Sunteți pe pagina 1din 30

Marine Structures 9 (1996) 71-100

1995 Elsevier Science Limited


Printed in Great Britain. All rights reserved
0951-8339/96/$9.50
ELSEVIER 0951-11339(95)00005-4

Nonlinear Dynamic Behaviour of Jack-up Platforms

N. Spidsoe & D. Karunakaran


SINTEF Structures and Concrete,Trondheim,Norway
(Received5 January 1995)

ABSTRACT

Jack-up platforms may, under extreme wave loading conditions, exhibit


significant nonlinear behaviour. This must be accounted for in the design of
such platforms, in order to ensure satisfactory structural safety.
In this paper an overview of the different sources of nonlinear platform
behaviour is presented. Furthermore, it is outlined how those nonlinear
effects may be modelled and handled in numerical simulation studies of
structural response, with a special focus on the estimation of extreme
response and dynamic amplification factors. Finally, it is discussed how
nonlinear dynamic response may be accounted for in the design of jack-up
platforms.
The discussions and outlines are illustrated by examples from numerical
simulation studies of jack-up behaviour.

Key words: jack-up, nonlinear dynamic response, extreme response.

INTRODUCTION

Today it is widely accepted within the offshore industry that the jack-up
platform concept generally is sensitive to a number of nonlinear dynamic
response effects. Consequently, there is a general understanding between
structural engineers that this must be accounted for in the structural
design analysis, both with respect to which type of analysis should be used
and which mathematical models should be involved in these analyses.
A far step towards an agreed industrial practice has been taken during

71
72 N. Spidsoe, D. Karunakaran

the joint industry project 'Jack-up site assessment procedures - - estab-


lishment of an international recommended practice'. Though there is no
final conclusion to this subject yet, it appears that time domain stochastic
dynamic response analysis is an analysis method which will be a central
element of the procedure. However, it seems not to be clear so far how this
method should be integrated in a practical design procedure.
Based on a wide experience from research studies and design analysis,
this is also the firm opinion of the authors. The objective of this paper is,
consequently, to contribute to the process towards an industrial practice
by bringing forward those experiences and results which may be relevant
for the ongoing discussion.
The paper is divided into three main parts. The first gives an updated
overview of the different sources of nonlinear platform behaviour and
how the underlying mechanism may be modelled for numerical analysis.
In the second part the important issues of using time domain stochastic
dynamic response analysis are discussed, as concerns the establishment of
simulation schemes and interpretation of simulation results. Finally, the
third part gives some suggestions on the practical use of the method in the
design process. In all three parts examples are given from earlier studies.
The platforms referred in these examples are briefly described in Appen-
dices A and B.

SOURCES OF N O N L I N E A R P L A T F O R M BEHAVIOUR

Though they may hardly be said to be independent it is convenient to


divide the sources of nonlinear platform behaviour into the following
main classes:
- - hydrodynamic loading effects,
- - structural properties,
- - dynamic effects.

Hydrodynamic loading effects

The hydrodynamic loading is often classified as inline forces, i.e. forces in


the wave and current direction; and transverse forces, i.e. forces normal to
the wave and current direction. Since the wave loaded part of jack-up
platforms consists of frameworks of slender elements, both these types of
forces are highly nonlinear with respect to the wave height and current
speed.
The inline forces are by far the most important. For all practical purposes
Nonlinear dynamic behaviour ofjack-up platforms 73

they are modelled by the so called extended Morison equation. ~' ~8 In


extreme wave situations the quadratic drag term of this load model is the
dominating term, with severe consequences both for the quasistatic and the
dynamic response. The quasistatic effects are mainly to the higher order
statistical properties of the response, and are well known from the works by
Pierson and Holmes 19 and Borgman. 2 These show that, as the relative
importance of the drag term increases in a pure wave load situ.ation, the
kurtosis of the force process increases compared to the wave process, i.e.
the extremes in the force process are higher than in the inducing process. If
current is included this implies that the skewness of the load process also
increases compared to the inducing process, leading to a further increase of
the extreme forces. In the response analysis the distributed forces are inte-
grated to the instantaneous wave surface elevation, assuming full correla-
tion of the forces along the platform legs. The integration accumulates the
properties of the distributed forces, but it also introduces a significant
additional skewness to the quasistatic response because of the integration
of the intermittent load process in the surface zone.
The resulting effect to the response is exemplified in Fig. 1 and in
Table 1, which gives some results from a simulation study of the TPG 500
Jack-up platform. 12 This figure demonstrates the highly non-Gaussian
quasistatic response in terms of sample probability distributions, whereas
the table gives estimates of statistical parameters for the different response
quantities. The extreme value parameter defined as the ratio between the
extreme response and the standard deviation of response should be espe-
cially noted, since this directly quantifies the non-Gaussian effect.
The adequacy of these results depends fully on the relevance of the
extended Morison equation. The validity of this model is continuously

I i I I
~q /1
X X X X Max LmO s e r ke8 / I X X X X ~ X kmo #let"LgS; / I
.J
- - - - - ROd L , I . ~ / I

0,9~ -
0.9990- 0.9990 -

0.9900 - 0 . 9 ~ -

0.9000 - 0.9000 -

0.5000 - 0.5000 -

O. I000 - i_ 0.I~- ~
i
, , P 3'5 ' '
0.0000 2.5555 4.8867 7.0000 0.~ 2 55 4.8667 7.0~3CI0
S T . DEV. OF PARENT S E R I E S ST. OEV. OF PAF~NT SF=J:~IES
Base shear" OTM
Bca~e Shear" O~i~lurn , . ~ momen t

Fig. 1. Samplemaxima probability distributions of quasistatic response-- TPG 500 Jack-


up.
74 N. Spidsge, D. Karunakaran

TABLE 1
Statistical Parameters of Nonlinear Quasistatic Response - TPG 500 Jack-Up

Response quantity A verage response statistics*


Skewness Kurtosis Extreme value parameter
Base shear 1.30 6.17 8.3
Overturning moment 1.47 6.91 8.8
Deck displacement 1.49 6.97 8.8
Shear at top of leg 0.99 6.32 6.6
Moment at top of leg 1.75 8.37 9.2
Shear at bottom of leg 1-27 6.05 8.1
Moment at bottom of leg 1.36 6-31 8.5
* Average estimates from 7 samples.

