Vous êtes sur la page 1sur 13

Copyright 1999, Offshore Technology Conference

This paper was prepared for presentation at the 1999 Offshore Technology Conference held in
Houston, Texas, 36 May 1999.
This paper was selected for presentation by the OTC Program Committee following review of
information contained in an abstract submitted by the author(s). Contents of the paper, as
presented, have not been reviewed by the Offshore Technology Conference and are subject to
correction by the author(s). The material, as presented, does not necessarily reflect any
position of the Offshore Technology Conference or its officers. Electronic reproduction,
distribution, or storage of any part of this paper for commercial purposes without the written
consent of the Offshore Technology Conference is prohibited. Permission to reproduce in print
is restricted to an abstract of not more than 300 words; illustrations may not be copied. The
abstract must contain conspicuous acknowledgment of where and by whom the paper was
presented.
Abstract
There is a steadily increasing demand for the use of jack-up
units in deeper water and harsher environments. To be
confident of their use in these environments there is a need for
jack-up analysis techniques to reflect accurately the physical
processes occurring. The common modelling of spudcan
footings as pinned or as linear springs is an over-simplification
of foundation behaviour, and can therefore lead to over-
conservative design. This paper details the implementation of
a work hardening plasticity model for spudcan footings on
dense sand (named Model C) into a dynamic structural
analysis program (named JAKUP). The motivation for this is
to have a balanced approach to the analysis of jack-up units,
with the non-linearities in the structural, foundation and wave
loading models all taken into account. Wave loading is
considered using NewWave theory, and the importance of
random wave histories shown by constraining the
deterministic NewWave into a completely random surface
elevation. Examples are shown to emphasize the differences in
predicted extreme response according to the various footing
assumptions: pinned, linear springs and Model C. The role of
plasticity theory for modelling foundations will be
highlighted. Further uses of JAKUP in the understanding of
the extreme response statistics and reliability of jack-up
platforms are discussed.
This paper is directed to those interested in the long-term
reliability analysis of jack-up units.
Introduction
Jack-ups were originally designed for use in the relatively
shallow waters of parts of the Gulf of Mexico. Due to their
economic importance within the offshore industry, there has
been a steadily increasing demand for their use in deeper
water and harsher environments
1
. There is also a desire for a
longer lasting commitment of a jack-up at a single location,
especially in the role of a production unit
2
. To be confident of
their use in these environments, there has been a need for
changes in analysis techniques to make them more accurate,
avoiding unnecessary conservatisms which were once
commonplace.
More realistic modelling of jack-ups, based upon the relevant
physical processes has been developed in a number of areas,
the most significant being;
dynamic effects,
geometric non-linearities in structural modelling,
environmental wave loading,
models for foundation response.
There has been considerable diversity in this development,
with studies using varying levels of complexity for different
aspects of the analysis. This is especially true for soil/structure
interaction, with some studies using detailed structural models
or advanced wave mechanics, whilst still using the simplest of
foundation assumptions (i.e. pinned footings)
3,4,5
.
The purpose of this paper is to describe a more balanced
approach to the analysis of jack-up units. The components of a
dynamic structural analysis program, named JAKUP, which is
capable of considering the major non-linearities in jack-up
response, are described. The motivation for this development
is to achieve confidence in all of the components affecting the
response. The principal characteristics of the structural,
foundation and wave loading models are detailed, and
example calculations given. Lastly, a discussion of the
applications of these models is presented.
The Structural Model
For the modelling of jack-up response, structural non-
linearities must be considered if reasonable accuracy is to be
achieved. In particular, nonlinearities due to the presence of
axial loads in the legs from the deck's weight need to be taken
into account. P - and Euler effects are both accounted for in
JAKUP by using Orans
6
formulation of beam column theory
to specify the stiffness matrix. Additional modifications to
produce the additional end rotations on the beam due to the
OTC 10995
On the Modelling of Foundations for Jack-up Units on Sand
M.J. Cassidy and G.T. Houlsby, The University of Oxford
2 M.J. CASSIDY AND G.T. HOULSBY OTC 10995
presence of shear have also been implemented
7
. The stiffness
matrix is derived in incremental form, as the structural
response is path dependent. Both the mass and damping
matrix are time invariant, with the former derived using cubic
Hermitian polynomial shape functions and the latter by use of
Rayleigh damping. Structural damping coefficients are defined
for the lowest two modes, i.e. surge and sway in a jack-up.
Because of the need to model non-linearities, analysis in the
time domain using numerical step-by-step direct integration
techniques provides the most versatile method to analyse jack-
ups. Within JAKUP, the Newmark = 0.25, = 0.5 method is
used, since it is an unconditionally stable and highly accurate
algorithm. Further details of the structural formulations and
dynamic solution techniques can be found in Thompson
8
and
Williams et al.
9
.
The Foundation Model
There has been much interest in recent years in the level of
foundation fixity developed by spudcan footings. If some
foundation fixity is taken into account, critical member
stresses (usually at the leg/hull connection) and other critical
response values are reduced
10,11
. With higher levels of moment
restraint, the natural period of the jack-up is also reduced,
usually improving the dynamic characteristics of the rig. It is
still widely accepted practice, however, to assume pinned
footings (infinite horizontal and vertical and no moment
restraint) in the analysis of jack-ups
12,13
. This results in over-
conservative results. Another approach used in jack-up
analysis, and an improvement on use of pinned footings, is the
use of linear springs. Unfortunately, linear springs, while easy
to implement into structural analysis programs, do not account
for the complexities and non-linearities of spudcan behaviour,
and this simplistic method can render unrealistic results which
may be unconservative.
The use of a strain hardening plasticity theory is seen as the
best approach to model soil behaviour with a terminology
amenable to numerical analysis. This is because the response
of the foundation is expressed purely in terms of force
resultants. Though first used as a geotechnical solution to
another problem by Roscoe and Schofield
14
, it has been
recently used in the examination of jack-up performance (see
for instance Schotman
15
, Martin
7
and Thompson
8
). In this
section, features are described of an incremental work
hardening plasticity model, named Model C, that has been
developed to represent spudcan footings in the analysis of
jack-up units in sand. Model C is based on a series of
experimental tests performed at the University of Oxford by
Gottardi and Houlsby
16,17
.
Components of Model C. Model C has four major
components:
(1) An empirical expression for the yield surface in three
dimensional vertical, moment and horizontal loading
space (V, M/2R, H) (note: moment is normalised by the
radius of the spudcan, R) Once the yield surface is
established, any changes of load within this surface will
result only in elastic deformation. Plastic deformation can
result, however, when the load state touches the surface.
(2) An empirical strain-hardening expression to define the
variation of the size of the yield surface with the plastic
component of vertical displacement. Though the shape of
the yield surface is assumed constant, it expands with
vertical plastic penetration (as the footing is pushed
further into the soil).
(3) A model for elastic load-displacement behaviour within
the yield surface.
(4) A suitable flow rule to allow prediction of the footing
displacements during yield.
Details of these four components follow, with typical
parameter values given in Table 1. The sign convention used
throughout is that suggested by Butterfield et al.
18
, and is
shown in Fig. 1.
Yield Surface. The yield surface is defined by the best fit of
the experimental data, and has a functional form of
( )
2
0 0 0
2
0 0
2
0 0
2 2 2
0 , 2 ,
V m h
R M aH
V m
R M
V h
H
H R M V f

