life estimation of tubular joints in offshore structures. The parametric
equations, which are used to determine the hot spot stress at a tubular joint, are all based on
the databases from static experiments. Therefore, they do not recognise the dynamic loads
which are applicable to the offshore platforms. As the experiment and analytical solutions
showed in this study, whether the loads applied to a joint are static or dynamic makes a
substantial difference to the calculated values of stress or strain concentration factors.
A reliability approach was introduced by Tebbett and Lalani (1984) to the S C F parametric
equations. The reason for introduction of such approach was the inconsistent results
obtained from different formulae or tests on the joints with the same specifications.
Dynamic load effects on S C F estimation can be considered the best by employing a
frequency dependent factor in those equations. Another solution can be employed based on
a reliability approach in the form of a constant factor, because the dominant loads on
offshore structures have low frequencies. However, in both solutions there is a need to
evaluate those factors realistically using analytical or numerical dynamic methods of analysis
supplemented by experimental investigations.
Chapter 8
SEMI-RIGIDITY EFFECTS ON BEHAVIOUR
OF OFFSHORE TOWERS
Chapter 8: Semi-Rigidity Effects on Behaviour of Offshore lowers
150
8.1 Introduction
In order to study the effects of joint rigidity on the behaviour of offshore structures, a tower
is analysed in this chapter for two states of rigid and flexible joints. To be able to compare
the results with another reference, the geometry of the tower is taken similar to the one
used in the 'Node flexibility and its effect on jacket structures,' report U R 2 2 by the
Underwater Engineering Group (UEG, 1984).
T w o aspects of joint stiffness are studied in this chapter. The first aspect is the effect of
joint stiffness on the results of structural analysis including bending moments, axial forces
and natural frequencies. This is generally similar to what was carried out in the report
U R 2 2 and is required to apply the results in fatigue analysis. However, the method of
flexibility implementation in the analyses herein, which uses a continuous mass model for
beam members, is different from U R 2 2 study and is described in Chapter 3. The second
aspect of study in this chapter is the effect of joint stiffness on the fatigue life of tubular
joints. The fatigue life is calculated for a wave condition taken from the book 'Dynamics of
fixed marine structures,' (Barltrop, 1991). To the knowledge of the author this part of the
study is not carried out before. The purpose of the analyses herein is to investigate the
effects of joint rigidity on the structural behaviour of a tower frame. This will indicate how
much variation a designer will achieve in the results of his analysis assuming flexible
joints. In this study a frame with rigid joints is called a conventional model.
All computer programs used in this chapter are written by the author. A continuous mass
formulation is used in the analysis of the towers. Therefore, there was no need to assume
extra nodal points on beam elements between the joints in dynamic analyses.
8.2 Tower geometry and joint specifications
U R 2 2 study, from which the geometry of the towers in this chapter is taken, has considered
three frames with different configurations. The frames in U R 2 2 have been intended to be
representative of the platforms in the North Sea. In order to show the semi-rigidity effects
on the structural behaviour, the most flexible tower configuration in U R 2 2 study was
selected and analysed here. This tower and the geometry of its members are shown in
Figure 8.1.
Chapter 8: Semi-Rigidity Effects on Behaviour of Offshore lowers
151
In order to be consistent with the U R 2 2 study the tower with rigid joints is called CP1 and
the one with flexible joints M P 2 (see Section 2.4 for the tower identification names). The
non-dimensional parameters of the joints in CP1 and M P 2 frames are reported in Table 8.1.
The chords of the joints in the tower are assumed to be stiffened with joint cans at the
intersections. Also brace stubs are considered for the horizontal braces. To strengthen a
joint, the chord section is usually thickened in the connection zone. This section with
higher strength is called joint can. Similar zone on brace is called brace stub.
10000
17.0x33.5 '
18045
22054
26955
tane =0.1
y = 45
32946
36240
dim. in m m
Figure 8.1. Geometry of the tower analysed
Table 8.1. Geometrical properties of joints used in the analysed towers (units in millimeters)
frame with:
rigid joints
semi-rigid joints
name
CP1
MP2
d45
900
900
braces
t45 d90
25
25
750

t90
25

D
1700
joint stiffness based on
T d45 d90 D/2T
33.5 900 25.4
d/D
0.53
Stiffness of the joints was accommodated in the analyses by using the modified stiffness
matrix for the members, described in Chapter 3. Therefore, both frames with rigid and
semi-rigid joints were modelled using the centerlines of the members. Joint cans were not
modelled separately but the stiffness of the joints were calculated using the chord can
geometry.
Chapter 8: Semi-Rigidity Effects on Behaviour of Offshore Towers
152
Because the focus of this study is on single braced joints, the inclined braces have been
assumed to have hinged connections to the chords; i.e. they do not produce any eccentricity
or bending stiffness at the joints. Therefore, only T-joints were considered to be semi-rigid,
and the inclined braces were assumed as truss members.
8.3 Stiffness determination of the joints of the tower
The stiffness of the T-joints in M P 2 was determined by using F E Method in conjunction
with the theoretical model described in Chapter 3. A Finite Element analysis of a T-joint
using shell elements with the following details was carried out for its natural frequencies.
_>1700mm, 7=33.5 mm, 1=13.5 m
d=900 m m , r=25 m m , 1=5 m
The calculated results for the natural frequencies and stiffness of the T-joints in M P 2 tower
and also the values used by TJR22 Report for the same joints are shown in Table 8.2. In this
table, a model constructed from one dimensional beam elements is called the line model. It
is seen that the calculated stiffness of the T-joints using the method of this study are close to
those used in U R 2 2 study especially IPB results. Therefore, to maintain the consistency of
the input data with U R 2 2 study, the stiffness values of U R 2 2 were also used in the analyses
of this chapter.
Table 8.2. Natural frequency and stiffness of the T-joint in MP2 tower
Mode Joint type Natural frequency (Hz) Joint stiffness
Line model FE model This study UR22 study
IPB T 31.68 15.65 292 M N m 379 M N m
Axial T 21.15 21.51 375 M N / m 559 M N / m
8.5 Loading
Four load cases were selected which are the same as U R 2 2 study. A concentrated load of
10 M N was applied at the top of the tower to study the transverse deflection changes. To
investigate the effects of joint flexibility on the internal forces, three wave loads derived
from a representative 100 year storm wave passing through the towers with three phase
angles were considered (see Table 8.3 and Figure 8.2). Loading produced by sea currents
was not considered in the analyses.
The square root of the sum of the squares of the instantaneous responses to two waves 90
phase angle apart was calculated to obtain the response amplitude of the bending moments
(Barltrop, 1991). This simple method is used to make an estimate to the response of a
structure under sinusoidally varying waves with various phase angles. Therefore, the
Chapter 8: Semi-Rigidity Effects on Behaviour of Offshore Towers
153
amplitude of bending moments was calculated using the results of load cases 2 and 3 from
Table 8.3. The response amplitude, x_ is generally defined as:
* _ = i / ( * _ + * 3
2
) C
8
-
1
)
where x
a
= response amplitude,
x-i = response from load case 2,
*3 = response from load case 3.
Response amplitudes of bending moments are compared to determine the overall effect of
allowing for joint flexibility. To compare the distribution diagrams of a response such as
axial force or bending moment, a further wave load with phase angle 45 (case 4) was
chosen. It applies both horizontal sway forces and local loading on individual vertical,
inclined and horizontal members and therefore includes all the important effects of a wave
loading (UEG, 1984).
Table 8.3. Details of the load cases in the analysis of towers
Load case
number
Load type and direction Numerical
details
1
Concentrated load, P in x direction applied to
Node 10
Linearly varying distributed loads acting
perpendicular to all members derived by using
linear (Airy) Wave Theory and Morison's
Equation for member loading.
P=10MN
<|> = 0
o
4) = -90
4> = -45
Mean water level, d = 85m Wave period, T = 16sec
Wave height, H = 30m Wave phase angle,*
M W L
Direction of w a v e propagation
<|> = 0
= -90
$ = ^ 5
Figure 8.2. Wave definition for load cases 2,3 and 4 (UEG, 1984)
The wavelength (_.) was calculated to be 360.5m by using Equation 8.2 (Barltrop, 1991,
Equation 6.97):
2 UJ\
27t
In
\^J
(8.2)
Chapter 8: Semi-Rigidity Effects on Behaviour of Offshore Towers 154
in which g is the acceleration due to gravity and T is the wave period. Wave loading was
calculated using a Fortran program written by this author where linear (Airy) Wave Theory
and Morison's equations, given in Table 8.4 below (Barltrop, 1991), were implemented.
Table 8.4t Linear (Airy) Wave Theory equations (Barltrop, 1991)
Surface elevation = y cos 27i( j - j) J = 9f cos A
H
2
Horizontal particle velocity: u
Vertical particle velocity: w
Horizontal particle acceleration: du/dt
Vertical particle acceleration: dw/dt
Force/unit length of a member: F
(perpendicular to member axis)
_ nH cosh(k(z + d))
14 ;__ ~ COS A
w
Tsinh(kd)
nHsinh(k(z + d))
sin A
Tsmh(kd)
du 2n
2
Hcosh(k(z + d))
dt T
2
sinh(JW)
dw -2n
2
Hsmh(k(z + d))
sin A
COS A
in which
dt T
2
sinh(kd)
F = 0.5C
d
pD\U\U+C
m
pAU
H = wave height,
d = water depth,
z = distance above mean water level,
L = wavelength,
and k = 2-K/L.
T See Figure 8.2 for a better explanation of the notations used in this table.
Drag and inertia coefficients (C
d
and C^) for cylindrical sections were assumed to be 0.7
and 1.8, respectively (Barltrop, 1991). The analyses were carried out under static loads
where dynamic magnification was not considered.
The wave forces calculated for load cases 2, 3 and 4 are shown in Figure 8.3. The
specifications of the waves are reported in Table 8.3 and Figure 8.2. The forces were
calculated for each member between two nodes and assigned to those nodes. As it is seen
in Figure 8.2, the wave crest in load case 2 is at the top of the tower. However, in load
cases 3 and 4 the water surface is lower than the frame top level.
It is seen in Figure 8.3 that the forces in load cases 2, 4, and 3 corresponding to phase
angles 0, -45, and -90 are decreasing in the order of magnitudes. However, the response
of a structure to wave loading will be a combination of the responses to the waves with
various phase angels, and should be calculated using an appropriate method.
Chapter 8: Semi-Rigidity Effects on Behaviour of Offshon? Towers
155
a) Load case 2, = 0 b) Load case 3, = -90 c) Load case 4,= -45
Figure 8.3. Wave forces calculated for load cases 2, 3, and 4
As it is seen in Figure 8.3, the maximum force intensity occurs at the top of the tower in
each load case and the wave force decreases towards the seabed. The forces are calculated
perpendicular to the members. However, at a node where a horizontal brace meets the leg
the sum of the forces will not be perpendicular to the leg member, due to the force
component from the brace.
8.6 Results
A detailed report is given in this chapter from the analyses carried out on the frames M P 2
and CP1. Where numerical values are reported, the locations selected for the output forces
or deflections have been chosen to correspond with those in U R 2 2 report.
8.6.1 Deflections
Load case 1 was selected in U R 2 2 for studying the deflection changes. However the
deflections herein are calculated for all four load cases with the maximum values as shown
in Table 8.5. It is seen in Table 8.5 that the difference between the maximum deflections of
CP1 and M P 2 is about 10%, comparing 3 4 0 m m and 373mm, for the location under the
point load of 10MN.
Chapter 8: Semi-Rigidity Effects on Behaviour of Offshore Towers
156
The maximum deflections under wave loads indicate a negligible change when the joint
stiffness is employed in the analyses. These are the cases where the loads are distributed
over the members. The reason for the small differences of displacements is because the
lateral stiffness of the tower depends on its overall stiffness as a cantilever, which in turn
depends on the cross sectional area of the legs and their distance from each other.
Table 8.5. Deflection changes when joint flexibility is incorporated in analysis
Frame
CPl
MP2
Load Case 1
0.3401
0.3732
%
9.7
max deflection (meter)
Load Case 2 %*
0.0679
0.0689 1.5
Load Case 2 %*
0.0259
0.0263 1.5
Load Case 3
0.0494
0.0500
%*
1.2
* %: Deflection change of M P 2 with respect to CPl.
The deformed shapes of CPl and MP2 under the point load of 10MN in load case 1 are
shown in Figure 8.4. It is noted that almost all the deflection change in the frame with semi-
rigid joints is due to the axial flexibility of joints 9 and 10.
i-WW^
k=559E9MN/m
k=623E9MN/m
A5=10 (~+ ^ - ) = 0.034 m = 34 mm
10MN
rigjd=dashed line (CP1)
semi-ngid=sloid line (MP2)
0.5m
I 1 1 h
Figure 8.4. Deflected shape of CPl and MP2 frames for load case l(point load at the top corner)
To verify whether the deflection difference between CPl and M P 2 is related to the
flexibility of joints 9 and 10, beam element 13 which connects the two joints is considered.
The axial force in this element is 1 0 M N . It is seen in Figure 8.4 that the total axial
deflections due to the springs at the ends of beam 13 is 34mm. From Table 8.5 the
difference in the displacements of CPl and M P 2 is: 373 - 340 = 3 3 m m which is very close
to 34mm. This confirms that the flexibility of the joints under point loads introduce the most
changes of deflections between the two models. This is important when the effects of
impact loads are considered on a platform. The U E G study, U R 2 2 , reported a change of
1 6 m m or 5 % for M P 2 in load case 1, which is almost half the value calculated here. Tall
platforms, though, may show more variation in lateral deflections when joint flexibility is
considered as reported by B o u w k a m p (1981).
Chapter 8: Semi-Rigidity Effects on Behaviour of Offshore Towers
157
8.6.2 Axial forces
The diagrams of axial forces for load cases 2, 3 and 4 and the changes of the forces between
the conventional and flexible joint analyses are shown in Figure 8.5. There was generally a
very small difference between axial forces resulting from the analyses on CPl and M P 2 . The
small variation of axial forces may be assigned to the unique load path exist in the towers.
Further analyses are required on the towers with redundant members and different
geometrical configurations to study the effect of joint stiffness on axial force distribution.
0 1 0 M N
I I I I I I
load case 2 (<|> = 0) load case 3 (| =-90) load case 4 (<)> = -45 )
a) Axial force diagrams for different load cases
0 0.1 MN
l l l l l l
$ = 0 ^ = -90 <|> = -45
b) Axial force changes
Figure 8.5. Axial forces for load cases 2, 3 and 4 and their variations between flexible joint analysis and
conventional analysis (MP2-CP1)
Chapter 8: Semi-Rigidity Effects on Behaviour of Offshore Towers
158
The axial forces for a number of elements corresponding to load case 4 are reported in
Table 8.6. It is seen that the maximum change is due to element 11 and is 0.18 M P a or
4.3%. The maximum change in the response amplitude of axial force occurs in member 10
and is 4.9%.
Table 8.6*. Axial forces, their changes and axial stress changes for load case 4.
Chord
45 Brace
Member No. 2
15
Analysis
t Element numbers referred to here, are shown in Figure 8.1.
16
Horizontal
Brace
11
Axial force
(MN)
Axial force
change (%)
Axial stress
change (MPa)
CPl
MP2
CPl
MP2
CPl
MP2
3.452
3.453