discussed both as concerns the model concept and the choice of coeffi-
cients. Based on comparisons of measurements and simulations, see for
instance Refs 3 and 24, it seems reasonable to conclude that this model
gives a satisfactory representation of the forces and quasistatic response,
provided that the coefficients are properly chosen. C o m m o n practice is to
use flow independent coefficients. It is, however, clear from experiments
that the coefficients are dependent on the flow pattern around the
members and on the leg roughness, 26 and thus may vary significantly over
the structure. There are reasons to believe that the coefficients are signifi-
cantly lower in the splash zone than for the submerged part of the struc-
ture. 26 Industrial practice neglects this, and it appears thus that the normal
choices of coefficients (i.e. Cm = 2-0, Ca = 0.7-1.0) will therefore lead to
reasonable, but most probably conservative results, as concerns the
nonlinear effects to the response. A final verification of this demands full-
scale measurements, but so far necessary comparisons of such to simula-
tions has not been published, though they ought to be available from
different measurement programmes.
Transverse forces are normally not included in global response analyses.
These forces may be modelled as are the drag forces and will consequently
have the same statistical properties. Experiments have indicated that
locally these forces may also be high, but that they are not well correlated
along the legs. If so, they will not have any effect in the global analysis,
and they are therefore neglected in such analysis. However, whether this is
correct can only be confirmed through full-scale experiment.
It is also well known that, under special current conditions, transverse
forces may cause vortex induced vibrations of individual legs. Whether
this is a problem for a specific platform must be checked in each case, as
this is not included in global dynamic response analysis procedures.
Nonlinear dynamic behaviour of jack-up platforms 75

Structural properties

Three different sources o f nonlinear platform behaviour are directly rela-


ted to the structural system:
- - P-~ effects
- - soil structure interaction
- - deck-leg interaction.

P-~ effects

The so called P-6 effect is directly related to the displacement and load
levels in the structural element.2 For a wave loaded structure this leads in
principle to time and load dependent structural system stiffness. Conse-
quently, both the structural flexibility and the natural frequencies may
vary with time, thus effecting both the quasistatic and resonant response.
For jack-up platforms, however, the deck loads will dominate the leg
forces. The major P-6 effect will thus be a static contribution to the linear
stiffness, which only has a minor influence to the dynamic properties of
the platform. This is demonstrated in Table 2, which gives results from a
simulation study, based on the above assumptions.

Soil-structure interaction

Traditionally the soil-structure interaction of a jack-up platform is


modelled for dynamic response analysis as a pinned connection at each leg
or by a set of linear springs and dashpots at the bottom of each leg.

TABLE 2
Comparison Between Dynamic Response Estimated With and Without Including P-6
Effect TPG 500 Jack-up Platform

Response quantity With P-6 effect Without P-6 effect


To = 5.32 sec To = 5.19 sec
St. dev. of Extreme St. dev. of Extreme
response response response response
Base shear (MN) 3.19 24-07 3.16 24.20
Overttmaing moment (MNm) 313-4 2180 295,6 2120
Deck displacement (m) 0-082 0.562 0,077 0.545
Shear at top of leg (MN) 0.77 4.77 0.76 4.80
Moment at top of leg (MNm) 43-90 277.5 41,0 266.1
Shear at bottom of leg (MN) 1.16 8.74 1.14 8-77
Momenlt at bottom of leg (MNm) 75.3 540-4 71.4 528.4
76 N. Spidsoe, D. Karunakaran

However, it is well known that many soil materials exhibit elastoplastic


behaviour. The platform foundation system may thus act as a nonlinear
system with significant hysteretic behaviour. This will influence both the
stiffness and damping properties of the platform. The structural flexibility,
natural frequencies and damping may thus be time dependent, having
consequences for both quasistatic and resonant response.
Nonlinear soil-structure interaction behaviour may be modelled assum-
ing that the foundations behave as a kinematic hardening material with
nested yield surfaces. 16 The foundation is then represented by a set of
nonlinear springs at the so called decoupling point at the leg base, i.e.
where there is no coupling between rotations and translations. Each spring
is modelled by a so called back-bone curve which defines the yield surfaces
and load-deflection characteristics and thus describes the isotropic kine-
matic hardening behaviour, as demonstrated in Fig. 2. These back-bone
curves are established in a separate analysis based on simplified elastic and
plastic models which combines limit equilibrium methods and elastic half-
space solutions for the soil material behaviour. 2s
This soil model implies a hysteretical behaviour of the soil-structure
interaction. An example of such hysteretic behaviour of the foundation is
shown in Fig. 3.
The effects of using models like this are demonstrated in Table 3 which
compares calculated dynamic response with nonlinear foundation beha-
viour and linearized foundation behaviour. The comparison is carried out
for MSC CJ62 jack-up with foundations characterized by the back-bone
curves shown in Fig. 4. Further information on this foundation are given
by Skotheim. 23
These results indicate that the linearized analysis gives very close
standard deviation of response compared to the nonlinear analysis for
the mudline responses. The maximum deviation is about 2%. However,
the standard deviation of the axial force in the leg is lower in the line-
arized analysis than in the nonlinear analysis. Furthermore, there are

OS
fl "fs "'""

Fig. 2. Foundation back-bone curves.


Nonlinear dynamic behaviour of jack-up platforms 77

j
F i

Fig. 3. Hysteretical behaviour corresponding to the back-bone curve shown in Fig. 2.

TABLE 3
Ratio of Dynamic Response Using Linearized and Nonlinear Soil Stiffness

Response quantity Ratio between response using linearized and nonlinear


soil-structure interaction
St. deviation of response Extreme response
Base shear 1.002 0-974
Overturning moment 0.997 0.940
Axial force in leg 0.963 0.796
S.F. in leg at M W L 1.005 0.097
B.M. in leg at M W L 0.944 0.799
S.F. in leg at - 3 6 0.996 0.964
S.F. in leg at mudline 0-994 0.994
B.M. in leg at mudline 1-022 1.187

12 45, i , , , 450.
10 _ ,oL
~'3oh'
~6 ~.2501-
i 2o ~2ooI-
,,o4 Liaearizgl.... . . . .
~Lm~iz~l '
2 51-
01 ' o.,~1 ' o.~ ' o.~ 'o.o4 o o.b~"'o'.1 o.~5 o'.2 o.:5 ~ 0.11 0~2 0b3 0.h
Displacemont[m] Displacgmcnt[ml Rotation[radsl
Horizontalload-deflectioncurve Verticalload-deflectioncrave Moment-rotationcurve
Fig. 4. Foundation characteristics.