,
_

,
_



( )
( ) 2
2
0
1
2
0
2
2
2
1
1
2 1
2 1
1


+

,
_

,
_

1
1
]
1

V
V
V
V
..........(1)
This is a cigar shaped surface, as shown in Fig. 2, which
may be described as an eccentric ellipse in section on the
planes of constant V, and approximately parabolic on any
section including the V-axis.
0
V determines the size of the
yield surface and indicates the intersection of the yield surface
with the V-axis (H = 0 and 0 2 R M ). Furthermore,
0
V is
governed by the strain hardening law. The dimensions of the
yield surface in the horizontal and moment directions are
determined by
0
h and
0
m respectively.
Strain Hardening. Using Eqn 1 the size of the yield surface
is defined solely in terms of the pure vertical load capacity.
The variation of V
0
with plastic vertical displacement w
p
defines a hardening law for the yield surface, and is defined
from a combination of an empirical fit to experimental data for
flat circular footings on dense sand and a theoretical bearing
capacity approach for the conical section of the footings.
Experimental Evidence on Flat Plates. From the
experimental evidence of flat circular footings on dense sand,
the following formula defines the vertical capacity with
embedment:
2
0
0
2 1

,
_

,
_

,
_

pm
p
pm
p
m
pm
p
w
w
w
w
V
kw
kw
V ........................(2)
OTC 10995 ON THE MODELLING OF FOUNDATIONS FOR JACK-UP UNITS ON SAND 3
where k is the initial plastic stiffness,
m
V
0
the peak value of
0
V , and
pm
w the value of plastic vertical penetration at this
peak. Numerical values for k,
pm
w and
m
V
0
were derived
from the experiments, and the fit of this closed form solution
is shown in Fig. 3. A formula that models post-peak work
softening as well as pre-peak performance was essential,
however, Eqn 2 unrealistically implies 0
0
V as
p
w .
Therefore, it can only be used for a limited range of
penetrations. For jack-ups in dense sand, loading post-peak
would not be expected; however, for a complete foundation
model, Eqn 2 can be altered to:
2
0
0
2
0
1
1
2 1
1