0.03

0.01
0.6598
0.6623

0.38

0.02
-2.189
-2.194

0.23

0.07
1.704
1.701

-0.18

-0.04
0.2387
0.2284

-4.3

-0.18
8.6.3 Bending moments
Figure 8.6 shows the bending moment diagram of CPl for load case 4 (<|> = -45) and the
difference between the results of bending moments in M P 2 and CPl.
0.3MNm
(a) (b)
Figure 8.6 a) Bending moment diagram for frame MP2 (Load Case 4)
b) Bending moment difference between MP2 and CPl (MP2-CP1)
Table 8.7 also reports the response amplitudes of the bending moments at joint 7 of CPl
and M P 2 and their differences. Joint 7 was selected because it showed the maximum
variation of bending moment between the too frames.
"S
a
I
-Cs
.8
1
as
1
Is e
o
eb
_S
OH
I
s
3
I

I
u
1
u
is
S3 8
A
m
p

d
i
f
f
a
U
2
s
2
2
OH
u
2
OH
u
2
I
Q
03
+
o
O
Ov
en
o
U
o
O
Ov
en
o
o
CS
U
1
5
s
u
.5
o
_.
a
u
u
03
oo "1 oq r-;
*' en
; q
CO -H
33
^_ __ oo /->
r oo ^ __
vo o S
^ CS S3 S3
d o d o 9 9
en
-H
9
CS OV
en r~
en <i
oo r-
d d
-H r-
O vp
es o
d d
o
00
VO
en
en
CS vo
en ^H
5 vo
6b vo
d d
2 " r.
r~ \o r-
oo i en
r~
Ov
d
VO
Ov
d
00
cs
cs d
cs
d
o oo -H r~
oo r- - cs
~+ 1-H oo en
d o d o
38
vo cs
d d
cs cs
d d
8 3
f vo
d d
S O v
en
IT) -H
d d
r~ cs
o o
+ +
cs
-9
00 vo
CS ON
CS r+
d d
+ +
O VO Ij 00
oo * q o
99 +9
o 9g
9'-
ir> en
VO ^ H
o d
9?
r~ cs
00 O
t -H
d d
ir> r r oo r~ ov
cs
-H -t
/5fl
Chapter 8: Semi-Rigidity Effects on the Behaviour of a Frame
160
It is seen in Table 8.7 that the maximum amplitude change of bending moment is 26.7% at
the brace section and 24.8% at the chord section. There is no change of moment in diagonal
members because they were assumed to have hinged connections to the leg members. It is
noted in Figure 8.6 that there are bigger changes of bending moments from CPl to M P 2 in
the members towards the top of the tower compared to the members at the levels further
down. This is because the vertical water particle velocity is largest towards the top of the
structure. The water particle velocity becomes zero at the seabed. Therefore, horizontal
braces at the top bays will be under loadings of large magnitudes. This will, consequendy,
cause a bigger change of bending moment in the joints of top bays.
8.6.4 Dynamic characteristics
The first nine natural frequencies and mode shapes of CPl and M P 2 were calculated in
order to compare the dynamic characteristics of the two frames. Similar to U R 2 2 study, a
deck with 5000 tonnes mass was considered in the dynamic analyses. The members were
assumed unflooded but the added mass and the members' own mass were considered. The
added mass, M^, for a tubular section submerged in water and undergoing oscillation is
calculated from Equation 8.3 (Hallam, 1978).
M
am
= pnd
2
(8.3)
in which p is the water density and d is the tube diameter. The added mass in the axial
direction of a member is assumed to be zero.
The natural frequencies, reported in Table 8.8, indicate bigger variations for the frequencies
of the higher mode shapes compared to the lower modes. The maximum change is 21.7%
where the sixth natural frequency of CPl has decreased from 2.48Hz to 2.03Hz. The
second and third frequencies are due to the local modes of the inclined braces and have not
changed at all. Rest of the natural frequencies have changed less than 4%. It is, therefore,
necessary to consider joint stiffness when the load frequency might be close to a natural
frequency, in other words when the frequency margin is small to avoid coincidence of the
structural natural frequency and loading frequency. Consequently, dynamic response of the
structure will become critical and very sensitive to the loading and demands high accuracy
in analysis.
Another difference which is observed between the dynamic behaviours of CPl and M P 2 is
an extra natural frequency when the joint stiffness is modelled in the analysis. The seventh
natural frequency in M P 2 does not exist in CPl. This can be also counted as one of the
important reasons for considering the flexibility of joints in the structural analysis.
Chapter 8: Semi-Rigidity Effects on the Behaviour of a Frame 161
Mode
1
2
3
4
5
CPl
0.388
0.760
1.132
1.667
1.763
Table 8.8.
MP2
0.379
0.760
1.132
1.658
1.752
Natural frequencies of towers
Frequency ratio
(CP1/MP2)
1.024
1
1
1.005
1.006
Mode
6
7
8
9
CPlandMP2(Hz)
CPl
2.475