significant differences between the extremes obtained by the two types


of analyses.
The total overturning moment is for this structure mainly taken as axial
force in the legs and partly as leg bending moment. As the load level
78 N. Spidsee, D. Karunakaran

increases, the rotational stiffness decreases more than the vertical stiffness
due to their different stiffness properties, see Fig. 4. When the rotational
stiffness decreases more than the vertical stiffness, the leg bending moment
also decreases, while the axial force increases. This gives that a redis-
tribution of the moment takes place in the nonlinear analysis due to the
stiffness reduction. In the linearized analysis, there is no nonlinear soil
stiffness effect and hence there is no redistribution of moments at high
load levels. This is seen from Fig. 5, where the total overturning moment,
the axial force in the leg and the moment at the bottom of leg are shown
for both linearized and nonlinear analysis at an extreme load cycle.
It is seen from Fig. 5 that the axial force in the nonlinear analysis has
higher maxima than in the linearized analysis and that the bending moment
in the bottom of the leg has less peaked maxima in the nonlinear analysis
than in the linearized analysis. Due to this moment redistribution, the axial
force in the leg has slightly higher standard deviation of response in nonlinear
analysis than in linearized analysis. Furthermore, this redistribution produ-
ces a non-Gaussian response with larger maxima than the linearized analysis
for the axial force, see Fig. 6. On the other hand this redistribution gives
smaller maxima for bending moment in the bottom of the leg from nonlinear
analysis than the linearized analysis, see Fig. 7. The other response quantities
which have similar behaviour to the axial force are the bending moments in
the upper part of the structure. Furthermore, the shear force in the upper
part of the structure has similar behaviour to the moment at the base.
Due to this redistribution phenomenon, the extreme response predicted by
linearized analysis deviates significantly from the response derived by nonlin-
ear analysis. The axial force in the leg estimated using linearized analysis is up
to 25% smaller than the extreme response from nonlinear analysis, so is the
bending moment at the upper part of the structure where the maximum
deviation is up to 29%. Furthermore, the deviation between the nonlinear and
linear response increases as the foundation moment capacity increases.

Deck-leg interaction

The deck-leg connection is one of the highly loaded parts of a jack-up plat-
form. The interaction between the deck and the leg may be nonlinear due to
the guide clearances, system backlash, brake slippage, etc. This connection is
often modelled by nonlinear spring stiffnesses) 5 These nonlinear springs
may principally imply hysteretic behaviour, with effect both to the natural
frequencies and the structural damping. However, the deck-leg interaction
will only have significant influence on the moment distribution on the leg
section within and around the connection. Hence, for a global dynamic
analysis it may be idealized by linearized springs, which is often used.
Nonlinear dynamic behaviour ofjack-up platforms 79

1400 i i I i i

//
1200 _ e~ No~m:~---- ....

1000

8OO
60O
1
400
200

0
-200 - -tO"
-400 I I I I I
280 285 290 295 300 305 310
T i m e [sec]

Overturning moment

16
L i I i i i j
14 /~ Nordlnoar

~ 6
4
2
0
-:2
-4 I
280 285 290 295 300 305 310
Tinae [see]

Axial force in leg

I I
t~ Nonlinear
ii
Linear ....
150

i
50

t
-50 I I I
280 285 290 295 300 305 310
Time [see]

Moment at bottotn of leg

Fig. 5. Schematic representation of redistribution of moments.


80 N. Spidsoe, D. Karunakaran

i I Z i I I I I I

. . . . Wm
tbut t ~ / /

0.9990 - 0.9990
0.9900 - 0.9900
0 9000
0.9000 -

0.5000-
0.4000- ,S
F I
4
I
8
I
t2
I
t6
I I
20
L 0 .SO00
0 . 4000

0.0
'2"
4.4
I I
8.8
I
t5.2
'
Ii
47.6 22.0
MN MN
Force Force
Non L ,ne(~r 8 0 t L mode t L tnoer" t z e d 8 0 k L mode L

Fig. 6. Axial force in leg - - n o n l i n e a r a n d linearized soil stiffness.

I I I I I I I I I I
X X X X MOX kmo 81b"tile, /
- - -- ~ RoI=I La tgh /
WetNLL - 3 /
/ n

o.999o-
0.9990 - -~-'~
o .ssoo - I o.ssoo - x,x/~x~eX~ex ^ I
0.9000 - f X " " I
o.9ooo - i
0.5OOO -
0.9000 - I
0. I000 - 0. tO00- I
I
0 ~0 ,00 '1~0 2~0 ~-0 ZOO 0 5~0 ,00 ,;0 2~0 2~0 300
MNm MNm
Moment Moment
Non L tnaoP s o t t modg L L tnaeP t z o d s o tt mode L

Fig. 7. Bending m o m e n t in leg at mudline - - n o n l i n e a r a n d linearized soil stiffness.

Dynamic effects

These are effects due to


- - damping mechanism,
- - superharmonic excitation,
- - springing and ringing.

Damping

As for any bottom fixed offshore platform, the damping of a jack-up


platform consists of structural damping, soil damping and hydrodynamic
damping.
The structural damping is commonly modelled as proportional damp-
Nonlinear dynamic behaviour ofjack-up platforms 81

ing using a Rayleigh model specified by the modal damping of the lowest
natural modes of the structure.
If a linearized model is applied for the soil-structure interaction, the
damping may be specified through the damping coefficients of dashpots.
Alternatively, the soil damping is, however, normally negligible for slender
structures such as a jack-up platform. If it is accounted for, this may be
done by including it in the Rayleigh model used for the structural damp-
ing. Tile drag damping which dominates the hydrodynamic damping is
implicitly given by the extended Morison equation used for the wave
loading.
There are three different damping mechanisms involved:
- - the linear viscous mechanism modelled by the Rayleigh model;
- the nonlinear drag damping mechanism, modelled by the extended
-

Morison equation;
- - the hysteretic soil damping mechanism, modelled by the kinematic
hardening model.
These mechanisms have different properties and effects to the dynamic
response which may have significant impact to the response extremes.
The most important and least known property of the linear viscous
damping mechanism is the so called Normalization effect, i.e. its ability to
generate a Gaussian resonant response independent of the statistical
nature of the excitation process. 4 This effect is demonstrated in Fig. 8 and
Table 4, which summarize results from a study on damping for the TPG
500 jack-up. This example shows that, even if the resonant response is very
high in the rms sense, its contribution to the extreme response is signifi-
cantly less because the total response is less non-Gaussian than the quasi-
static response because of the nearly Gaussian resonance generated by the
damping mechanism. The drag damping mechanism has a similar effect,
though the mechanism is different. As also shown by Brouwers4 this
damping mechanism generates a non-Gaussian resonant response with a
lower Kurtosis coefficient than in a Gaussian process, i.e. with a statistical
nature opposite to the quasistatic response. When the resonant and
quasistatic response is added an even stronger Normalization effect to the
total response occurs than generated by the linear viscous damping model.
The nonlinear properties of the hysteretic soil damping mechanism
depend highly on the shape of the back-bone curve. The more nonlinear
this curve is, the wider will the hysteretic curve be at the high load cycles.
Consequently, the damping in these response cycles will be higher than in
the lower. This is similar to the drag damping mechanism. Though it has
not been investigated so far, this means that one may assume that a
Normalization effect similar to what is found for the drag damping will
82 N. Spids~e,D. Karunakaran

I I I I I I I /
I
x x x x Mclx tmo Isr-Ueli X X X X M o x ~.mo 6 i s P L,e 8
-I
. . . . R o W t e Lgh . . . . R o W L Lgh
. . . . WoLbuLL-S F Lt
<= . . . . WetbuL L-3 r t l