,
_

,
_

,
_

,
_

,
_

,
_

pm
p
p pm
p
m
pm
m
pm
p
p
p
p
w
w
f w
w
V
kw
V
w
w
f
f
kw
V ..........(3)
where
p
f is a dimensionless constant that describes the
limiting magnitude of vertical load of
m
V
0
. (i.e.
m p
V f V
0 0

as
p
w ).
Assuming the same shape of the vertical strain hardening law,
k and
pm
w

can be defined for different sized footings by
introducing the dimensionless parameters f and
p
, where:
p pm
R w 2 ...............................................................(4)
v
RGk f k 2 ................................................................(5)
where G is a representative shear modulus and
1
K the vertical
elastic stiffness factor (see below). The values of f and
p

were determined from the experimental results as 0.1 and 0.03


respectively.
m
V
0
can be calculated for different sands from bearing
capacity theory as:
3
0
R N V
m


..............................................................(6)
where

N is the dimensionless bearing capacity for a circular


footing (see for example Bolton and Lau
19
).
Conical Spudcans Theory. The theoretical case of a
conical footing, as depicted in Fig. 4(c), can be split into two
cases. Firstly, a conical section of varying radius (r) and
penetration ( ( ) 2 tan < R w
p
) and secondly, after full
penetration of the conical section ( ( ) 2 tan R w
p
). Before
full penetration of the conical section, theoretically, the
vertical load is proportional to the cube of the current radius of
the penetrated section. This is represented by segment AB
in Fig. 4(d), and written as
( ) ( )
3 3
0
2 tan
p
Theory
w N r N V ..................(7)
With the conical section of the footing fully penetrated, BC
on Fig. 4(c), the footing is assumed to penetrate at a constant
load (neglecting the geometric effects of embedment).
Conical Spudcans - Combined Method. By combining these
theoretical and experimental ideas, a more realistic strain
hardening law can be derived for spudcans in dense sand. Figs
4(e) and 4(f) outline this new model. Using an N

value
appropriate for the geometry and roughness of the spudcans
and friction angle of the sand, the experimental flat footing
curve shape can be normalised by the theoretical maximum
m
V
0
, as shown in Fig. 4(e) and written as:
2
0
2
0
0
0
2 1
1
2
2 1
2 1

,
_

,
_

,
_

,
_

,
_

p
p
p p
p
m
p
p
p
p
p
m
p
m
Exp
R
w
f R
w
V
kw
R
w
f
f
V
kw
V
V
....(8)
where
m
V
0
is given by Eqn 6. It should be noted that the
plastic penetration has been normalised by the vertical
penetration at peak load (Eqn 4).
Until full penetration of the conical footing the appropriate
value of
p p
R w 2 is a constant and equal to
( ) 2 tan 2 1
p
. From this value a constant factor
m
Exp
fp
V V s
0 0
can be determined from Eqn 8., as
indicated in Fig. 4(e). Before full penetration of the cone
( ( ) 2 tan < R w
p
), the load penetration curve from the
theoretical model is used, but scaled by s
fp
, thereby,
consistently combining the two methods. This is shown as
section A B in Fig. 4(f) and can be written as
( )
3 3
0
) 2 tan( V
p fp fp
w N s r N s .....................(9)
After full penetration of the conical section, Eqn 9 can still be
applied, however, s
fp
will now vary according to Eqn 8, thus
following the original Model C experimental shape. If
( ) 2 tan R w
pm
, the shape of the experimental hardening
law will result in a peak load at ( ) 2 tan R w
p
. This is
illustrated as section B C of Fig. 4(f). Conversely, if
( ) 2 tan < R w
pm
, the entire response after embedment is
predicted from the post-peak section of the experimental
curve. The maximum load occurs just as full embedment is
reached, as shown in section B D of Fig. 4(f). For realistic
4 M.J. CASSIDY AND G.T. HOULSBY OTC 10995
values of the parameters it appears that the latter case is more
usual.
Elastic Behaviour. Elastic response of the soil needs to be
defined for any increments within the yield surface. Finite
element work has shown that cross coupling exists between
the horizontal and rotational footing displacements
18
, with a
linear elastic incremental force-displacement relationship of
the form

,
_

1
1
1
]
1

,
_

e
e
e
du
Rd
dw
h
k
c
k
c
k
m
k
v
k
GR
dH
R dM
dV
2
0
0
0 0
2 2 ...................(10)
where G is a representative shear modulus. Eqn 10 is
implemented in Model C with the elastic constant values
evaluated by Bell
20
for a Poissons ratio of 0.2. The shear
modulus is estimated as:
a a
p
R
g
p
G

2
..........................................................(11)
where
a
p is atmospheric pressure, the submerged unit
weight of sand and g a non-dimensional shear modulus factor.
Values of
v
k ,
m
k ,
h
k ,
c
k and g are given in Table 1.
Plastic Potential. The experimental evidence did not support
the application of associated flow in the vertical/horizontal
and vertical/moment planes and a plastic potential function g
was defined to relate the ratio between the plastic
displacements as follows:

,
_

,
_

H
g
2R M
g
V
g
du
2Rd
dw
p
p
p
..............................................(12)
Plastic displacements occur when the force point is located on
the yield surface and in the direction normal to the plastic
potential, which for Model C is defined as
0 0 0 0
2
0 0
2
0 0
2
2
2
0
V m
R / M
V h
H
a
V m
R / M
V h
H
g
m h m h