2.539
3.024
MP2
2.034
2.469
2.565
2.923
Frequency ratio
(CP1/MP2)
1.217

0.990
1.035
Plots of stiffness versus frequency for CPl and M P 2 are shown in Figure 8.7a and 8.7b,
where the locations at which the stiffness becomes zero are natural frequencies. The
correspondence between the natural frequencies and mode shapes of the two towers can be
constructed by investigating the mode shapes, shown in Figure 8.8 to 8.16. It is seen that
point A in Figure 8.7b, which corresponds to the 7th natural frequency, is missing from
Figure 8.7a. The mode shapes shown in Figure 8.13 correspond to the sixth natural
frequencies, co
6
, shown in Figure 8.7a and 8.7b. However, mode shape 7 of M P 2 in Figure
8.14 does not have any counterpart in CPl.
Stiffness
..--25
X10
"
10
oo 0.2 '6.4 6.6 "'"as i:6' i'._ V.4 i!- ii"""2.0 2.2 2.4 2 . 6 _ . B 3 . 0 3 . 2
Frequency (Hz)
a) CPl - model with rigid joints
Figure 8.7. Relationship between stiffness and frequency for a) CPl and b) MP2 towers (cont.)
Chapter 8: Semi-Rigidity Effects on the Behaviour of a Frame
162
Stiffness
X10-
2 9
- 1 0 T*TTTTTTTTTTTTTTT*"rn IIIIIIIM|IMIIIIII|MIIIII'M|IIIIIIIII|MMIIIII|MIM II1111 rT*T1TTT'| I I I V I I ril | TT1 *1 TTTT'I^VVI 1111 f| *1 *T* TTII | I M 1111 TF> IMIIril|l IMII lll|
0.0 0.2 0.4 0.6 0.B 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4 2.6 2.B 3.0 3.2
Frequency (Hz)
b) MPl - model with semi-rigid joints
Figure 8.7. Relationship between stiffness and frequency for a) CPl and b) MP2 towers (cont.)
M o d e shapes 1 to 9 of CPl and M P 2 are shown in Figures 8.8 to 8.16. The changes in
mode shapes are most significant for higher mode shapes. It is seen that the bouncing action
in mode shape 5 is stronger in M P 2 than in CPl. Noting mode shape 9, the amplitudes of
displacements in M P 2 are smaller than those in CPl.
Furthermore, the axial joint flexibility in mode shape 6 has resulted in bigger transverse
displacements in the top bay of M P 2 compared to CPl. This becomes important, similar to
the effect of axial flexibility on transverse deflections, when impact loads from ships and
loading barges are applied to the top of a platform.
Chapter 8: Semi-Rigidity Effects on the Behaviour of a Frame
163
a) CPl - rigid joints b) MP2 - semi-rigid joints
Figure 8.8. Mode shape 1 of towers CPl andMP2
a) CPl - rigid joints
b) MP2 - semi-rigid joints
Figure 8.9. Mode shape 2 of towers CPl andMP2
Chapter 8: Semi-Rigidity Effects on the Behaviour of a Frame
164
a) CPl - rigid joints b) MP2 - semi-rigid joints
Figure 8.10. Mode shape 3 of towers CPl and MP2
a) CPl - rigid joints b) MP2 - semi-rigid joints
Figure 8.11. Mode shape 4 of towers CPl andMP2
Chapter 8: Semi-Rigidity Effects on the Behaviour of a Frame
165
a) CPl - rigid joints b) MP2 - semi-rigid joints
Figure 8.12. Mode shape 5 of towers CPl and MP2
a) CPl - rigid joints
b) MP2 - semi-rigid joints
Figure 8.13. Mode shape 6 of towers CPl and MP2
Chapter 8: Semi-Rigidity Effects on the Behaviour of a Frame
166
This mode shape does not
exist for CPl
a) CPl - rigid joints
b) MP2 - semi-rigid joints
Figure 8.14. Mode shape 7 of towers CPl andMP2
a) CPl - rigid joints
b) MP2 - semi-rigid joints
Figure 8.15. Mode shape 8 of towers CPl and MP2
Chapter 8: Semi-Rigidity Effects on the Behaviour of a Frame
167
a) CPl - rigid joints
b) MP2 - semi-rigid joints
Figure 8.16. Mode shape 9 of towers CPl and MP2
8.7 Fatigue life estimation
The results of a structural analysis may be direcUy used for fatigue life prediction of the
tubular joints in a structure. The maximum stress range for a particular joint is calculated
through multiplying the nominal stress range of the brace meeting at the joint by the
appropriate stress concentration factor. W h e n different types of loadings are involved (IPB,
O P B or axial) an interaction formula, or conservatively, Equation (8.4) m a y be used.
A/ks = 4/a S C F
a
+ A/JPB SCFjp
B
+ A/oPB SCFQPB
(8.4)
in which A/
a
, A/n>B and A / O P B are the nominal brace stress ranges due to axial, inplane and
out of plane loadings, respectively. A/H_ is the hot spot stress range. In order to evaluate the
effect of joint flexibility on fatigue life prediction of tubular joints a typical wave loading is
considered on towers C P l and M P 2 , and the fatigue lives of joints 5 to 8 (see Figure 8.1)
are calculated. The wave specifications are taken from Barltrop (1991) and are shown in
Table 8.9. The significant wave height, H
s
, is the average height of the waves that observers
will typically report. The mean zero crossing period, T_, is obtained from the mean time
between the up-crossings of the mean water level.
Chapter 8: Semi-Rigidity Effects on the Behaviour of a Frame
168
Table 8.9. Individual wave occurrences '' in a sea state ofH
s
=3m, 7_= 7 sec, normalised to 1 year
Wave height (meters) Wave period range
Min Max 1-5 5-7 7-9
9-11 11-13 13-15 15-20
13
12
11
10
9
8
7
6
5
4
3
2
1
60
13
12
11
10
9
8
7
6
5
4
3
2
1
2
5079
96949
1
22
554
7666
58555
242227
496608
244095
0
3
69
1220
13449
91053
366080
809832
814392
194854
5
271
6707
66778
213342
99300
9
1922
42917
49993
74
11002
27451
5
5414
31317
102030 1049728 2290952 386403 94841 38527 36736
Number of waves/year 3999214
To determine the maximum stress range at a certain location on the towers, it is required to
calculate the stress range per unit wave height for each wave height. A n example of such a
relationship is shown in Figure 8.17, where the stress response has a very large increase at
the natural period.
E
25
20 -
15 -
10 -
S -
0