0,9990- 0.9990-
0,9900- 0.9900-
0,9000- 0.9000-

O.,~O00- 0.S000-
O,JO00- 0.I000-
i i i i i
0 I
ST.
I
DEV. OF
;
PARENT
*'S E R I E S ,0 5
T.
4
OEV. OF
S 8
PARENT SERIES
I0

Momon t M o m o n t.
Ouoe L s t o t k c r-elserise D~,,,inomc response

I I I I I I I I
x x x x M o x Lmo l e t Les x x x x Mox Lmo e e P Le8
-I - - " ~ --'- ~ R o u l " kgh / -- -- -- RO W t e ~.gh

//
We t b u t I . - S ~
=e /
/ /
0 8990 - 0.9990
0.9900 - 0.9900
0 . SO010 - 0.9000
0 .SO00 - O.SO00
O. I000 - 0.t000
i I i l i
2 4 6 8 t0
ST. DEV. OF PARENT SERIES ST. OEV. OF PARENT SERIES
Homn Momen ~,
Dynom LC w o v e response R o e n o ni Peepnee

Fig. 8. Maxima probability distributions for overturning moment of TPG 500 jack-up
with 3% linear damping only.

also occur for this soil damping mechanism. This is probably partly the
explanation of the different statistical nature observed between linear and
nonlinear soil structure interaction models demonstrated in Fig. 6.
The different damping mechanisms are not independent, but work
together. This is demonstrated by the results given in the last part of
Table 4. From these it is seen that when the linear damping reduces the
dynamic amplification of the rms response increases. This is as expected
since the total damping level reduces. However, it is also noted that the
dynamic amplification of the extremes is almost independent of the linear
damping level. The explanation of this is that the reduction of linear
damping is compensated for by the Normalization effect both 'from the
linear damping itself and also from the drag damping which becomes
more important as the linear damping level reduces. From a design point
of view this is interesting because it implies that, as long as both these
damping models are involved, it is not important with an accurate speci-
fication of the linear damping level. However, since neither the linear
Nonlinear dynamic behaviour of jack-up platforms 83

TABLE 4
Dynamic Amplification Factors (DAF) for Various Damping Combinations TPG 500
Jack-up

Case identification Estimated Base shear Overturning moment


damping
DAFfor St. DAFfor DAFfor St. DAFfor
deviation extremes deviation extremes
Linear clamping only
0-5% Ln. damping 0.52% 2.47 1.39 3.89 1.83
1.0% L~n. damping 1.03% 1.93 1.24 2.90 1.51
2.0% Liin. damping 2.07% 1.58 1.18 2.22 1.34
3.0% Liin. damping 3-11% 1.43 1.18 1.93 1.29
4.0% Lin. damping 4.15% 1.36 1.18 1.76 1.29
5.0% Lin. damping 5.19% 1.31 1.17 1.65 1.30
Linear damping and drag damping
No Lin. damping 2.76% 14-51 1.12 2.09 1.24
1.0% Lin. damping 3-10% 1.41 1.12 1.90 1.22
2.0% Lin. damping 4.16% 1.34 1.12 1.73 1.22
3.0% Lin. damping 5.21% 1.29 1.12 1.63 1.23

damping mechanism nor the drag damping mechanism may be regarded


as verified, this effect should be utilized with care.
For further discussion on the nonlinear damping properties, reference is
made to Spidsoe et al. 27

Superharmonic excitation

Super.harmonic excitation is directly related to the nonlinear hydro-


dynamic loading and could therefore just as well be classified as an effect
of thi,;. The effect is well known. 25 It is caused by the higher order load
components of the nonlinear load process which increase the excitation
level in the high frequency range where the natural frequencies of the
platform are located and thus increase the resonant response compared to
only natural excitation at these frequencies.
The, effects of superharmonic excitation depend on

--- the load pattern,


--- the ratio between wave peak frequencies and natural frequencies,
--- the degree of natural excitation at the natural frequencies.

The superharmonic properties of Morison equation are described by


Gudrnestad and Connor 6 who show by Taylor expansion that if the wave
peak frequency is COo,then superharmonic components are generated at
84 N. Spidsoe, D. Karunakaran

- - 3090, 5O)o, ... if the loading is pure wave loading,


- - 2~Oo,4090, ... if current is present,
- - 2090, 4090, ... by the integration of forces to the instant surface
elevation.

The effects of superharmonic excitation are thus strongest when the


higher order components match the natural frequencies of the platform,
i.e. when the ratio between wave peak period corresponds to multiples
of the highest natural periods. Since the wave period varies from cycle
to cycle in an irregular wave process, the pattern of superharmonic
components will also vary with time. Superharmonic excitation has
therefore clear similarities to springing and ringing which will be
discussed later.
One implication of superharmonic excitation which should be
mentioned is its influence to the Normalization properties of the linear
viscous damping model. As shown by Brouwers, 4 the full Normalization
effect requires that the excitation is broad-banded around the natural
frequencies. This is normally true for direct excitation in the high
frequency range. However, the spectral properties of the superharmonics
are not well known and it may very well be that in many cases they are
narrow-banded and thus prevent Normalization. This may be especially
the case when the major super-harmonic matches exactly the natural
frequency and the direct excitation at these frequencies is low.
The lower the first natural frequency of the platform is, the stronger is
the direct excitation at this frequency. The relative importance of super-
harmonic excitation is thus highest at high natural frequencies. I n these
cases, however, resonance is generally of minor importance for the total
response. It is thus difficult to give a general statement on when super-
harmonic excitation is most important. This is demonstrated by the
example study carried out for the TPG 500 jack-up summarized in Table 5.
Further studies on the effects of superharmonic excitation are also
presented by Haver. 7

Springing and ringing

Springing and ringing are dynamic response phenomena which recently


have been highlighted during the design of large volume offshore struc-
tures with dynamic behaviour such as gravity platforms and tension leg
platforms. Both phenomena are caused by higher order wave loads in the
surface region which may induce significant resonant response. There are
different and ambiguous definitions of springing and tinging. The follow-
ing definitions are, however, adopted in the joint industry project 'Higher
Nonlinear dynamic behaviour of jack-up platforms 85

TABLE 5
Eflbcts of Superharmonic Excitation to the Response of the TPG 500 Jack-up
Response quantities To = 3.3 s To = 5.3 s To = 6 s

Base shear 0-64 0.86 0.46 0.93 0.28 0.98


Overturning moment 0.65 0.84 0.48 0.80 0.29 0-88
Deck displacement 0.66 0.82 0.48 0.93 0.28 0.94
Shear at top of leg 0.70 0-88 0.52 0.92 0.30 0-85
Moment at top of leg 0.66 0-86 0-49 0.80 0.29 0-81
Shear at bottom of leg 0,64 0.93 0.48 0-80 0.28 0-92
Moment at bottom of leg 0.64 0.92 0.48 0.80 0.29 0-91
Note: R~r is the ratio between the standard deviation of resonant response calculated
without direct wave excitation at the resonant frequency and with direct wave excitation.
Re is the ratio between the extreme dynamic response estimated without direct wave exci-
tation at the resonant frequency and with direct wave excitation.