,
_

,
_




( )
( )
( ) ( )
4 3
4 3
4 3
2
0
2
0
2
4 3
4 3
1




,
_

,
_

,
_

+
V
V
V
V
......(13)
where
0
V is a dummy variable which is adjusted so that the
potential passes through the current load point.
h
and
m

are non-dimensional association factors relating the shape of


the plastic potential to that of the yield surface. They are
written as hyperbolic functions in terms of plastic
displacement histories as:
( )
( )
p p
p p h
h
w u k
w u k
+
+


............................................(14)
( )
( )
p p
p p m
m
w R k
w R k
+
+


2
2
........................................(15)
where k determines the rate of change of the association
factors. For no previous radial displacements,
h
and
m
equate to 1 and associated flow is assumed. The rates at
which
h
and
m
vary in Model C are depicted in Fig. 5.
The use of Plasticity Theory for Spudcan Footings. The
major advantages of Model C in the analysis of jack-up
response are:
it is formulated in a form amenable to numerical analysis,
allowing it to be implemented into structural analysis
programs.
it accounts for the non-linearities of combined loading on
sand in a consistent manner.
a direct indication of yielding is given. Furthermore,
movement of the spudcan footings can be evaluated, with
differentiation between upwave and downwave leg
behaviour possible. Sliding of spudcan footings,
therefore, can be evaluated directly.
a realistic interpretation of spudcan fixity allows for more
accurate dynamic analysis. Model C will give
significantly different dynamic responses to pinned and
fixed footing assumptions.
The Wave Loading Model
For structures such as jack-ups, which respond dynamically, it
is important to simulate the random, spectral and non-linear
properties of ocean loading. The extreme dynamic response
depends not only on the load being currently applied, but is
also affected by its load history. Therefore, the most accurate
methods of estimating extreme response are based on random
time domain simulation of the ocean surface and the
corresponding kinematics. However, acquiring confidence in
random time domain simulation results is computationally
time consuming, and so for convenience deterministic wave
theories, for example Airy and Stokes V, are still widely used
for calculating wave loading on jack-ups. However,
comparisons of deterministic regular wave and validated
random wave theories show that the regular wave theories
tend to overestimate wave kinematics and the fluid load
21
.
Moreover, regular wave theories assume all wave energy is
concentrated in one frequency component rather than the
broad spectrum of the ocean environment and hence give an
unrepresentative dynamic response.
OTC 10995 ON THE MODELLING OF FOUNDATIONS FOR JACK-UP UNITS ON SAND 5
NewWave Theory. NewWave theory, a deterministic method
described by Tromans et al.
21
, accounts for the spectral
composition of the sea, and can be used as an alternative to
both regular waves and full random time domain simulations
of lengthy time periods. The method involves the
superposition of directional, linear wavelets with an extreme
crest associated with the superposition of all wavelet crests at
a specific point in space and time. The surface elevation
around this extreme wave event is modelled by the statistical
average of the elevations associated with all occurrences of
this extreme event. This shape is also given by the
autocorrelation function of the Gaussian process defining the
sea-state.
The deterministic formulation of NewWave allows it to be
conveniently and efficiently implemented into structural
analysis programs, such as JAKUP. As it is based on linear
theory, the water wave particles can be easily obtained once
the water surface has been established. Care must be taken,
however, in describing their values in a crest, as linear wave
theory has no validity at or above the mean water surface. A
number of stretching approaches are commonly used, such as
Delta stretching
22
or Wheeler stretching
23
. Due to the
substantial separation of jack-up legs the ability to evaluate
accurately the time varying surface elevation and kinematics
at two spatial positions is important. An approximate solution
method for the dispersion relation for plane waves in a
constant water depth, as outlined by Newman
24
, has been
implemented in JAKUP.
Constrained NewWave. NewWave theory also can be used
to produce a time series of a random surface elevation, by
constraining a NewWave with a pre-determined height within
a completely random background. This is performed in a
mathematically rigorous manner such that the constrained
sequence is statistically indistinguishable from the original
random sequence. A procedure outlining this can be found in
Taylor et al.
25
. Constrained NewWave allows for the easy and
efficient evaluation of extreme response statistics; achievable
without the need to simulate many hours of real time random
seas most of which is of no interest.
Example Analyses
An example jack-up analysis for the structure shown in Fig. 6
will be outlined. The mean water depth is 90m, and the rig
size typical of one used in harsh North Sea conditions. Fig. 6
represents an equivalent beam model, with the corresponding
stiffnesses and masses of the beams shown. The hull is also
represented by a beam element with a rigid leg/hull
connection. Though non-linearities in the leg/hull jack houses
are recognised as significant, they were not included in this
analysis. The hydrodynamic modelling of the leg is performed
by idealising the detailed lattice leg as an equivalent vertical
tubular section according to the SNAME
26
procedures, with
the equivalent hydrodynamic coefficients shown.
NewWave Example. Fig. 7 shows the NewWave wave
profile evaluated in JAKUP for the upwave and downwave
legs of the example structure. The sea-state can be described
by the Pierson Moskowitz wave energy spectrum, with a
significant wave height (H
s
) of 12m and a mean zero crossing
period (T
z
) of 10s. The peak NewWave crest has been focused
on the upwave leg at the reference time (t = 0s).
It is essential for the relative motion between the structure and
the water to be considered in the analysis of jack-ups, as a
significantly larger response is predicted if relative velocity
effects are ignored (see for instance Chen et al.
5
and Manuel
and Cornell
27
). Therefore, the extended Morison equation is
used to relate the horizontal kinematics to the hydrodynamic
loads on the jack-up legs.
For the example jack-up, Fig. 8 describes the wave forces in
the time domain calculated on each leg for the NewWave
shown in Fig. 7. The forces are the sum of the loads applied at
each node on the leg. The rig is assumed to have two legs to