/
7
'*\
il
ll
| 1 . H (wave height) = 1 m
I \ / /
H = 5m
I \ / / t H -10m
I \ / / / /
H = 1Sm
i i I 1 1 1 1 1
- H 1 1 1 1 1 i "
10 12 14 16 18 20
Wave period, T (sec)
Figure 8.17. Typical relationship between wave period and stress per wave height (after Barltrop, 1991)
Since the scope of this thesis is concerning the structural aspects more than the loading
details, the stress range per unit wave height was calculated only for 1 meter high waves.
The stress range per unit height of the higher waves will be bigger than the range of the 1
meter high waves, as it is seen in Figure 8.17. Therefore, the fatigue life predictions which
Chapter 8: Semi-Rigidity Effects on the Behaviour of a Frame
169
are made here are not very accurate. This will not, however, affect the comparisons
between the fatigue lives of the joints in CPl and M P 2 towers.
In order to calculate the nominal stress values, a series of dynamic analyses were carried
out. There was a different analysis performed for every wave period and wave height,
according to Table 8.9, to construct the curves of stress range per wave height of each
joint. A phase angle of -45 was assumed for the waves. A deck mass of 5000 tonnes and
the added mass effects were also included in the analyses.
The curves of the hot spot stress range per unit wave height at joints 5, 6, 7 and 8 are
shown in Figure 8.18. Comparing the results of rigid and semi-rigid analyses, it is clearly
seen that lower stress values have been obtained in the latter case.
C P 1 (rigid joints)
o e e e o M P 2 (semi-rigid joints)
8q
Joint 6
i i i i i i i i i i i i i i i i i i i i i i i i i i i i I I i
4 6 8 10 12 14 16 18
Wave period, T (sec)
(a)
CP1 (rigid joints)
ooooo M P 2 (semi-rigid joints)
0 - i i i | i i i | i i i | i i i | i i i | i i i i i i i | i i i | i i i |
0 2 4 6 8 10 12 14 16 18
Wave period, T (sec)
(b)
Figure 8.18. Stress range per wave height of different wave periods for joints a) 5 and 6 and b) 7 and 8
The stress concentration factor equation by Efthymiou and Durkin (1985) and Equation
(8.4) were used to convert the brace nominal stress range to the hot spot stress range. The
values of S C F were calculated as: SCF
A
= 20.5 and SCF^ = 4.8.
The S-N curve T from 'Offshore Installations: Guidance on design, construction and
certification,' (DEn, 1990) with the equation _v* = 1.458xl0
12
(AoT
3
was used to evaluate
the fatigue damage corresponding to the number of wave occurrences. Table 8.10 to 8.13
report the damage per year of the joints, calculated for different wave heights and periods.
Chapter 8: Semi-Rigidity Effects on the Behaviour of a Frame 170
Table 8.10. Fatigue damage of joint 5 per year (xl0
6
)for each wave
Wave height
(meters)
Min Max 1-5
CPl MP2
5-7
CPl MP2
height and 1
Wave period range (seconds)
7-9 9-11
CPl MP2 CPl MP2
11-13
CPl MP2
wave period
13-15
CPl MP2
15-20
CPl MP2
13
12
11
110
9
8
7
6
5
4
3
2
1
60
13
12
11
10
9
8
7
6
5
4
3
2
1
0
0
0
0
0
0
0
2
8
13
6
0
0
2
6
9
4
0
0
2
29
192
711
1346
1085
236
2
0
2
22
149
554
1048
845
183
2
0
2
23
84
58
1
0
2
20
74
51
1
0
2
10
0
0
2
8
0
0
2
0
0
2
0
0
1
0
0
1
0
0
0 23 21 3603 2805 168 148 12 10
Frame CPl (includes rigid joints):
Total damage: DjXlO
6
= 3809
Fatigue life (years) =263
Model M P 2 (includes semi-rigid joints):
Total damage: DjXlO
6
= 2987
Fatigue life (years) =335
Table 8.11. Fatigue damage of joint 6 per year (x!0
6
)for each wave height and wave period
Wave height Wave period range (seconds)
(meters)
Min Max
13
12
11
110
9
8
7
6
5
4
3
2
1
60
13
12
11
10
9
8
7
6
5
4
3
2
1
1-5 5-7
7-9
9-11 11-13 13-15 15-20
CPl MP2 CPl MP2 CPl MP2 CPl MP2 CPl MP2 CPl MP. CPl MP2
0
0
0
0
0
0
0
1
5
7
3
0
0
1
3
4
2
0
0 0
Frame CPl (includes rigid joints):
Total damage: _V<10
6
= 5722
Fatipue life (years) =175
0
4
43
284
1054
1994
1607
349
3
0
3
32
216
800
1514
1220
265
2
0
4
47
171
118
2
0
3
37
133
91
2
0
4
17
1
0
3
14
1
0
3
0
0
2
0
0
1
0
16 10 5338 4052 342 266 22 18 3
Model M P 2 (includes semi-rigid joints):
Total damage: >_xl0
6
= 4349
Fatigue life (years) =230
0
1
0
Chapter 8: Semi-Rigidity Effects on the Behaviour of a Frame
171
Table 8.12. Fatigue damage of joint 7 per year (xlP
6
)for each wave height and wave period
Wave height Wave period range (seconds)
(meters)
Min M a x
1-5
5-7 7-9 9-11
11-13
CPl MP2 CPl MP2
13-15 15-20
CPl
MP2 CPl MP2 CPl MP2 CPl MP2 CPl MP2
13
12
11
110
9
8
7
6
5
4
3
2
1
60
13
12
11
10
9
8
7
6
5
4
3
2
1
0
1
1
0
1
0
0
1
9
67
242
365
162
3
0
1
9
67
242
365
162
3
0
4
47
315
1169
2211
1782
387
3
0
3
34
228
847
1602
1292
281
2
0
2
25
89
62
1
0
2
18
65
45
1
0
1
6
0
0
1
5
0
0
1
0
0
1
0
0
0
0
0
0
0
1 849 849 5918 4289 179 131
1 1
0 0
Frame CPl (includes rigid joints):
Total damage: DgXlO
6
= 6956
Fatigue life (years) = 144
Model M P 2 (includes semi-rigid joints):
Total damage: Dp<10
6
= 5277
Fatigue life (years) = 190
Table 8.13. Fatigue damage of joint 8 per year (xlP
6
)for each wave
Wave height
(meters)
Min M a x 1-5
CPl MP2
5-7
CPl MP2
height and
Wave period range (seconds)
7-9 9-11
CPl MP2 CPl MP2
11-13
CPl MP2
wave period
13-15
CPl MP2
15-20
CPl MP2
13
12
11
:io
9
8
7
6
5
4
3
2
1
60
13
12
11
10
9
8
7
6
5
4
3
2
1
0
0
0
0
0
0
0
0
0 0
2 1
25 17
186 132
670 473
1010 713
447 316
8 6
2348 1658
Frame CPl (includes rigid joints):
Total damage:
Fatigi ie life (years)
DgXlO
6
= 5609
= 178
0
2
25
170
629
1191
960
208
2
3187
0
2
20
131
485
917
739
161
1
2456
0
1
10
36
25
0
72
0
1
9 0
33 1
23 3
0 0
66 4
0
1 0 0 0
3 0 0 0
0 0 0 0
4 0 0 0
Model M P 2 (includes semi-rigid joints):
Total damage:
Fatigue life (years)
D_xl(r = 4184
= 239
0
0
0
0
Chapter 8: Semi-Rigidity Effects on the Behaviour of a Frame
172
The fatigue lives obtained for the joints 5, 6, 7, and 8 and their changes after the joint
stiffness was considered are shown in Table 8.14. The fatigue life estimates of all four joints
have increased about 3 0 % on average in M P 2 model where the joints were assumed to be
flexible. However, this result is just an example and shows the significance and possible
advantage of considering stiffness of joints in analysis. Further study is required to draw
general conclusions.
Table 8.14. Fatigue life predictions and their changes when joint stiffness is considered in analysis
Joint Fatigue life (years) Ratio
CPl (rigid joints) M P 2 (semi-rigid joints) (CP1/MP2)
5 263 335 1.27
6
175 230 1.31
7 144 190 1.32
8 178 239 1.34
8.8 S u m m a r y
The aspects of the structural behaviour investigated in this chapter, to recognise the effects
of joint flexibility on the analysis of a 100m fixed offshore tower, were:
1) global deflections,
2) axial forces,
3) bending moments, and
4) dynamic characteristics.
There was almost no change of lateral deflections when distributed loads were acting on the
towers. However, under a point load the flexibility of the joints in the vicinity of the load
can generate extra deformations which may be up to 1 0 % of the maximum deflection of the
tower.
The axial forces were almost the same when joint stiffness was considered in the analysis.
This equality exists mainly because there was a unique load path for axial loads. However,
further studies are required on the structures with redundant members and different
geometrical configurations.
W h e n the stiffness of the joints was included in the analysis, bending moments had
variations up to 2 7 % at the brace sections and 2 5 % at the chord sections . The overall
changes of bending moments were more significant in the leg members than in the brace
members.
Chapter 8: Semi-Rigidity Effects on the Behaviour of a Frame 173
The flexibility allowance in the dynamic analysis altered the higher natural frequencies the
most The sixth frequency decreased by up to 2 2 % . The first natural frequency decreased by
only 2.4%. Furthermore, there was an extra m o d e shape when the joint stiffness was
considered in the analysis.
Fatigue life of the tubular joints increased by up to 3 0 % on average when flexibility was
considered in the analysis. The results obtained herein indicate the significance of joint
flexibility effects on the fatigue life of offshore structures. Further, analytical and experimen-
tal studies on the structures with different geometrical configurations are required to
produce more evidence and to draw conclusions on the effects of joint stiffness
consideration on the fatigue life prediction of tubular joints.
Chapter 9
SUMMARY AND CONCLUSIONS
Chapter 9: Summary and Conclusions
175
9.1 Summary
Offshore steel structures are benefited extensively from the advantageous properties of
tubular sections. However, thin walls of tubes create some problems in design and analysis
of tubular joints, such as: joint flexibility, ultimate strength, stress concentration, and
fatigue. Perhaps ultimate strength and fatigue impose the most crucial criteria on the
methods used for design of tubular joints. However, joint flexibility and stress concentration
are directly involved in the calculation of stress values, and consequendy influence the
estimation of the fatigue life of a tubular joint
Offshore structures and particularly tubular joints are introduced in Chapter One. The scope
of this thesis is explained and the views of different codes of standards about joint flexibility
and its consideration is quoted. Tubular joints are commonly assumed to be rigid in the
structural analysis, and the main reason for this assumption is the lack of an efficient method
of joint modelling (Barltrop, 1991). The highlights of the work carried out in this thesis are
as follows:
1) an investigation into the flexibility behaviour of tubular T-joints using natural
frequencies as an indication of joint rigidity. The stiffness matrix of a beam with semi-
rigid ends and a continuous mass property is calculated for two cases of bending
deformations only and bending plus shear deformations. The dynamic approach is for
the first time used herein to calculate the stiffness of tubular joints.
2) Establishment of parametric formulae for stiffness of tubular T-joints based on the
natural frequencies obtained from Finite Element analysis. The modes considered are
inplane bending, out of plane bending, and axial deformation of brace.
3) A n investigation into the effect of dynamic strain reading on the stress or strain
concentration factor of tubular T-joints. This parameter is being presently ignored in
design practice and research.
4) A study into the effect of joint flexibility consideration on the fatigue life of tubular
joints in offshore structures. Tubular joints are presently assumed to be rigid when
designed for fatigue.
The work on the stiffness or flexibility of joints, in general, has been started in the 1910s.
However, it was completely related to the joints made of open sections. Tubular joints have
been attended, though, in the 1970s and the main reason has been their application in the
Chapter 9: Summary and Conclusions 775
offshore structures. As, this thesis focuses on the tubular joints in particular, thus the
literature survey concentrates on these types of joints. The different investigation methods
of tubular joints is described in Chapter T w o . This chapter reports a review of the available
literature on the flexibility of tubular joints. The various methods of analysis used to study
the behaviour of tubular joints are described as:
1) analytical methods,
2) experimental and semi-experimental methods, and
3) numerical methods.
Analytical procedures are still used by researchers in view of their numerical accuracy and
theoretical bases. The only known experimental work on the stiffness of tubular joints,
which is on araldite specimens by Fessler (1981), was described. Most studies have used the
Finite Element (FE) method to determine the joint stiffness and they hardly include any
comparison with experimental data. All parametric formulae for the stiffness of tubular
joints have been derived by curve fitting the measured data. A semi-experimental model may
establish a better mathematical form for the stiffness formula (refer to Section 2.3.2).
The studies on the effects of flexibility consideration on the behaviour of tubular structures
are also reviewed in Chapter T w o . Bending moments are influenced the most, when
flexibility of the joints is introduced in the analysis. However, the literature survey shows
various effects on the same response in the structures, implying that the effects of joint
flexibility depends on the type, geometry and configuration of the structure. There was only
one study into the joint stiffness effects on fatigue estimate which was carried out on two
simple one-storeyed frames (Recho et al., 1990).
Regarding stiffness determination of joints, the aim of most studies is to calculate an
equivalent spring stiffness for the joint model. There is also another approach to define a