Order Load Effects o f Large Volume Offshore Structures', N T H , D N V


and SINTEF:

- - springing is the steady state effect,


- - ringing is the transient effect.

Three examples of measured ringing effects on large volume structures


derived from model test data are shown in Fig. 9. These show the typically
fast build up followed by a damped decay of the resonant response which
gives the bell-shaped response history which is the origin of the concept
ringing.
Though springing and ringing effects may be significant for large
volume structures, it is an open question whether they are relevant for
drag-dominated structures. On the basis of the results from the above
mentioned joint industry project it appears that some higher order load
mechanisms which cause the effects on large volume structures also m a y
be of importance for slender structures. Arguments for this m a y also be
found in the so called slender body theory presented by Rainey. 21 It is also
obvious that these p h e n o m e n a have similarities to the superharmonic
excitation discussed earlier. Furthermore, a ringing-like response has been
observed from the full-scale measurement o f jacket platform response. A n
example o f this is shown in Fig. 10. However, though the springing and
ringing p h e n o m e n a are present it is not sure that they have the same
importance for slender structures. F o r large volume structures they are
important because resonance amplifies the m a x i m u m quasistatic response
induced by inertia wave forces. Due to the drag force dominance the
86 N. Spidsoe, D. Karunakaran

i I I J

1650 1850
Bending moment Time (sec)
Monotower gravity platform

J llJ I l
t -- -t- - - --I-- --(

, ,
2250 2450
Overturning moment Time (sec)
Four-legged gravity platform

I i i i I

I i 'l i I
8600 8800
Tether tension Time (sec)
Tension leg platform

Fig. 9. Examples of ringing events observed from model tests of large volume structures.

quasistatic response of slender structures is phase shifted, and the ampli-


fication occurring for large volume structures thus need not take place for
these structures. Simulation studies with slender single columns based on
the Morison equation for the wave forces have indicated this.
It has been checked by careful examination of simulated resonant
response in the example studies referred to earlier in this paper whether
Nonlinear dynamic behaviour ofjack-up platforms 87

Transient resonant response


/ /
6
4
2
0
-2
-.v
-6
0 $0 ~00 IS0

Significant wave height: H, = 4.6 m

Fig. 10. Ringing phenomenon seen from full-scale measurements - - F R I G G DP2 jacket.

the employed wave load models cause ringing effects in these cases. No
such effects have been observed. This means, however, that the problem is
not irrelevant. Improved or extended wave and wave load models may
give different result. It is therefore recommended that ringing and spring-
ing are investigated further for slender offshore structures.

ANALYSIS OF NONLINEAR DYNAMIC RESPONSE

Since jack-up platforms are highly dynamically sensitive, the design sea
state may be different from the extreme sea state which defines the 100
year wave height. This implies that, in principle, a long-term response
analysis which includes the contributions from all possible sea states to the
extreme response and which includes all the nonlinear effects discussed
above should be employed in dynamic response analysis of these plat-
forms. A procedure for this type of analysis tailored for drag dominated
platforms is described by Farnes 5 and Karunakaran. z2 As will be discus-
sed in the next section such analysis is, however, hardly applicable as a
practical design tool. For most jack-ups, it will not be required since the
extreme dynamic response in most cases will be dominated by extreme sea
states. This is demonstrated in Table 6 which gives results from a nonlin-
ear long-term dynamic response analysis of TPG 500 jack-up with varying
dynamic properties.
The long-term procedure is therefore not further discussed here, and it
is focused on short-term nonlinear response analysis. The following
discussions are, however, also relevant for long-term analysis as the short-
term modelling of the extreme dynamic response forms the basis for this
analysis also.
The short-term analysis of the nonlinear dynamic response requires
stochastic time domain simulation techniques. This is because the different
nonlinear phenomena involved may only be properly modelled in the time
domain as discussed above, and because realistic representation of the
dynamic behaviour of the platform demands stochastic modelling of the
88 N. Spids~e, D. Karunakaran

TABLE 6
Nonlinear Long-term Response of TPG 500 Jack-up Platform

Natural period (s) Base shear ( M N ) Overturning moment ( M N m )


Long term Percentage Long term Percentage
response contribution from sea response contribution from sea
states below 7 m states below 7 m

4.40 22.72 0 2171 0


5.30 21.74 0 2055 0
6.00 22.81 0 2563 80
7.00 23.17 0 2783 88
8.00 21-67 0 2419 20

Note: When the percentage contribution from sea states below 7 m is zero, the extreme
long term is completely dominated by the 100 year sea state.

load and response process. Such analysis may be performed as an inte-


grated analysis where the waves, hydrodynamic forces and dynamic
response are calculated simultaneously at each time step or by a stepwise
procedure where the wave process, the force process and the dynamic
response process are established sequentially.
Since the nonlinear behaviour of jack-up platforms is mainly related
to the loading process and to the soil-structure interaction, the nonli-
nearities are connected to the load vector on the right hand side of the
response equation and to a few response degree of freedom systems
which may be isolated from the total structural system. A stepwise
analysis procedure may thus be applied for such platforms, and this is
recommended due to its computational efficiency compared to inte-
grated analysis. A tailor-made analysis procedure for drag dominated
fixed offshore platforms based on this principle which includes all the
nonlinear effects discussed above is presented by Karunakaran. m This
procedure has shown to be a very applicable tool in wide number of
jack-up studies.
Independent of analysis procedure, time domain stochastic dynamic
response analysis implies two basic problems:
-- how to decide the simulation scheme, i.e. select time step, sample
length and number of samples;
- - h o w to derive estimates of extreme response from simulated
samples.
Except for the selection of time step which has to be decided on the
basis of the natural frequency of a structure and the property of solution
algorithms for the dynamic response equation, these problem areas are
Nonlinear dynamic behaviour of jack-up platforms 89

interrelated. This is because the accuracy of a chosen extreme value esti-


mator depends on sample length and number of samples. Principally there
are two different estimation methods:
--sample statistics applied directly to the simulated response
samples;
-- sample statistics based on extrapolation of simulated response
samples.
The first alternative requires a sample length corresponding to the
specified storm duration, i.e. very long samples, of the order 3-6 hours.
Even with rather simplified structural models and modern computer
resources this requires very high computer demands. The second alter-
native is thus the most attractive, from a practical point of view. A central
element of this alternative is the extrapolation of simulated samples. There
are different ways of doing this. The analysis procedure described by
Karunakaran 1 is based on extrapolation of sample probability distribu-
tions using Weibull probability models. This extrapolation procedure is
investigated through an extensive simulation study applying the MSC
CJ62 jack-up platform described in Appendix B. With a reference case
consisting of 100 independent six hour long samples, the applicability of
the Weibull model was investigated and the statistical properties of the
Weibull extreme value estimator evaluated. From this study it was
concluded that the Weibull models give a satisfactory fit to jack-up
response provided that the model parameters are estimated by the method
of matching statistical moments for global response maxima, see Farnes. 5
This is demonstrated in Fig. 11 which compares sample extreme value
distributions to Gumbel extreme value distributions fitted to sample
extremes and Weibull estimated extremes.
The findings on the extreme value properties of the Weibull estimated
extremes are summarized in Fig. 12 which shows how the accuracy of the
extreme estimates relate to sample length and number of samples.
Based on the results from this study the following procedure for simu-
lation and estimation of extreme response is recommended:

(1) choose a practical simulation length based on required structural


modelling complexity and available computer resources;
(2) decide a target value for the coefficient of variation, Car, for the
average extreme response;
(3) simulate N samples (of the order 10), fit Weibull models to the
sample maxima distributions by the method of matching statistical
moments for global maxima and calculate estimates of extrapolated
,;ample extremes for each sample by
90 N. Spidsoe, D. Karunakaran

3
>. 2

0 ~ = = , ~ v C m m ~ l dist. f r o m flttext W e i b u l l
-1 f o r 14 s e t s o f 7 s a m p l e s e a c h -

-2 I I I I I I
14 16 18 20 22 24 26 28
Base Shear [MN]

Fig. 11. Sample extremes plotted on Gumbel probability paper - - base shear.

0.16 I I I I I I I I I
22rain-sample
0.14 I'~ 4 5 r a in - s a m p le - - - - - --
II 90rain-sample ....
0.12 -~ 180min-sample ........
I1~ 360rain-sample -----
0.1
go
0.08

._= 0.06

c; 0.04

0.02

0
0 10 20 30 40 50 60 70 80 90 100
Number of samples

Fig. 12. C.o.v. in average extrapolated extreme response vs simulation length and number
of samples - - base shear.

0.57722 '- ]
= + [(ln ( N s ) ) h + - - fl (ln (Ns))~]-~ (1)
where

/~ -- Weibull l o c a t i o n parameter
tr -- Weibull scaling parameter
/3 -- Weibull shape factor
Ns -- number o f m a x i m a in storm duration;

(4) calculate the average ~xe standard deviation, ~x, a n d coefficient o f


variation Cxe for the extreme response as f o l l o w s

1 N
-flx,= ~ .~-'~~x,)i (2)
7"7
Nonlinear dynamic behaviour of jack-up platforms 91

axe = ((#xe); - xe) 2 (3)


i=l

Cxe = ax,
_ (4)
#x,

where N is the number of independent simulation samples;


(5) calculate required number of samples to meet the target value of
C.O.V., C T from:

NR = (5)

(6) simulate additional samples if needed, i.e. NR > N, estimate corre-


sponding extreme responses, update the average extreme response
and use this as the design extreme value.
Though Fig. 12 is based on a specific example structure, it may be
generalized to serve as a guide to select initial parameters to this proce-
dure.

DESIGN ANALYSIS

It follows from the above discussions that the highly nonlinear dynamic
behaviour of jack-up platforms implies a wide range of significant uncer-
tainties related to structural analysis methods, to mathematical models for
structural behaviour, loading and response to the parameters involved in
these methods and models. On this background it is, at least from an
academical point of view, tempting to recommend that structural relia-
bility analysis should be included in the design process. Examples of such
analyses are presented by Jensen e t al., 8 L~seth e t al. 14 and Karunakaran
e t al. 13 Furthermore, a tailor made procedure for fixed, drag-dominated
offshore platforms based on nonlinear long-term dynamic response
analysis is established and demonstrated by Karunakaran. 12 However,
even if the tools are available and the analysis technique is proven, it
seems obvious that reliability analysis can not be a practical design tool.
This is because it requires specially qualified personnel, it is extremely
computer demanding and, maybe most important, it requires statistical
information with sufficient confidence on a wide number of models and
parameters which may not be available for a specific platform. There is,
however, a sufficient basis for extensive reliability studies of jack-ups as a
class of structures. Thus, it is possible to use this type of analysis for
92 N. Spidsoe, D. Karunakaran

establishment of calibrated codes which gives the design engineer consis-


tent load factors and resistance factors to be applied to extreme values
derived from the design structural analysis. Calibration of load factors by
reliability method for drag dominated structures is demonstrated by
Olufsen. 17
Adopting the concept of load factors, the design problem is then, from
the response point of view, reduced to calculate the extreme response. In
order to be consistent with the expected nonlinear stochastic dynamic
behaviour of jack-up platforms, this should formally be done using the
nonlinear long-term response analysis or, if there exists a well defined
design sea state, by short-term stochastic analysis. However, due to the
complexity of the analysis methods, their computer demands and inherent
uncertainties, this approach is not very efficient. With respect to design
structural analysis it appears thus to be attractive to establish a practical
procedure which links the sophisticated nonlinear long term and short
term analysis methods with the traditional design wave methods in a
manner which is consistent with a complete reliability analysis.
Based on the experiences from a wide number of jack-up studies the
following procedure is suggested:
(1) identify the short term design sea state by long-term analysis;
(2) estimate dynamic amplification factors by Short term stochastic
response analysis;
(3) establish extreme values for the dynamic response by combining
design wave analysis and the estimated dynamic amplification
factors;
(4) establish design response values by applying calibrated load factors
to the estimated extreme dynamic response.
The first step is required since the design sea state for a jack-up due to
its dynamic behaviour may be different from the sea state which defines
the 100 year design storm. As shown by Karunakaran et al.ll it is required
to use a nonlinear procedure in order to obtain correct identification.
However, as demonstrated in the previous section, this investigation is not
needed for any platform as it appears that for structures with low natural
periods, below 6 s, the extreme sea state will be the design sea state.
Estimation of dynamic amplification factors, D A F , is done based on the
following definition:

OAF -- Xe'dyn (6)


Xe, qs
where Xe, dr, is the extreme dynamic response and Xe, qs is the extreme
quasistatic response, Which analysis method should be used for calcula-
Nonlinear dynamic behaviour of jack-up platforms 93

tion of these extreme responses depends on the structural systems linearity


properties and on the design sea state. If the design sea state is low, i.e. the
sign&ant wave height is below 6m and the structural system in this sea
state behave linearly, then a traditional frequency domain analysis22 may
be applied. The calculation of DAI; is then straight forward as the
extremes are directly given from the analysis in a unique way. In all other
cases a. nonlinear method which implies time domain simulations of
dynami.c response must be involved, rising the, question on how a reliable
estimate may be obtained from simulated response samples. Since the
basis for calculation of DAF is the estimate of extreme responses, it
appears reasonable to apply the following estimator for DAF in such
cases:

DAF = f $ DAFi (7)


r=l

Here, the index i indicates individual sample estimates while N is the


number of independent samples. This implies that the problems related to
choice of simulation length and number of samples which were discussed
on extremes are also relevant for DAFs. The results presented in the
previous section on extremes are thus applicable for DAFs. However, if
the same wave process realization is used for the calculation of both
quasistatic and dynamic response samples, the statistical uncertainties
related to the simulation procedure is removed. Hence the variability of
the DAF will be less than for extremes, from which it follows that shorter
and fewer samples may be used for DAFs than for extremes in order to
achieve the same target variability. This is exemplified in Table 7, which
compares coefficients of variation for extremes and DAFs for the TPG
500 jack-up platform.
When the DAFs are established, the extreme values for a dynamic
response are derived by multiplying these with extreme responses, Xe,DWA,
calculated by a traditional design wave analysis
X, = DAF- X,,. (9)
Extreme values which are consistent with the long-term analysis are
thus derived through an extension of traditional methods which should
not be regarded as too problematic, though it may involve advanced
analysis methods.
There is one problem related to this procedure which should be
94 N. Spideoe, D. Karunakaran

TABLE q
Dynamic Response and Corresponding D A F s with C.O.V.s - - TPG 500 Jack-up

Response quantity Standard deviation o f response Extreme response


Resp. c.o.v. DAF c.o.v. "Resp. c.o.v. D A F c.o.v.
Base shear 3.19 0.022 1.22 0.016 24.1 0.096 1.12 0.045
Overturning moment 313.4 0.036 1-50 0.030 2180 0-098 1.19 0.057
Deck displacement 0.082 0.037 1.55 0.033 0.562 0.097 1.20 0.058
Shear at top of leg 0.77 0.046 1.54 0.035 4.77 0.083 1.45 0-080
Moment at top of leg 43.9 0.045 1.73 0.038 277.5 0-103 1.17 0.064
Shear at bottom of leg 1.16 0.022 1-25 0.019 8-74 0.084 1.21 0.035
Moment at bottom of leg 75.3 0.032 1.43 0.028 540.4 0.092 1-22 0.048
Note: Shear in MN, moment in M N m and displacement in m.

mentioned. Formally eqn (9) required that there be consistency between


the quasistatic response calculated by the design wave method and by the
nonlinear stochastic analysis. Since the wave theories applied in these two
methods are basically different, i.e. describe different flow conditions, it is
not possible to obtain consistent extremes with the same drag coefficients.
Provided that a drag coefficient is specified for the design wave conditions,
this requires that a drag coefficient be used in the stochastic analysis must
be derived though a calibration of these analyses. Practically, this may be
done using equal structural models and response models in the two
analysis methods and adjusting the drag coefficient in the stochastic
analysis until equal estimates of the quasistatic response are obtained.
Normally, this will give higher drag coefficients in the stochastic analysis
than in the design wave analysis.

CONCLUSIONS

Sources of nonlinear behaviour of jack-up platforms and modelling of


their underlying mechanisms have been overviewed and exemplified
through simulation studies. Furthermore, structural analysis techniques
which may account for this behaviour are discussed with respect to
available methods, their practical use and integration in design proce-
dures.
From the given discussions and examples the following conclusions may
be drawn:

- - the major sources of nonlinear behaviour are the wave loading, the
damping mechanisms and the soil-structure interaction. The
Nonlinear dynamic behaviour ofjack-up platforms 95

nonlinear effects to the extreme dynamic response of jack-ups are


significant and must be accounted for in the design process.
Mathematical models for the underlying phenomena, which are
applicable in structural analysis exist. However, several of these
models may not be regarded as fully verified;
the inclusion of nonlinear dynamic platform behaviour in the
design process implies that nonlinear stochastic dynamic response
analysis based on time domain simulation methods must be
applied. For most platforms a short term analysis method is
sufficient with respect to estimation of extremes. The estimation of
extremes from simulated samples may then be considered compu-
tationally efficient and statistically consistent by using Weibull
models for extrapolation of short samples following the procedure
discussed in this paper. For platforms with high natural periods, a
long term analysis may be required, at least to define the short
term design sea state;
the complicated and nonlinear dynamic platform behaviour
implies that a wide number of significant uncertainties are intro-
duced to the design process through the included mathematical
models, analysis methods and the practical use of these methods.
A calibrated code giving consistent load and resistance factors
should therefore be established through an extensive reliability
analysis for jack-ups as a class of structures. For the design of
specific platforms these load factors should be used together with
extreme responses calculated by traditional design wave analysis
and dynamic amplification factors derived from stochastic
dynamic response analysis in order to obtain consistent design
values for the dynamic response.

Apparently, all elements which are needed to establish an industrial


design practice are available. What remains is thus the industrial agree-
ment which is needed to initiate the final implementation and work for
necessary verification of the procedure.

ACKNOWLEDGEMENT

The experiences and examples presented in this paper are derived mainly
from studies carried out on contract for Statoil. The company is highly
acknowledged for release of the results. Furthermore, O. T. Gudmestad,
M. Ba~rheim and S. Haver, all from Statoil, have, through many valuable
discussions and suggestions during the projects, contributed significantly
96 N. Spidsee, D. Karunakaran

to the content o f this paper, and they are highly acknowledged for this
support.

REFERENCES

1. Borgrnan, L. E., Computation of the Ocean-Wave Forces on Inclined


Cylinders. J. Geophysical Research, 39 (1958) 885-888.
2. Borgman, L. E., Random Hydrodynamic Forces on Objects. Ann. Mathe-
matics, Statist. 38 (1967) 37-51.
3. Brathaug, H-P., Karunakaran, D. & Spidsoe, N., Morison equation fitted to
measured forces from irregular waves on a vertical pile: estimated coefficients
and model evaluation. SINTEF Report STF71 A89025, Trondheim, 1990.
4. Brouwers, J. J. M., Response near resonance of nonlinearly damped systems
subjected to random excitation with application to marine risers. Ocean
Engng, (1982).
5. Fames, K-A., Long-term statistics of response in non-linear marine struc-
tures. Dr. Ing thesis, Norwegian Institute of Technology, 1990.
6. Gudmestad, O. T. & Connor, J. J., Linearization Methods for Nonlinear Drag
Force Loading Terms. Massachusetts Institute of Technology, Boston, 1980.
7. Haver, S., On the relative importance of various load mechanisms regarding
an idealised Jack-up. STATOIL research Report, 1990.
8. Jensen, J. J., Madsen, H. O. & Pedersen, P. T., The effect of nonlinear wave
force model on the reliability of a jack-up platform. In Proc. 3rd IFIP Working
Conf. on Reliability and Optimization of Structural Systems, Berkeley, 1990.
9. Karunakaran, D., Nonlinear dynamic response analysis of TPG 500 jack-up
platform with parametric study. SINTEF Report STF71 A90010, Trond-
heim, 1991.
10. Karunakaran, D., Procedure for nonlinear dynamic response analysis of
offshore structures - - both for extreme and fatigue response. SINTEF
Report STF71 A91016, Trondheim, 1991.
11. Karunakaran, D., Spidsoe, N. & Gudmestad, O. T., Selection of design sea
states for Jack-up platforms with dynamic behaviour. In Proc. First Int.
Offshore and Polar Engineering Conference, Edinburgh, Scotland, 1991.
12. Karunakaran, D., Nonlinear dynamic response and reliability analysis of
drag-dominated offshore structures. Dr. Ing. thesis. Division of Marine
Structures, Norwegian Institute of Technology, 1993.
13. Karunakaran, D., Spidsae, N. & Leira, B. J., Prediction of extreme dynamic
response of jack-up platforms using a nonlinear time domain simulation
method. In Proc. 12th Int. OMAE Conf., Glasgow, Scotland, 1993.
14. Loseth, R., Mo, O. & L~tsberg, I., Probabilistic analysis of a jack-up plat-
form with respect to the ultimate limit state. In Proc. First European Offshore
Mechanics Symposium, Trondheim, 1990.
15. Mommaas, C. J. & Dedden, W. W., The development of a stochastic
nonlinear and dynamic jack-up design and analysis method. In Proc. Second
Int. Conf. On The Jack-up Drilling Platform, London, 1989.
16. Mrrz, Z., On the description of anisotropic hardening. J. Mech. Phys. Solids,
15 (1967) 163-175.
Nonlinear dynamic behaviour of jack-up platforms 97