windward, so for the upwave legs the figures shown are the
total for two legs. For this example, the environmental force is
purely wave loading, with no wind or current included.
The corresponding horizontal deck displacements due to this
NewWave are shown in Fig. 9 for three foundation cases:
pinned, Model C and linear springs. Pinned footings represent
infinite horizontal and vertical stiffness, but no rotational
stiffness. Model C is the strain hardening plasticity model for
sand, whilst linear springs uses finite stiffness values as in the
elastic region of the Model C case (Eqn 10). Though only
horizontal deck displacements have been shown, any measure
of structural response can be determined by JAKUP. After the
NewWave passes, the structure can be seen to be vibrating in
its natural mode. With increased rotational fixity the natural
periods decrease, with approximate values of 9, 5, and 5
seconds for the pinned, Model C and linear spring footings
respectively. As expected the pinned footings give the largest
horizontal deck displacement over the time period and the
fixed case the lowest. The pinned case can be seen, for this
example, to be rather conservative compared to the Model C
footings with a peak displacement close to a factor of four
greater (as would be expected from a quasi-static linear model
of a simplified jack-up).
In the example shown in Fig. 9, the load combinations on the
Model C footings were contained entirely within the yield
surface, thus giving a response identical to the linear spring
case. By increasing the NewWave crest amplitude to = 15m
or = 18m, as shown in Fig. 10, the increased loading caused
plastic displacements in the Model C footings, shifting the
entire foundations and leaving a permanent offset in the
displacement of the deck. This yielding of the sand footings
occurred during the peak of the NewWave. This direct
indication of yielding is a major benefit in using elasto-plastic
formulations for the spud-can footings. The natural period
after this event may also be modified by the plastic behaviour.
6 M.J. CASSIDY AND G.T. HOULSBY OTC 10995
Constrained NewWave Example. Fig. 11 illustrates the
surface elevation of a NewWave with a crest elevation of 15m
embedded in a random sea characterised by H
s
= 12m and T
z
=
10s. The wave has been constrained, such that at about 59.34s,
its peak collides with the upwave leg of the jack-up; the
surface elevation for the downwave is also displayed. The
corresponding deck displacements with time are shown in Fig.
12 for the pinned, Model C and linear spring foundation
assumptions. For this example, the peak displacements have
been increased compared to just the equivalent NewWave
(Fig. 10). This is due to the random background and the
structural memory it causes; displaying that for dynamically
sensitive structures, such as jack-ups, the response is not only
conditional on the present applied load, but also on the load
history. As was the case for a jack-up loaded exclusively by a
NewWave, the assumption of pinned footings is clearly
illustrated in Fig. 12 as overly conservative. The linear springs
can be seen to yield lower displacements than the Model C
footing due to the greater stiffness exhibited. In addition,
Model C indicates a permanent horizontal displacement of the
jack-up due to the passage of this extreme wave.
Applications
Though only one Constrained NewWave example was shown
here, one of the main benefits of the constraining technique is
that the probability distribution of the extreme response can be
estimated without the need to simulate many hours of real
time. For a storm associated with one sea-state, shorter time
periods can be used with a logical combination of crest
elevations to simulate responses for the expected wave sizes
within that sea-state. Convolution with the probability of
occurrences of crest elevations allows for the compilation of
response statistics. With knowledge of long-term sea
conditions, long term extreme exceedence probabilities for
response design properties of interest in the reliability of jack-
ups (for example lower leg-guide moments) can be evaluated.
The capacity to evaluate the difference between modelling
assumptions, not just for singular loading examples, but for
long-term conditions, is now possible. For example, the
difference in long term probability of exceedence values
between a pinned and an elasto-plastic model (Model C) must
be of interest to the designers and operators of jack-up units.
JAKUP, therefore, has immediate applicability in the
understanding of extreme response statistics and the reliability
of jack-up platforms. In any reliability analysis, the results can
only be judged on the accuracy of the individual components
used in the analysis. This is especially true with highly
interactive and non-linear processes, as seen in jack-ups. With
inappropriate and highly conservative assumptions, such as
pinned footings, not only are the reliability results inaccurate,
but the level of uncertainty in them can be unacceptably high.
A complete sensitivity study of the important random
variables involved in the analysis of jack-ups, when using the
Model C foundation models, is being conducted at the
University of Oxford. Not only will the importance of
parameters within an elasto-plastic model be evaluated, but
also their relative importance compared to other random
variables commonly used in reliability analysis.
Conclusions
An elasto-plastic model, entitled Model C, appropriate for the
modelling of foundations for jack-up units on sand has been
detailed. Furthermore, the advantages of its application as the
foundation model for a plane frame analysis program
(JAKUP) have been demonstrated. As jack-up units are
dynamically responding structures, the importance of random
time history analysis has been shown through use of the
NewWave and constrained NewWave wave loading methods.
The inclusion of the spectral properties of the ocean make
these deterministic methods advantageous over the widely
used regular wave formulations. Further applications,
especially relevant to the reliability analysis of jack-up units,
have been highlighted.
Nomenclature
a = eccentricity of the yield surface
d
C = drag coefficient
f = yield function
f = Initial plastic stiffness factor
g = plastic potential function
g = non-dimensional shear modulus factor
G = shear modulus
0
h = dimension of yield surface (horizontal)
H = horizontal load
s
H = significant wave height
k = initial plastic stiffness
k = rate of change of association factor
v
k = elastic stiffness factor (vertical)
m
k = elastic stiffness factor (moment)
h
k = elastic stiffness factor (horizontal)
c
k = elastic stiffness factor (combined)
0
m = dimension of yield surface (moment)
M = moment load