non-dimensional rigidity factor between 0 and 1 to state the degree of fixity of a joint
The research on the stiffness aspects of tubular joints, on one hand, has concentrated mostiy
on the joints as a discrete element of a structure. O n the other hand the effects of joint
stiffness on the behaviour of structures have only been studied analytically. The high cost of
experiment perhaps has prevented testing of large scale specimens.
This study offers Dynamic Deformation method to derive the stiffness matrix of a beam with
semi-rigid ends. The formulations developed include bending deformations and/or shear
deformations. In these formulations, the inertia forces have been modelled as continuous
properties. The method of analysis in this thesis uses the resonant frequencies to calculate
the stiffness of a joint. A new definition is given for the rigidity factor of joints compatible
with the moment rotation relationship of beams with semi-rigid joints. The stiffness matrix
Chapter 9: Summary and Conclusions 777
of an axially loaded member with flexible ends is also derived. A definition is introduced for
the rigidity factor of a joint loaded axially.
The definitions introduced for joint rigidity factor, a number between 0 and 1 to express the
stiffness of a joint as a percentage of full fixity are:
Bending stiffness: v
b
= J=J> Axial stiffness: v
a
=
l
-=rr~
The experimental part of this study includes eleven tubular T-joints tested for inplane and
out of plane bending modes. The test specimens are selected to bound the c o m m o n ranges
of parameters used in offshore structures. A survey is conducted on the geometry of tubular
joints to recognise these ranges.
The dynamic experimental method adopted here, uses the measured natural frequencies of
the joints. The main piece of equipment in the experiments was a Fourier analyser which
was employed to measure the natural frequencies. Test results are reported and compared
with the theoretical values from Dynamic Deformation method when joints are considered
to be rigid. There is a significant difference observed between the measurements and
calculated natural frequencies by Dynamic Deformation method which is attributed to the
joint flexibility. Various methods, including computer analysis and modal analysis, were
used to confirm the correspondence between the natural frequencies and mode shapes of a
joint. W h e n natural frequency of a mode of vibration is measured, assurance should be made
that it corresponds to the desired mode shape.
The stiffness method was used to study the effect of support conditions on the natural
frequencies of T-joints in inplane bending and out of plane bending modes. The results
indicated that there is hardly any effects due to the support conditions on the natural
frequencies of a T-joint in inplane bending mode. Out of plane bending, however, requires
fixed support conditions, otherwise instability will occur for the joint specimen. Despite the
minimal effects of support conditions on the inplane bending vibrations, a special
arrangement was used to minimise the rotational fixity of the chord supports.
The dynamic method used for flexibility determination of tubular joints can be performed
easily and inexpensively. The shortcomings of this method are:
1) the need of a suitable arrangement for support conditions in some deformation
cases, and
2) fairly high capital cost of measurement equipment involved in order to conduct the
dynamic tests.
Chapter 9: Summary and Conclusions ^ ^ ^ ^ ^ ^
1?Q
The advantages of the method are:
1) its ability to capture the stiffness and mass properties of a joint, including any
deficiencies and imperfections such as those created during manufacturing or those
caused after structure is being in service.
2) The speed of the test procedure can be counted as another advantage over the
static methods of testing.
The parametric formulae developed in this study for the stiffness of tubular T-joints are
based on the results of an extensive Finite Element analysis. Therefore, Chapter Five is
dedicated to the Finite Element method and the general steps used to model the T-joints.
There is a brief look at the different types of F E techniques to model the shell structures.
Plate elements and degenerated shell elements have been used successfully in the studies
before. The plate element is used herein to model the T-joints. The computer package
ALGOR used for conducting the F E analyses is introduced. A bench mark study on the plate
element in ALGOR is performed by modelling a cylinder under internal pressure. The results
of F E analyses by ALGOR show a good agreement with the calculated results from the
theory of elasticity. The performance of the three dimensional solid elements in modelling of
a T-joint was also investigated. It was found that plate elements can make more realistic
predictions of the natural frequencies of a T-joint compared to three dimensional elements.
The effect of support conditions on the frequencies of in plane bending mode was also
investigated through F E analyses. It was observed that the effects of support conditions are
even less than what stiffness method showed in chapter Four.
The results of the F E analyses of the test specimens are reported in Chapter Five. The mean
absolute error of the predicted natural frequencies by F E models were 3.1% compared to
the measurement results. This indicates that the Finite Element meshes used to model the
joints were appropriate. The agreement of out of plane bending results was not as good as
inplane bending for the T-joints with large brace diameters. The main reason is attributed to
the ground connection of the support assembly.
The description of the parametric study of the joint stiffness is given in Chapter Six. There
were totally 270 F E analyses, covering the c o m m o n ranges of joint parameters used in
offshore structures. The F E analyses included three modes of deformations in tubular joints,
i.e. inplane bending, out of plane bending and axial deformation of brace. The chord length
to diameter ratio was chosen to be 6.85. The results of the F E analyses for all three modes
indicate that the joints with large brace diameters or small chord diameter to thickness ratio
possess a relatively high stiffness. The results also show that these types of joints are most
sensitive to non-dimensional parameters such as dJD or DI1T in which D and T are diameter
and thickness of the chord, and d is the diameter of the brace. The effect of thickness ratio
(t = tIT) on the stiffness of T-joints is presently ignored. This effect is investigated for the
Chapter 9: Summary and Conclusions fjg
first time in this study. The results of the FE analyses indicate a significant change of joint
stiffness when x is taken into account.
A set of equations are established for the stiffness of T-joints based on F E results which
include X ratio in addition to other parameters. The curve fitting of the F E results has
maximum errors of 2 5 % , 1 8 % and 2 9 % error in inplane bending, out of plane bending, and
axial stiffness formulae, respectively. The stiffness of tubular T-joints varies with x ratio as
follows:
1) inplane bending stiffness: p(r/m+i.s5)
x
o._s ^
2) out of plane bending stiffness: x
21
, and
3) axial stiffness: p(-Y/27.3
+
o.6T)
T
o.9i
A good agreement was obtained when the results of the developed equations herein and
those from other studies were compared. The average absolute differences of 15.4%, 15.2%
and 27.4% were observed from the comparisons of the joint stiffness values in each mode of
deformation i.e. inplane bending, out of plane bending and axial deformation of brace,
respectively.
The concept of stress concentration factor in analysis and design of tubular joints is
presented in Chapter Seven. Various methods of investigations on the stress or strain
distribution of tubular joints are briefly described. The most popular procedures are strain
gauge reading and Finite Element method. There are variations observed and reported for
the calculated or measured stress concentration factors by other researchers. The reasons
for such variations are also presented. The reliability approach to the stress concentration
factor and the parameters involved are stated. In this approach different design values of
S C F are selected according to the level of confidence required in different applications e.g.
primary or secondary joint or structure (Tebbett and Lalani, 1984). Dynamic strain reading
is discovered as a factor that introduces some variations to the S C F of tubular joints. This
variable is presently ignored in design practice of offshore structures and even in research
works. A series of tests were conducted in this thesis to study the effects of load frequency
on the strain concentration factor of T-joints. The test results show a significant variation of
strain concentration factor with respect to load frequency. A series of analyses were also
carried out using the stiffness method to study the dynamic effects of loading on the strain
concentration factor. The findings of stiffness method analyses indicate a variation in strain
concentration, however it was not as high as the experimental results.
A work has been initiated on the relationship between joint stiffness and stress concentration
factor of tubular T-joints (Kohoutek and Hoshyari, 1993). This thesis has not included this
topic, but the data used in the published paper are taken from this study.
Chapter 9: Summary and Conclusions ^ ^ ^ ^ 180
The effects of considering joint stiffness on the behaviour of an offshore tower is studied in
Chapter Eight. The tower modelled is 100 meter high and represents the structures in the
North Sea. The deflections, axial load and bending moment distributions under static loads
are compared between two analyses, one with rigid joints and one with semi-rigid joints.
The first nine natural frequencies and mode shapes are also calculated and compared. The
geometry and specification of the tower and the applied loadings are taken from the report
U 2 2 by Underwater Engineering Group (1984). There are four load cases considered in the
analyses. Load case 1 consists of a horizontal point load at the top of the tower for studying
the deflection changes. Load cases 2 to 4 are derived from a 100 year storm wave passing
through the tower with different phase angles. The programs used in this chapter for
structural analysis and wave force calculation are written by the author. The theoretical
model described in Chapter Three is employed in the structural analysis program used
herein. A deck mass of 5000 tonnes and also added mass effects were included in the
dynamic analyses.
A fatigue analysis was also carried out on several joints of the towers, and the results were
presented. A wave specification in a particular sea state was used to calculate the maximum
stress range per unit wave height of the joints under study. The results indicate a significant
change of fatigue life estimation in tubular joints when the stiffness of the joints are
introduced in the analysis.
Chapter 9: Summary and Conclusions
181
9.2 Conclusions
In connection with the studies carried out in this thesis there are a number of conclusions
that can be made as follows:
1) At present, flexibility of joints is not included in the analysis of offshore structures, due to
the lack of an efficient method of joint modelling.
2) According to the literature survey, in spite of the importance of fatigue criteria in design
of tubular joints in offshore structures, the research on the joint stiffness has not covered the
fatigue life changes. Furthermore, the adverse effect of fatigue fracture on the stiffness of
tubular joints and subsequentiy its influence on the behaviour of whole structure have not
been investigated before.
3) The literature survey shows that there is no experimental study into the effects of joint
stiffness on the behaviour of offshore structures.
4) Although the joints with p = 1 are extensively used in offshore structures, most
researchers in this field have concentrated on the joints with a range of P ratio between 0.2
and 0.8.
5) The F E analyses indicate that there is effectively no influence from support conditions on
the natural frequencies of a T-joint under inplane bending vibration.
6) The fixity of supports was critical for the joints with p = 1 in out of plane bending mode.
The worst correlation between F E analysis and test results was also for this type of joints.
Support conditions of the test specimens became most critical for the stiffer joints. These
conditions should be as close as possible to those assumed in the theoretical model.
7) The inclusion of X ratio in the parametric formulae for stiffness of tubular joints is