17. Olufsen, A., Uncertainty and reliability analysis of fixed offshore structures.
Dr. Ing Thesis, Norwegian Institute of Technology, 1989.
18. Penzien, J. & Tseng, S., Three-dimensional dynamic analysis of fixed offshore
platforms. In Numerical Methods in Offshore Engineering, eds O. C. Zienkie-
wicz et al., John Wiley & Sons, NY, 1978.
19. Pierson, W. J. & Holmes, P., Irregular wave forces on a pile. ASCE J.
Waterways and Harbours, WW4 (1965) 1-10.
20. Przemieniecki, J. S., Theory of Matrix Structural Analysis, Mc-Graw Hill,
USA, 1968.
21. Rainey, R. C., A new equation for calculation of wave loads on offshore
structures. J. Fluid Mech., 204 (1989).
22. Sigbj6rnsson, R., Stochastic theory of wave load processes. Engineering
Structures, 1, (1979) 58-64.
23. Skotheim, A. A., Foundation analysis of jack-up for Sleipner, B., Geovest
report, 1992.
24. Spidsoe, N., Brathaug, H-B. & Skjhstad, O., Nonlinear random wave loading
on fixed offshore platforms. In Proc. OTC, Houston, USA, 1986.
25. Spidsoe, N. & Karunakaran, D., Effects of superharmonic excitation to the
dynamic response of offshore platforms. E&P Forum Workshop on Wave and
Current Kinematics and Loading, IFP, Paris, 1990.
26. Spidsoe, N., Summary report - - effect of surface elevation. SINTEF Report,
STF'71 A89027, Trondheim, 1990.
27. Spidsoe, N., Karunakaran, D. & Gudmestad, O., Nonlinear effects of
damping to dynamic amplification factors for drag-dominated offshore plat-
forms. In Proc. 11th Int. Conf. OMAE, Calgary, 1992.
28. Svano, G., Madshus, G. & Lango, H., On the validity of nonlinear spring
idealization of soil structure interaction. In Proc. 2nd European Conf. on
Structural Dynamics, Trondheim, 1993.

A P P E N D I X A - - T P G 500 J A C K - U P

Structural model

The jack-up platform considered is supported by three legs. This platform


is designed as a production facility to operate at a water depth of 110 m.
The top side mass of the structure is 16 800 tons. The legs are spaced 7 2 m
apart. Each leg consists of a triangular truss leg structure and supported
by p o d foundation which is piled into the sea bottom. The typical m e m b e r
diameters in the leg truss work range from 0.2 m to 0-8 m. A general view
of the platform is shown in Fig. A. 1.
A three dimensional F E M c o m p u t e r model is prepared. The leg truss
work i,; idealized as a string of beam elements with equivalent stiffness
properties. The c o m p u t e r model consists of 31 nodes and 33 elements. The
leg soil interaction is modelled by linear springs.
The first two natural periods of the structure are 5-3 s which are sway
98 N. Spidsoe, D. Karunakaran

J~

Fig. A.1. General view of TPG 500 Jack-up platform.

modes. The next mode is a torsional mode with a natural period of 4.3 s.
The other higher modes are found below 0-6 s,
The soil and structural damping together is assumed to be 2% and
modelled as proportional damping (Rayleigh damping) at the first and
third natural modes. The nonlinear hydrodynamic drag damping of the
structure is included by using relative velocities in the Morison equa-
tion.

Hydrodynamic coefficients

The hydrodynamic coefficients used in this analysis are:

drag coefficient CD: 1.0


mass coefficient CM: 2.0
Nonlinear dynamic behaviour of jack-up platforms 99

Environmental conditions

100 year significant wave height Hs = 15.5 m


Corresponding mean value for peak period Tp = 16.5 s
Wave spectrum JONSWAP
Storm duration 6h
Current velocities (corresponding to 10 year return period)
al: mwl 1.15 ms -~
- - 50 m 0.70 ms -1
- - 100m 0.60ms -1
al: m u d line 0.60 ms -~

A P P E N D I X B w MSC CJ62 JACK-UP P L A T F O R M

Structural model

The jack-up platform considered is supported by three legs. This platform


is designed as a production facility to operate at a water depth of 108 m.
The legs are spaced 62 m apart. The three legs, each consisting of a trian-
gular truss leg structure, rests on spud-can foundation. The typical
member diameters in the leg truss work range from 0.2 m to 0.8 m.
A three dimensional F E M computer model of the platform is prepared
for the analysis. A general view of the structure is shown in Fig. B. 1. The
leg truss work is idealized as a string of beam elements with equivalent
stiffness properties.
The: top side mass is lumped to the top four nodes. The structural weight
and the added mass of the legs are lumped at appropriate nodal points.
The first two natural periods of the structure are both 7.5s which
correspond to sway modes in orthogonal directions. The next mode is a
toi'sional mode with a natural period o f 5.7 s.

Hydrodynamic coefficients

The hydrodynamic coefficients used in this analysis are:


drag coefficient CD: 1-0
mass coefficient Cg: 2.0

~ a l comlitions

100 year significant wave height Hs = 14.8 m


Corresponding mean value for the peak period Tv = 16.0s
100 N. Spidsoe, D. Karunakaran

Fig. B.I. General view of MSC CJ62 Jack-up platform.

Wave spectrum JONSWAP


Storm duration 6h

C u r r e n t velocities ( c o r r e s p o n d i n g to a 10 y e a r period):
at m e a n w a t e r level 1.10 m s - 1
-- 30m 0-75 m s -1
-- 50m 0 - 7 0 m s -1
at m u d line 0.70 m s -~

S-ar putea să vă placă și