N = Bearing capacity factor (peak)


a
p = atmospheric pressure
R = radius of circular footing
z
T = mean zero crossing period
u = horizontal footing displacement
p
u = plastic horizontal footing displacement
V = vertical load
0
V = maximum vertical load capacity when H = 0 and M = 0
m
V
0
= peak value of
0
V in strain hardening law
0
V = maximum vertical load for plastic potential
w = vertical footing displacement
OTC 10995 ON THE MODELLING OF FOUNDATIONS FOR JACK-UP UNITS ON SAND 7
p
w = plastic vertical footing displacement
pm
w = value of plastic vertical displacement
= crest elevation
h
= horizontal association factor
m
= moment association factor
0
= value of association factor with no displacements

= association factor as displacements infinity


= cone apex angle
1
,
2
= curvature exponents in f (yield surface)
3
,
4
= curvature exponents in g (plastic potential)
= stability parameter in Newmark method
= dissipation parameter in Newmark method
p
= Dimensionless plastic penetration at peak
= submerged unit weight of soil
= rotational footing displacement
p
= plastic rotational footing displacement
= Poisson's ratio
superscripts
e = elastic
p = plastic
Acknowledgements
Support from the Rhodes Trust and the Wroth Fund for the
first author is gratefully acknowledged.
References
1. Carlsen, C. A., Kjey, H. and Eriksson, K.: "Structural
behaviour of harsh environment jack-ups." The Jackup Drilling
Platform Design and Operation, Collins, London, pp. 90-136,
(1986).
2. Bennett, W.T. and Sharples, B.P.M.: "Jack-up legs to stand on?"
Mobile Offshore Structures, Elsevier, London, pp. 1-32, (1987).
3. Daghigh, M., Hengst, S., Vrouwenvelder and A, Boonstra, H.:
"System reliability analysis of jack-up structures under extreme
environmental conditions." Proc. of BOSS97 Behaviour of
Offshore Structures, Vol. 3, pp. 127-144, (1997).
4. Mommas, C.J. and Grndlehner G.J.: "Application of a
dedicated stochastic non-linear dynamic time domain analysis
program in design and assessment of jack-ups." Recent
Developments in Jack-Up Platforms, Elsevier, London, pp.152-
175, (1992).
5. Chen, Y.N., Chen, Y.K. and Cusack, J.P.: "Extreme dynamic
response and fatigue damage assessment for self-elevating
drilling units in deep water." SNAME Transactions, Vol. 98, pp.
143-168, (1990).
6. Oran, C.: "Tangent stiffness in plane frames." J. Struct. Engng
Div. ASCE. Vol. 99, No. ST6, (1973[a]).
7. Martin, C.M.: Physical and numerical modelling of offshore
foundations under combined loads, DPhil Thesis, Oxford
University, (1994).
8. Thompson, R.S.G.: "Development of non-linear numerical
models appropriate for the analysis of jack-up units." D.Phil
Thesis, University of Oxford, (1996).
9. Williams, M.S., Thompson, R.S.G. and Houlsby, G.T.: "Non-
linear dynamic analysis of offshore jack-up units." Report No.
2071/95, Oxford University Engineering Laboratory, (1997).
10. Chiba, S., Onuki, T. and Sao, K.: "Static and dynamic
measurements of bottom fixity." The Jackup Drilling Platform
Design and Operation, Collins, London, pp. 307-327, (1986).
11. Norris, V.A. and Aldridge, T.R.: "Recent analytical advances in
the study of the influence of spud can fixity on jack-up unit
operations." Recent Developments in Jack-Up Platforms,
Elsevier, London, pp. 424-450, (1992).
12. Reardon, M.: "Review of the geotechnical aspects of jack-up
unit operations." Ground Engineering, Vol. 19, No. 7, pp 21-26,
(1986).
13. Frieze, P.A., Bucknell, J., Birckenshaw, M., Smith, D. and
Dixon, A. T.: "Fixed and jack-up platforms: basis for reliability
assessment." Proc. 5
th
In. conf. on Jack-Up Platform Design,