recommended, especially for axial loading case. The thickness ratio x makes a difference of
up to approximately 4 0 % for out of plane and inplane bending stiffness, and 2 3 0 % for axial
stiffness of brace.
8) The parametric equations, which are presently used to determine the hot spot stress
values at a tubular joint, are all based on the databases from static experiments. These
equations do not differentiate between dynamic loads, which are applicable to the offshore
platforms, and static loads.
9) There is a variation of strain or stress concentration factor in tubular joints when
measurements or calculations are carried out in a dynamic fashion. For the range of
frequencies examined in this study (0.5 to 50 Hz), the following variations observed on the
strain concentration factors of tubular T-joints:
Chapter 9: Summary and Conclusions
182
inplane bending mode: 100%,
out of plane bending mode: 120%.
These values are measured at the frequencies much higher than the natural frequency of an
offshore platform. However, they indicate the variation which exist for stress concentration
factor due to the load frequency.
10) It is recommended that a frequency dependent factor or conservatively a constant factor
to be included in the parametric formulae for stress concentration factors of tubular joints to
take the effect of dynamic loading into account
11) The results of the analyses carried out on two towers, one with semi-rigid joints and the
other with rigid joints, to recognise the effects of joint flexibility are as follows:
a) The change in the maximum lateral deflection is negligible when wave loading is
applied to the towers. However, when a point load is acting, the flexibility of the
joints in the vicinity of the load can generate extra deflections which may be up to
1 0 % of the maximum deflection of the tower.
_>) Axial forces were almost the same for the towers analysed.
c) Bending moments had a maximum variation of 2 7 % at the brace sections and 2 5 %
at the chord sections. The overall change in bending moments were more significant
for the leg members than for the brace members.
d) The flexibility allowance in the dynamic analysis altered the higher natural
frequencies more significantly than the lower frequencies. A n extra mode shape was
also generated when the joints were assumed to be semi-rigid in the analysis.
12) A mean increase of 3 0 % resulted for the fatigue life estimates of tubular joints when
joint stiffness was considered in the analysis.
Chapter 9: Summary and Conclusions ^ 183
9.3 Future research work
Further investigation is required on the following topics to achieve more insight to the
stiffness of tubular joints and its effects on the behaviour of offshore steel structures:
1) Investigation of the effects of fatigue fracture on the stiffness of tubular joints and
consequently the whole structure.
2) Further study on the stiffness of the joints with diameter ratios greater 0.8. Diameter
ratio varies between 0.2 and 1.
3) In order to verify and substantiate the works already conducted on the stiffness of tubular
joints, there is a need to investigate the effects of joint flexibility on the different aspects of
structural behaviour by conducting experiment on reasonable sized frame specimens.
4) Further investigation is required to be able to draw conclusions on the relationship
between stress concentration factor and joint stiffness of tubular joints.
5) Experimental study on the axial stiffness of tubular joints.
6) Application of the dynamic method of joint stiffness determination used in this thesis for
other types of tubular joints such as Y, K, etc..
7) Further investigation into the effects of dynamic loading on the stress concentration
factor in tubular joints.
8) Further investigation into the effects of joint stiffness on the behaviour of offshore towers
with various configurations and geometries. This shall include the structural response of
three dimensional frames and also fatigue life estimate.
References 754
REFERENCES
Abel A. (1982). "Joints in tubular steel structures," Symposium, Research on Marine
Structures, University of Sydney, April.
Algor Processor Reference Manual, August 1989, Algor Interactive Systems, Inc.
American Petroleum Institute (1991), "Recommended practice for planning, designing and
constructing fixed offshore platforms," API RP2A, Nineteenth Edition, August 1, 1991.
Bathe KJ. (1982). "Finite element procedures in engineering procedures," Prentice-Hall
Inc.
Barltrop N. D. P. and Adams A. J. (1991). "Dynamics of Fixed Marine Structures"
Butterworth Heinmann, Third Edition.
Beale L.A. and Toprac A.A. (1967). "Analysis of in-plane T, Y and K welded tubular
connections," Welding Research Council Bulletin 115, October.
Behrens M., Kohoutek R., (1990). "Dynamic of Beam with Semi-rigid joints, Part II-
Experimental evaluation," Proceedings of Australian Vibration and Noise Conference, held
at Monash University 18-20 September, pp 344-348.
Bijjlard P.P. (1955). "Stresses from Radial Loads and external moments in cylindrical
pressure vessels," Welding Journal, pp. 608s-617s.
Bijlaard P.P. (1954). "Stresses from Radial Loads in Cylindrical Pressure Vessels," Welding
Journal, pp. 615s-623s.
Billington C.J., Lalani M. (1987). "Recent research and advances in the design of tubular
joints," Proceedings of the International Conference on Steel and Aluminium Structures,
Cardiff, U K , 1987, Steel Structures, Edited by RNarayanan.
Bjorhovde R., Colson A. and Brozzetti J. (1990) "Classification system for beam-to-column
connections," Journal of Structural Engineering, A S C E , Vol 116, No. 11, November.
Bouwkamp J.G., Hollings J.P., Maison B.F., Row D.G. (1980). "Effects of joint flexibility
on the response of offshore towers," 11th annual Offshore Technology Conference, paper
3901.
British Standards Institution (1982), Code of practice for fixed offshore structures, BS
6235.
Buitrago J., Zetdemoyer N., Kahilich J.L. (1984). "Combined hot-spot stress procedures for
tubular joints," 15th annual Offshore Technology Conference, paper 4775.
Chaudhury, Dover W.D. (1985). "Fatigue analysis of offshore platforms subject to sea wave
loadings," International Journal of Fatigue, pp. 13-19.
Chen B., Hu Y., Tan M. (1990). "Local joint flexibility of tubular joints of offshore
structures," Marine Structures, Elsevier Science Publishers Ltd, England, pp 177-197.
References 185
Chen T., W u S., Yang L. (1990). "The flexibility behaviour of tubular joints in offshore
platforms," Proceedings of the 9th International Offshore Mechanics and Arctic
Engineering Symposium, Publ by A S M E , pp. 307-312.
Matsui C, Morino S. and Kawano A. (1984). "Lateral-torsional buckling of trusses with
rectangular tube sections," Welding of Tubular Structures, Proceedings of the 2nd
International Conference.
Clough R.W., Penzien J. (1975). "Dynamics of structures? McGraw-Hill Book Company.
Clough R.W., Felippa CA. (1968). "A refined quadrilateral element for analysis of plate
bending," Proceedings of the Second Conference on Matrix Methods in Structural
Mechanics, Parti of 2, pp 399-440.
Connolly M. P., Hellier A. K., Dover W. D., Sutomo J. (1990). "A parametric study of the
ratio of bending to membrane stress in tubular Y and T-joints," Int. J. Fatigue , Vol 12, pp.
3-11.
Dawson T.H. (1983). "Offshore structural engineering," Prentice-Hall Inc., USA.
De Jong H., Wardenier J. (1987). "The effect of joint rigidity on the buckling behaviour of
compressed tubular members in trusses and frames," Proceedings of the International
Conference on Steel and Aluminium Structures, Cardiff, U K , 1987, Steel Structures, Edited
by R.Narayanan.
Demirbilek Z. (1989). 'Tension Leg platform: An overview of the concept, analysis, and
design," in Tension Leg Platform edited by Demirbilek Z., Published by American Society
of Civil Engineers.
Det Norske Veritas (1977). "Rules for the Design, Construction and Inspection of Offshore
Structures" Reprint 1981.
Dharmasavan S., Dover W. D. (1985). "Stress distribution formulas and comparison of 3
stress analysis technique for tubular joints," Transactions of the ASME, Vol 107, March.
Dharmasavan S., Dover W. D. (1981). "Stress concentration factors for tubular Y-joints,"
Ind International Conference on Integrity of Offshore Structures, Paper 16, University of
Glasgow.
Digre K.A., Brasted L.K., Marshall P.W. (1989). "The design of the Bullwinkle Platform,"
21 st annual Offshore Technology Conference, paper 6050.
Efthymiou M Durkin S. (1985). "Stress concentrations in T/Y and gap/overlapped k-
joints," Developments in Marine Technology, Vol 2, pp. 429-440.
Efthymiou M. (1985). "Local rotational stiffness of unstiffened tubular joints," Koninldijke/
Shell Exploratie en Produktie Laboratorium Report RKER.85.199.
Efthymiou M (1988). "Development of SCF formulae and generalised influence functions
for use in fatigue analysis," Proceedings of the Conference on recent developments in
tubular joints technology, Offshore Tubular Joints Conference, October, Surrey..
References 186
Fessler H., Spooner H. (1981). "Experimental determination of stiffness of tubular joints,"
2nd International Conference on Integrity of Offshore Structures, paper 28, University of
Glasgow.
Fessler H., Mockford P.B., Webster J.J., (1986). "Parametric equations for the flexibility
matrices of multi brace tubular joints in offshore structures," Proceedings of Institute of
Civil Engineers, Part 2, 81, December.
Fessler H., Mockford P.B., Webster J.J., (1986). "Parametric equations for the flexibility
matrices of single brace tubular joints in offshore structures," Proceedings of Institute of
Civil Engineers, Part 2, 81, December.
Forsyth P., Tebbett I.E. (1988). "New test data on the strength of grouted connections with
closely spaced weld beads," 20th annual Offshore Technology Conference, paper 5833.
Gibstein M.B., Moe E.T. (1981). "Numerical and experimental stress analysis of tubular
joints with inclined braces," International Conference on Steel in Marine Structures, Paris,
paper 6.3.
Gibstein M.B. (1978). "Parametric Stress analysis of T joints," European Offshore Steels
Research Seminar, paper 26, Cambridge.
Graff W.J. (1981). "Introduction to offshore structures," Gulf Publishing Company.
Grant J.E. (1968). "Dynamic response of steel structures with semi-rigid connections,"
PhD Thesis, Oregon State University.
Greste O. (1970). "Finite Element analysis of tubular K-joints " Ph.D. thesis, The University
of California Berkeley.
Gumey T.R. (1981). "Some comments on fatigue design for offshore structures," 2nd
International Conference on Integrity of Offshore Structures, paper 14, University of
Glasgow.
Hallam M. G., Heaf N.J. and Wooton L.R. (1978). "Dynamics of Marine Structures,"
CIRIA Underwater Engineering Group, U R 8 , 2nd Edition.
Head J.L., Tilley D.R. (1980). "Stress analysis of a welded tubular T-joint," Ind Annual
International Conference on BOSS, paper 1820, Dallas, Texas.
Hellier A. K., Connolly M. P., Dover W. D. (1990). "Stress Concentration Factors for
Tubular Y- and T-Joints," Int. J. Fatigue, Vol 12, pp. 13-23.
Hellier A. K., Connolly M. P., Kate R.F. , Dover W. D. (1990). "Prediction of Stress
Distribution in Tubular Y- and T-Joints," Int. J Fatigue, Vol 12, pp. 25-33.
Hellier A.K., Connoly M.P., Dover W.D., Corderoy D.J.H. (1990). "Parametric Equations
to Predict the Full Scale Stress Distribution in Tubular Welded Y and T-Joints,
Proceedings of the First Pacific/Asia Offshore Mechanics Symposium, Seoul, Korea, pp.
281-293.
References 187
Hoffman R.E., Sharifi P. (1980). "On the accuracy of different finite element types for
analysis of complex welded tubular joints," 12th annual Offshore Technology Conference,
paper 3691.
Holmas T., Remseth S., Hals T.E. (1985). "Approximate flexibility modelling of tubular
joints in marine structures," SINTEF Report No. STF71 A85016.
Hoshyari I. and Kohoutek R. (1993). "Rotational and Axial Flexibility of Tubular T-joints,"
The Third International Offshore and Polar Engineering Conference, Singapore, 6-11
June, pp 192-198.
Hoshyari I. and Kohoutek R. (1993). "Offshore Tower Analysis Considering Semi-Rigid
Joints," 13th Australian Conference on the Mechanics of Structures and Materials,
University of Wollongong, 5-7 July, pp 389-396.
Huang H.C. (1989). "Static and dynamic analysis of plates and shells," Springer-Verlag.
Irons B.M., Ahmad S. (1980). "Techniques of finite element" John Wiley & Sons.
Irvine N.M. (1981) "Comparison of the performance of modem semi-empirical parametric
equations for tubular joint stress concentration factors," Special and Plenary Sessions: Steel
in Marine Structures, International Conference, Paper 6.5.
Jones S.W., Kirby P.A., Nethercot D.A. (1983). "The analysis of frames with semi-rigid