Construction and Operation, City University, London, (1995).
14. Roscoe, K.H. and Schofield, A.N.: The stability of short pier
foundations in sand, British Welding Journal, August, pp. 343-
354, (1956).
15. Schotman, G.J.M.: The effects of displacements on the stability
of jackup spudcan foundations, Proc. 21st Offshore Technology
Conf., Houston, OTC 6026, (1989).
16. Gottardi, G. and Houlsby, G.T.: Model tests of circular footings
on sand subjected to combined loads, OUEL Report No.
2071/95, Department of Engineering Science, Oxford University,
(1995).
17. Gottardi, G., Houlsby, G.T. and Butterfield, R.: The
Plastic Response of Circular Footings on Sand under
General Planar Loading, Gotechnique, in press (1999)
18. Butterfield, R., Houlsby, G.T. and Gottardi, G.: "Standardized
sign conventions and notation for generally loaded foundations."
Geotchnique, Vol. 47, No. 5, pp1051-1054, (1997).
19. Bolton, M.D. and Lau, C.K.: Vertical bearing capacity factors
for circular and strip footings on Mohr-Coulomb soil, Can.
Geotech. Jour., Vol. 30, pp. 1024-1033, (1993).
20. Bell, R.W.: The analysis of offshore foundations subjected to
combined loading, MSc Thesis, Oxford University, (1991).
21. Tromans, P.S., Anaturk, A.R. and Hagemeijer, P.: A new
model for the kinematics of large ocean waves -Applications as
a Design Wave-, ISOPE-91 Conference, Edinburgh, (1991).
22. Rodenbusch, G. and Forristall, G.Z.: "An empirical model for
random wave kinematics near the free surface." Proc. 18th
Offshore Technology Conference, Houston, pp. 137-146, OTC
5097, (1986).
23. Wheeler, J.D.: "Method for calculating forces produced by
irregular waves." J. Petroleum Technology, March, pp. 359-367,
(1970).
24. Newman, J.N.: "Numerical solutions of the water-wave
dispersion relation." Applied Ocean Research, Vol.12 No. 1,
Elsevier Science Limited, (1990).
25. Taylor, P.H., Jonathon, P. and Harland, L.A.: Time domain
simulation of jack-up dynamics with the extremes of a gaussian
process, 14th Conference on Offshore Mechanics and Arctic
Structures (OMAE), Vol. 1-A, pp 313-319, Copenhagen,
Denmark, (1995).
26. SNAME: Guidelines for the site specific assessment of mobile
jack-up units. Society of Naval Architects and Marine
Engineers, New Jersey, (1994).
27. Manuel, L. and Cornell, C.A.: "Sensitivity of the dynamic
response of a jack-up to support modelling and morison force
modelling assumptions." Proc. Conference on Offshore
Mechanics and Arctic Structures (OMAE), Vol. 2, pp.243-250,
(1993)
8 M.J. CASSIDY AND G.T. HOULSBY OTC 10995
Table 1 - Parameters for Model C
Constant Dimension Explanation Constraints Typical
value
Notes
R L Footing radius (various)

F/L
3
Unit weight of soil 10kN/m
3
Saturated sand
g
- Shear modulus factor 4000
v
k - Elastic stiffness factor (vertical) 2.65
m
k - Elastic stiffness factor (moment) 0.46
h
k - Elastic stiffness factor (horizontal) 2.3
c
k
- Elastic stiffness factor (horizontal/moment
coupling)
-0.14
0
h - Dimension of yield surface (horizontal) 0.116 Maximum value of
0
/V H on
0 M
0
m - Dimension of yield surface (moment) 0.086 Maximum value of
0
2 / RV M on
0 H
a - Eccentricity of yield surface 0 . 1 0 . 1 < < a -0.2
1
- Curvature factor for yield surface (low stress) 0 . 1
1
0.82 1
2 1
gives parabolic section
2
- Curvature factor for yield surface (high stress) 0 . 1
2
0.99 1
2 1
gives parabolic section
3
Curvature factor for plastic potential (low
stress)
0 . 1
3
0.55
4
Curvature factor for plastic potential (high
stress)
0 . 1
4
0.65
h
Association factor (horizontal) 1.0-2.5 variation according to Eqn 15
and 5 . 2
h
m
Association factor (moment) 1.0-2.15 variation according to Eqn 16
and 15 2.
m

k Rate of change in association factors 0.125


f - Initial plastic stiffness factor 0.1

N - Bearing capacity factor (peak) 150-300


p
- Dimensionless plastic penetration at peak 0.03
u
w

Reference position
Current position
M
V
H
2R
Figure 1 - Sign convention (after Butterfield et al
18
)
OTC 10995 ON THE MODELLING OF FOUNDATIONS FOR JACK-UP UNITS ON SAND 9
V
H
M/2R
Figure 2 - Shape of yield surface in Model C
0
500
1000
1500
2000
2500
-1 0 1 2 3 4 5 6 7 8 9 10
w
p
(mm)
V

(
N
)
Experiments
Theory
Figure 3 - Strain hardening law of Model C
10 M.J. CASSIDY AND G.T. HOULSBY OTC 10995
R
(a) Flat plates
R
r

Displaced soil
neglected
R
tan(/2)
(c) Conical footings
V
oExp
V
oTheory
w
p
2R
p
1.0
1.0
s
fp
1
2
p
tan(/2)
(e) Normalised curves
w
p
V
o
V
omExp
= N

R
3
(b) Experimental evidence
w
p
V
o
V
omTheory
= N

R
3
Cubic
A
B C
R
tan(/2)
(d) Conical Footing: theory
w
p
V
o
Theory
A
B
C
R
tan(/2)
D
(f) Conical footing: combined model
Figure 4 - Adaptation of Model C for the conical shape of spudcans
OTC 10995 ON THE MODELLING OF FOUNDATIONS FOR JACK-UP UNITS ON SAND 11
0
0.5
1
1.5
2
2.5
3
0 1 2 3 4
u
p
/ w
p

or
2 R

/ w
p

h

o
r

m
horizontal
moment
Figure 5 - Rates of variation of
h


and
m
in Model C
M
e
a
n

w
a
t
e
r

d
e
p
t
h

9
0
m
80m
35.2m
Upwave Downwave
Two legs
Single leg
51.96m
Figure 6 - General layout of idealised jack-up used in the analyses
Values:
For a single leg:
E = 200GPa
I = 15m
4
A = 0.6m
2
M = 1.93x10
6
kg
A
s
= 0.04m
2
G = 80GPa
D
E
= 8.44m
A
h
= 3.94m
2
C
d
= 1.1
C
m
= 2.0
For hull:
I = 150m
4
A
s
= 0.2m
2
M = 16.1x10
6
kg
For spudcans:
R = 10 m
12 M.J. CASSIDY AND G.T. HOULSBY OTC 10995
-0.8
-0.6
-0.4
-0.2
0
0.2
0.4
0.6
0.8
1
-60 -40 -20 0 20 40 60
time (s)
deck displacement
(
m
)
pinned
Model C and linear springs
-5
-2.5
0
2.5
5
7.5
10
-60 -40 -20 0 20 40 60
time (s)
f
o
r
c
e

o
n

j
a
c
k
-
u
p

l
e
g
s

(
M
N
)
force on upwave legs
force on downwave leg
force in total
-12
-8
-4
0
4
8
12
-60 -40 -20 0 20 40 60
time (s)
s
u
r
f
a
c
e

e
l
e
v
a
t
i
o
n

(
m
)
upwave legs
downwave leg
Hs
= 12 m
T z
= 10 s
= 12 m
x = 0m x = 51.96m
upwave
leg
downwave
leg
Figure 7 - NewWave surface elevation at the upwave and downwave legs
Figure 8 - Horizontal force on the example jack-up's legs due to NewWave loading
Figure 9 - Horizontal deck displacements due to NewWave loading
OTC 10995 ON THE MODELLING OF FOUNDATIONS FOR JACK-UP UNITS ON SAND 13

-0.8
-0.6
-0.4
-0.2
0
0.2
0.4
0.6
0.8
1
-60 -40 -20 0 20 40 60
time (s)
d
e
c
k

d
i
s
p
l
a
c
e
m
e
n
t

(
m
)
12 m
15 m
18 m
Figure 10 - Horizontal deck displacements due to increasing amplitude NewWaves
-15
0
15
0 20 40 60 80 100 120
time (s)
s
u
r
f
a
c
e

e
l
e
v
a
t
i
o
n

(
m
)
upwave legs
downwave leg
Hs
= 12 m
T
z
= 10 s
= 15 m
x = 0m x = 51.96m
upwave
legs
downwave
leg
Figure 11 - Surface elevations at the upwave and downwave legs for a constrained NewWave
-1.2
-0.8
-0.4
0
0.4
0.8
1.2
1.6
0 20 40 60 80 100 120
time (s)
d
e
c
k

d
i
s
p
l
a
c
e
m
e
n
t

(
m
)
pinned
Model C
linear springs
Figure 12 - Horizontal deck displacements due to the constrained NewWave

Vous aimerez peut-être aussi