connections- A state-of-the-Art report," Journal of Constructional steel Research, Vol 3,
N o 2, pp 2-13.
Kawashima S., Fujitomo T. (1984). "Vibration analysis of frames with semi-rigid
connections," Computers & Structures, Vol 19, No. 1-2, pp. 85-92.
Knowles N.C., Razzaque A., Spooner J.B. (1976). "Experience of finite Element analysis of
shell structures," Finite Elements for thin Shells and Curved Members, Edited by D.G.
Ashwell and R.H. Gallagher, John Wiley & Sons.
Kohoutek R. (1980). "Spring-hinged beam," Report RR/Struct/04/1980, Department of
Civil Engineering, University of Melbourne, p 23.
Kohoutek R. (1985a). "Dynamic analyses of frames with semi-rigid joints," IEAust., Metal
Structures Conference, held at Monash University, Melbourne, L. Pham, ed., pp 7-10.
Kohoutek R. (1985b). "Analysis of beams and frames," Chapter 4 in Analysis and Design of
Foundations for Vibrations, pp 99-156; P. Moore, ed., 512pp, 1985, published by
A_A.Balkema,
Kohoutek R. (1990). "Dynamic of beam with semi-rigid joints, Part I-Analytical Model,"
Proceedings of Australian Vibration and Noise Conference; held at Monash University 18-
20 September, pp 339-343.
188
Kohoutek R. (1991a). "Dynamic tests of semi-rigid connections," Proceedings of The
Second International Workshop on Connections, Invited paper, Pittsburgh, April 10-12, p 9.
Kohoutek R. (1991b). "Analysis of frames for stability," Proceedings of The Second
International Workshop on Connections, Invited paper, Pittsburgh, April 10-12,9pp.
Kohoutek R. and Hoshyari I. (1991). "Dynamics of Tubular Semi-Rigid Joints,"
Proceedings for Offshore Mechanics and Arctic Engineering, Stavanger, Norway, June 23-
28, Vol. I, Part B, pp 579-585.
Kohoutek R. and Hoshyari I. (1991). "Flexibility of Tubular Joints in Offshore Structures,"
The First International Offshore and Polar Engineering Conference, Edinburgh, United
Kingdom, August 11-15, pp 61-66.
Kohoutek R. and Hoshyari I. (1992). "Parametric Formulae of Rigidity for Semi-Rigid
Tubular Joints," International Conference on Offshore Mechanics and Arctic Engineering,
Calgary, Canada, Vol. I, Part B, June, pp 605-611.
Kohoutek R. and Hoshyari I. (1992). "Dynamic Strain Measurement of Tubular T-Joints,"
Dynamic Loading in Manufacturing and Service, The Institutions of Engineers Australia,
Melbourne, 9-11 February, pp 131-136.
Kohoutek R. and Hoshyari I. (1993). "Relationship Between Rigidity and Stress
Concentration Factor for Semi-Rigid Tubular Joints," 12th International Conference on
Offshore Mechanics and Arctic Engineering, Glasgow, Scotland, 20 -24 June.
Kolousek V., (1973). "Dynamics in engineering structures," Praha-Academia, London-
Butterworth.
Kuang J.G., Potvin A.B., Leick R.D. (1975). "Stress concentration in tubular joints," 7th
annual Offshore Technology Conference, paper 2205.
Kurobane Y., Ogawa K., Ochi K. (1989). "Recent research developments in the design of
tubular joints," Journal of Constructional Steel Research, Vol 2, pp. 169-188.
Lalani M., Tebbett I.E. and Choo B.S. (1986) "Improved fatigue life of tubular joints," 18th
Offshore Technology Conference, paper 5306.
Lee G.C. (1968). "Offshore structures past, present, future and design consideration,"
Offshore, June 5, pp. 45-55.
Lionberger S.R. and Weaver W.Jr. (1969). "Dynamic response of frames with non-rigid
connections," Journal of the Engineering Mechanics Division, Proceedings of the American
Society of Civil Engineers, pp. 95-114.
Lothers J.E. (1960). "Advanced design in structural steel," Prentice-Hall, Inc.
References
189
M a S.Y., Tebbett I.E. (1988). "Estimations of stress concentration factor for fatigue design
of welded tubular connections," 20th annual Offshore Technology Conference, paper 5666.
Marshall Peter W. (1984). "Connections for welded tubular structures," FWP Journal, Vol
24, N o 10, Oct, pp. 7-22.
Marshall Peter W. (1984). "Connections for welded tubular structures," FWP Journal, Vol
24, Noll, Nov, pp. 59-78.
Monforton G.R. and Wu T.S. (1963). "Matrix analysis of semi-rigidly connected frames,"
Journal of the Structural Division, Proceedings of the American Society of Civil
Engineers, pp. 13-42.
Nethercot D.A. (1986). "The behaviour of steel frame structures allowing for semi-rigid
joint action," in Steel Structures, edited by M.N. Pavlovic, Elsevier Science Publishers, pp.
135-151.
Radenkovic D. (1981). "Stress Analysis in Tubular Joints," Special and Plenary Sessions:
Steel in Marine Structures, International Conference, Paris, Paper PS1, pp 53-93.
Rathbun J.C. (1936). "Elastic properties of riveted connections," Transaction of American
Society of Civil Engineers, 101, pp 524-563.
Recho N., Ritty B. and Ott F. (1990). "Numerical influence of flexibility on the fatigue
design of T-welded tubular joints," Proceedings of the 9th International Offshore
Mechanics and Arctic Engineering Symposium, Publ by A S M E , pp. 365-371.
Rodabaugh E.C. (1980). "Review of data relevant to the design of tubular joints for use in
fixed offshore platforms," Welding Research Council, Bulletin No. 256, January.
Sakamoto S. and Minoshima N. (1979). "Behaviour of T-joint of Octagonal tubes,"
Abstract, Annual Meeting ofAIJ, 1037-1038 (in Japanese).
Scordelis A.C., Bouwkamp J.G. (1970). "Analytical study of tubular tee-joints," Journal of
the Structural Division, Proceedings ofASCE.
Shinners C. D., Abel A. (1982). "Stress analysis of large scale tubular T-joints," Mechanical
Engineering Transactions - Institution of Engineers Australia, Vol M E 7 , N o 2, Jun, pp. 62-
67.
Simpson R.J., Venables R.K., Lalani M., Tebbett I.E., (1983). 'Towards a more rational
approach for the design of tubular joints in steel offshore structures," Design in Offshore
Structures, Thomas Telford Ltd, London, pp 77-83.
Smedely G.P. (1977). "Peak strains at tubular joints," Institution of Mechanical Engineers
Seminar, Corrosion fatigue in offshore installations, London, Sep.
Smedley P., Fisher P., (1990). "A Review of Stress Concentration Factors for Tubular
Complex Joints," Integrity of Offshore Structures-4, Glasgow.
References
190
Smedley P., Fisher P. (1991) "Stress Concentration Factors for Simple Tubular Joints,"
Proceedings of the First International Offshore and Polar Engineering Conference
Edinburgh, U K , pp. 475-483.
Souissi R. (1990). "Flexibility of tubular joints," in Tubular Structures Edited by E. Nierni
and P. Makelainen, Elsevier Applied Science.
Springfield C.W., Brunair R.M. (1989). " End-fixity determination of vibrating structural
components," Computers and Structures, Vol. 33, No. 2, pp. 453-458.
Steel Structures Research Committee (1931), First report, Department of scientific and
Industrial Research, H M S O , London.
Steel Structures Research Committee (1934), Second report, Department of scientific and
Industrial Research, H M S O , London.
Steel Structures Research Committee (1936), Final report, Department of scientific and
Industrial Research, H M S O , London.
Sterling G.H., Krebs J.E., Dunn F.B. (1989). "The Bullwinkle Project: An Overview," 21st
Offshore Technology Conference, paper 6049, pp 53-62.
SMS (Structural Measurement Systems) (1990). "The STAR system, theory and
application".
Tebbett I.E., Lalani M. (1986). "Recent development in the reassessment, maintenance and
repair of steel offshore structures," 18th Offshore Technology Conference, paper 5113, pp
305-314.
Tebbett I.E., Lalani M. (1984). "A new approach to stress concentration factors for tubular
joint design," 16th Offshore Technology Conference, O T C 4825.
Tebbett I.E. (1982). "The reappraisal of steel jacket structures allowing for the composite
action of grouted piles," 14th annual Offshore Technology Conference, paper 4194.
Timoshenko S., Goodier J.N. (1970). "Theory of elasticity" McGraw-Hill, 3rd edition.
Tolloczko J.A., Lalani M. (1988). "The implication of new data on the fatigue life
assessment of tubular joints," 20th annual Offshore Technology Conference, paper 5662.
Tolloczko J.J.A., (1991). "Fatigue of tubular joints in offshore structures," in "Structures
subjected to repeated loading, stability and strength," edited by R.Narayanan and
T.M.Roberts, Elsevier Applied Science.
Toprac A.A., Johnston L.P. and Noel J. (1966). "Welded tubular connection: an investiga-
tion of stresses in T-joints," Welding Journal, January.
Ueda Y., Rashed S.M.H., Nakacho K.(1990). "An improved joint model and equations for
flexibility of tubular joints," Journal of Offshore Mechanics and Arctic Engineering, Vol
112, pp 157-168.
References
Ueda Y., Rashed S.M.H. (1990). "Modern method of analysis of offshore structures
structures," Proceedings of the First Pacific/Asia Offshore Symposium, Seoul, Korea.
Ueda Y., Rashed S.M.H., Ishihama T., Nakacho K. (1986). "Flexibility and Yield Strength
of Joints in Analysis of Tubular Offshore Structures," Proceedings of the 5th International
Offshore Mechanics and Arctic Engineering Symposium, Publ by A S M E , pp. 293-301.
UK Department of Energy (1990), Offshore installations: Guidance on design, construction
and certification, Fourth Edition, London: H M S O .
Underwater Engineering Group (1984). "Node flexibility and its effect on jacket
structures," U E G Publication, UR22.
Underwater Engineering Group (1985). "Design of tubular joints for offshore structures"
U E G Publication, UR33.
Visser W. (1974). "On the structural design of tubular joints," 6th annual Offshore
Technology Conference, paper 2117.
Kuang J.G., Potvin A.B., Leick RD. (1975). "Stress concentration in tubular joints," 7th
annual Offshore Technology Conference, paper 2205.
Wilson W.M., Moore H.F. (1917). 'Tests to determine the rigidity of riveted joints in steel
structures," University of Illinois, Engineering Experiment Station, Bulletin No. 104,
Urbana, U S A .
Wordsworth A.C., Smedley G.P. (1980). "Stress concentrations at unstiffened tubular
joints," Lloyd's register of shipping, London, European offshore Steel Research Seminar,
paper 31.
Wordsworth A.C. (1975). "The experimental determination of stresses at tubular joints,"
BSSM/RINA joint conference on measurement in the offshore industry, September.
Wordsworth A.C. (1981) "Stress concentration factors at K and KT tubular joints,"
Conference on Fatigue in Offshore Structural Steel, Inst of Civil Engineers, Westminster,
London, February.
Wordsworth A.C. (1981). "Stress concentration factors at K and KT tubular joints,"
Fatigue in offshore structural steel, paper 7, ICE, London.
Wordsworth A.W. (1987) "Aspects of stress concentration factors at tubular joints," Steels
in Marine Structures Conference
Young C.R., Jackson K.B. (1934). "The relative rigidity of welded and riveted
connections," Canadian Journal of Research, II, No.l and No. 2.
Zienkiewics O.C. (1977), "The Finite Element Method," McGraw-Hill, 3rd edition.
APPENDICES
Appendices: Appendix A
193
Appendix A
Stiffness matrix of beam with semi-rigid ends
The following sign conventions have been used for bending deformations and moments to
derive the stiffness matrix of a beam with semi-rigid ends.
c-
M
--)
V
vory
Bending moments and shear forces
Bending moments and deflections
Figure Al. Positive sign conventions for beam element
Considering an infinitesimal beam element under external loads and inertia forces, as shown
below:
V+dV) M+dM
dx
Figure A.2. Beam element loaded and being under equilibrium conditions
The equilibrium condition requires that:
3 V d
2
v
dx dt
2
which by using the following two equations for the beam members:
BM .. .
rT
d
2
y _
u
= V and EI^r = M
dx dx
2
the differential equation for dynamic deformation of a beam element can be obtained as:
(A.1)
(A.2)
Appendices: Appendix A
194
-~g
+
,.o
(A.3)
Assuming \(x,t) is separable as \(x,t) = y(jc)sin(_or) and no intermediate load is acting on the
beam, then:
i4 t4
(EIT~ma>
2
y)sincor = 0, therefore EI-m(0
2
y = 0
dx* dx*
y
(A.4)
Assuming A, = /(mo)
2
/__-)
l/4
, Equation (A.4) can be simplified to:
dx
4
(A.5)
The solution for the above equation is:
ten , retell
y(x) = CicosQudl) + C_sin(AJc/Z) + G e ^ ' + C<e
(A.6)
Considering the following boundary conditions:
8, = y(0) 5, = y(/)
e
i=
__m
+/(
o)
e..=+-7^+/(/)
(A.7)
in a matrix form will result:
A = <_>C
(A.8)
where A ^
fS/1
8*
C_
Aj)
VC
4
J
and
Appendices: Appendix A
195
<_> =
1
COSA
^
2
aX(y) COSA- (y)sinA
0
sinA
X
/
Assuming that the positive directions for the degrees of freedom of a beam element are as
shown below:
6j(3)
8i(1)
Figure A.3. Degrees of freedom for beam element
the end forces will be:
Vi = V(0) = _-/y
m,
(0)
Mi = -M(P) = -EIy"(0)
Vy = -V(/) = -EIy
m
(l)
Mj = M(l)=EIy"(t)
(A.9)
which can be shown in a matrix form as:
P = _i/QC
(A. 10)
r-\r^\
)Vj[
in that E and / are Modulus of Elasticity and moment of inertia, respectively. P =n
M
f'
V,
V
M
t
MjJ
C =
c_
C
3
* and Q =
0
* 3 ,
(y) sinA

-
* 3 ,
(y)cosA
0

A
3
*
(7)e
4
2
.3
<7>
-<7)V
4>
2
L -(y)
2
KA -(7)
2
sinA (y)V ( 7 ) ^ j
Appendices: Appendix A
196
Substituting from Equation (A.8) into (A. 10) yields:
P = F/QO
1
A
(A.11)
from which the stiffness matrix can be obtained as:
k = EIQQr
x
(A. 12)
Matrix k can be shown in a similar form to the stiffness matrix of a beam with rigid joints as
given by Equation (A. 13). This is the form which Kolousek (1973) used to show the
stiffness matrix of a beam using continuous mass model. The coefficients F, are called
frequency functions.
k =
frF_<X) JTFM ffF,(A) f F A )
^F
2
(X) fF
5
(X) |fF
6
(X) fF
7
(A)
FT FT FT FT
fF
3
(X) -JiFsiX) ~F
%
(X) -JT^CX)
FT FT FT FT
_ yrF
4
(X) TTF
7
(X) TTF
9
(A) F
1 0
( A )
(A. 13)
The Frequency Functions are very long and cumbersome since the matrices C and O have
lengthy trigonometric and hyperbolic expressions. Frequency functions of a beam with semi-
rigid joints modelled with a continuous mass property is given in Table A.l. The variables
used in Table A. 1 are defined as follwes:
X = l
' mO)
2
V
/4
El
(A. 14)
"'-Jul'
J
kji
(A.15)
Appendices: Appendix A
197
Table A.1. Frequency Functions of beam with semi-rigid joints (shear and axial deformation effects not
included)
Den = --^-^Oi+c^XcosA+sinAHAe
2
^
Ft = -A
3
[2^(XiAH<MnA-cosAXl-e
u
tt^ Den
F
2
= X
3
[(e
2X
-l)(l + CMXjA
2
) + A(cti + otj)(l + e
2
* + 2e
x
cosX) + 2e_iX(l -ctiOjX
2
]/Den
F
3
= -X
2
(- a,XcosX + e^OjAcosX - sinX + e^sinX + ctj XsinX + e^OjXsinX)/ Den
F
4
= -X
2
(l + e
2
*- - ctiX + e^ctiX - 2e
x
cosX + 2^-CXMDX)/ Den
F
5
= -X't-^cXjX^ar-X-cosXXl-e^cx^ Den
F
6
= X
2
(l + e
2
* - ctjX + e^ctjX - 2e
x
cosX + 2e
x
ajXsinX)/ Den
F
7
= X
2
(- ctiXcosX + e^ctiXcosX - sinX + e^sinX + ctiXsinX + e
u
aiXsinX)/ Den
F
g
= -X[(l - e
n
)(cosX - 2ctjXsinX)+ sinX(l + e
2
*)]/ Den
F
9
= -X(-l + e
2X
-2e_iX)/D_
F
w
= -X[(l e^XcosX - 2aiXsinX>i- sinX(l + e
2
*-))/ Pen ,
Appendices: Appendix B
198
Appendix B
Stiffness matrix of truss member with semi-rigid ends
The following sign conventions have been used to derive the stiffness matrix of an axially
loaded member with semi-rigid ends:
P< 4** =
r H
Figure B.l. Positive sign convention for axial deformation
The differential equation of deformation for a member axially loaded can be written by
applying the equilibrium equation on an element shown in Figure B.l. Therefore:
- = EA?^ (B.l)
dx dx
j
SA
|_*-
m
^*
=
_0 (B.2)
dx
2
dt
2
in which E is the Modulus of Elasticity, A is the sectional area and m is the mass per unit
length. h(x,t) describes the longitudinal displacement of the member. Assuming a solution
for Equation
(B.2) as h(x,t) = K(x)sin(G>0 and also \jr = * J^Tj-, it can be written that:
^
+
( ^ )
2
u = 0 (B.3)
dx
2
I
The particular solution of the above equation is:
u(x) = C5sin(\|tt//) + C_cos(\|tt//) (B-
4
)
Assuming the positive degrees of freedom for as shown in Figure B.2 below:
Appendices: Appendix B
199
n,(l) 11,(2)
Figure B.2. Degrees of freedom for a truss member
The boundary conditions for a truss member with semi-rigid joints are:
T H = ( 0 ) -
k'
^=1/(7)4
P(D
(B.5)
in which k[ or k- refers to the axial stiffness the member ends. Equation (B.5) can be
written in a matrix form as:
A_ = <D_C_ (B.6)
where A = | M , *
a
= . , _ , and C = {
C
J
L x|j J
a
L sm \f cos \|/ J'
a
I Q J
The relationships between end forces and displacements are:
p
{
= -P(0) = -EAu\0) Fj = P(0 = F_4u'(/)
(B.7)
which in a matrix form, they will be:
P_ = __AQ_C_
(B.8)
in which P
a
= j p J, and Q,, =
/
0
^cos \jr ^ i n \j/
Substituting from Equation (B.6) into (B.8) yields:
P =EAQ
a
*;
l
A
(B.9)
from which the stiffness matrix can be obtained as:
k_ = E4Q_<&
-I (B.10)
Appendices: Appendix B
200
The coefficients of matrix k_ are given in Table B.l.
Table B.l. Stiffness matrix of a truss with semi-rigid ends
Den = \ircos\|/ (ft +ft ) + sin\|f (1- pifty
2
), pi = EAIk'J, ft = EAlk)l
k
u
= (A/0v(
c
os\|/ - pjsin\|r)/Den k
u
= -(A//)v/Den
k
2
i = -(EAJl)\\r/Den k
n
= (2_A//)\|/(cos\{/ - pi_in\|r)/Den
Appendices: Appendix C
201
Appendix C
M o d e shapes 1, 2, and 3 for the tested T-joint specimens
The mode shapes of a series of models corresponding to the tested T-joints were calculated
using ALGOR Package. Therefore ID beam elements and stiffness method were used to
model the joints. Although the joints are assumed to be rigid in ALGOR, yet the overall
picture of the mode shapes are correct
mode 1 mode 2 mode 3
Figure C I . First three mode shapes ofTl for IPB mode.
mode 1
mode 2
mode 3
Figure C.2. First three mode shapes ofT2for IPB mode.
mode 1
mode 2
mode 3
Figure C.3. First three mode shapes ofT3for IPB mode.
Appendices: Appendix C
202
mode 1 mode 2 mode 3
Figure C.4. First three mode shapes ofT4for IPB mode.
mode 1 mode 2 mode 3
Figure C.5. First three mode shapes ofTSfor IPB mode.
mode 1
mode 2
mode 3
Figure C.6. First three mode shapes ofT6for IPB mode.
mode 1
mode 2
mode 3
Figure C.7. First three mode shapes ofT7for IPB mode.
Appendices: Appendix C
mode 1 mode 2
mode 3
203
Figure C.8. First three mode shapes of 78 for IPB mode.
mode 1 mode 2 mode 3
Figure C.9. First three mode shapes ofT9for IPB mode.
mode 1
mode 2 mode 3
Figure CIO. First three mode shapes of TIO for IPB mode.
mode 1
mode 2
mode 3
Figure C I 1. First three mode shapes of Til for IPB mode.
Appendices: Appendix D
204
Appendix D
List of sub-programs in A L G O R (used in the analysis of joints)
AEDIT program for editing the ASCII data files.
DECODS generates an ASCII file from a graphic file.
SD2 (Superdraw) C A D program.
SSAPO static analysis module.
SSAP1 dynamic analysis module which performs eigenvalue analysis.
SUBSTRUC program for gluing two structure together.
SVIEW program for viewing a model, its deformed shapes and stresses.
Appendices: Appendix E
205
Appendix E
Description of T B C 3 and G C S 8 finite elements (Ashwell and Gallagher, 1976)
TBC3
Description:
Flat shell element oriented in the global cartesian system.
Number of Nodes: 3 at the vertices
Nodal coordinates: x, y, z
Degrees of freedom:
X, Y, Z, RX, RY, RZ at each node.
Geometric properties required:
ti, ti, ts nodal thicknesses.
Material properties required:
(i) Isotropic- E Young's Modulus
v Poisson's ratio
Loading:
Nodal point loads or moments may be applied. Constant or linearly varying normal pressure
loads may be applied.
Stress Output-
Three membrane stresses Gxx, O > and Gxy and three moments per unit length M , M and
M,y are output at each node in the local axis system. The membrane stresses are constant
across the element
Figure E.l. TBC3 Element
GCS8
Description:
Arbitrary curved thin shell quadrilateral element with varying ftiOa^^c^y^^).
Number of Nodes: 3 (Vertices and midsides) *
Nodal coordinates: x, y, z
Degrees of freedom:
a) At corner nodes: X, Y, Z.
b) At midsides: X, Y, Z, RX, RV.
Geometric properties required:
tu th to A, ts, ft, ti, h (Thickness at the nodes).
Figure E.2. GCS8 Shell element
Material properties:
(i) Isotropic- E Young's Modulus
being determined by the node ordering on the element topology card.
Appendices: Appendix E 206
vPoisson's ratio
a-(Linear Expansion coefficient)
p-(Density of Material)
Loading:
Nodal point loads, pressure load and temperature loads may be applied.
Stress Output:
Membrane stresses (a, < % and Gxy) and Bending Moments (A/_x, Myy and Mxy) in local
directions only, are output at each node.
Appendices: Appendix F
207
Appendix F
Effect of chord length (Efthymiou 1985)
The local joint stiffness obtained from a F E model depends on the length of chord modelled.
If the chord length is small, the restrained chord-ends will tend to stiffen the joint and hence
lead to higher stiffness coefficients. In offshore platforms the chord length between nodes is
typically 20 diameters or more, so the proper stiffness coefficients are those relating to large
chord lengths.
A long chord (length diameter) loaded in an arbitrary manner at some sections OTigure
F.l) will become locally distorted. These distortions decay (in some manner) with distance
from the load. The decay length, x^, depends on the type of loading. Disturbances due to
axisymmetric loading (Figure F.lb) decay very fast (xj = V4D) but non-symmetrical load
(Figure F.lc), such as that caused by brace loading leads to inextensible bending of the
chord, which could take much longer to decay. In the finite element model it is important to
choose a correct chord length (at least equal to the decay length for the particular
geometry); otherwise the stiffness coefficients may be grossly overestimated.
Supported end
ry load
a)
b)
c)
Figure F.l. a) Circular cylinder loaded at end-section, b) axisymmetric load and c) non-axisymmetric bad
Appendices: Appendix F 208
The decay length for some relevant idealised situations can be estimated analytically using
thin shell theory as follows:
a) Axisymmetric loading on a circular cylinder
The displacements, stresses, bending moments etc. decay according to e
_Xx
, where x is the
distance from the load and
___^!_. (R.)
R
2
t
2
Decay is sufficient when XJC > 3, leading to x = 2.33 jRt. For typical Rlt ratios, say 20, the
decay length is x = 0.52/?.
b) Non-axisymmetric load (inextensional bending)
Displacements, stresses etc. decay according to <T
Rc(p)
, where x is the distance from the
load and p is given by the fourth order equation:
12(l-v
2
)/?
2
p4
+ m*(m
2
- 1 )
2
= 0 (F.2)
The parameter m indicates the harmonic order round the cylinder periphery. For m - 2,
Equation (F.2) reduces to:
"-^p'tU.0 (F.3)
from which the first real solution for p is obtained as:
n
_ V2
r
12 t
2
m (F 4)
P
" 2 S-v
2
R
2)
Considering decay to be sufficient when Re(p)x/_?>3 and using (F.4) leads to 233R>jR/t for
Rlt =20, x = 10/?, i.e. 20 times more than the case of axisymmetric load.
The above result relates to m = 2, which has a severe ovalising effect like that found in a
double X-joint with two pairs of equal and opposite forces. It can be considered an upper
bound on the decay length for all tubular joints. This suggests that in an F E model of, say, a
T/Y joint the total length of chord required will be a little less than 2x10* = 10D. It was
established in task I that modelling a length of chord equal to 6D is adequate.
Appendices: Appendix G 209
Appendix G
Parametric SCF Formulae for inplane bending of T-joints
SCF for chord SCF for branch
Kuang et al. (1975)
-0.04-.0.60-.0.86
0.702p-
OO4
Y
a60
T
Wordsworth andSmedley (1878)
0 75<v
o.6
T
o.8
(16
po.25_
07
r32)
DNV (1977)
2i ,_ 0.38 -.1.05
[1.65-l.l(j_-0.42)
2
]Y
038
T
UEG* (1985)
0 75y
o.6
x
o.8
(1>6
po.25 .
0
.7P
2
)Q
,m
Efthymiou andDurkin (1985)
I 45R
T
0.85y(l-0.68P)
Hellier (1990)
2.31 a
00033
Y
a326
x
95
exp[(0.00154Y
0.0323)/P
2
-0.0248]
Lloyd's Register (1991)
1.22T'
8
PY
(1
"
a68P)
-0.38 , 0_3 ~ 0.38
l^oip-
03
^
0
-
23
*
1 + 0.63 SCFc
[0.95-0.65(p-0.4)
2
]Y'
39
T
29
1 + 0.63 SCFc
I + O ^ P T
0
- ^
1 0 9 0
-
7 7
^
0 332 a
00053
Y
a37
x
a296
exp(- 0.00436/p
2
+ 1.4)
l+x
a2
YP(0.26-0.21p)
Q' =1 for Y < 20 and 480/7(40 - 0.833Y) for 20 < Y ^ 40.

You might also like