You are on page 1of 22

Marine Structures 9 (1996) 3-24

Elsevier Science Limited


Printed in Great Britain.
0951-8339196]$9.50
ELSEVIER 0951-8339(95)00001-1

Failure Probability of a Jack-up Under Environmental


Loading in the Central North Sea

J. W. van de Graaf, P. S. Tromans, L. Vanderschuren


Shell Research B.V., Koninklijke/Shell Exploration and Production Laboratory, Rijswijk,
The Netherlands
&
B. H. Jukui
Shell Internationale Petroleum ~laatschappij B.V., The Hague, The Netherlands
(Received 5 January 1995)

ABSTRACT

Ti~e annual probability of failure has been calculated for a jack-up subjec-
ted to extreme storm loading in the central North Sea. Statistics of extreme
loads were generated using models that account for the combined behaviour
of winds, waves and current in severe storms. The fluid loading is calculated
using well validated theories for the kinematics of extreme random waves
and currents and hydrodynamic models of jack-up legs. Non-linear push-
over analysis is used to evaluate the ultimate strength and the overturning
resistance of the jack-up unit. In the non-linear analysis, windward spudcan
sliding is included by means of a force constraint method. The likelihood of
structural failure or overturning is the probability that the environmental
load will exceed the overturning or structural resistance of the rig. Physical
uncertainty in the extreme environmental load dominates the problem and
is fully analysed. The effects of physical uncertainty in resistance of the
structure are discussed. The results indicate a failure rate of order 10-5 per
year.

Key words: long-term extreme load statistics, generic load models, non-
linear push-over analysis, reliability assessment, structural collapse risk.
© Shell Research B.V.
4 J. IV. van de Graaf et al.

INTRODUCTION

Existing site assessment practices for jack-up rigs are based on conven-
tional design methods for fixed steel structures, sometimes adopting a load
resistance factor design approach. These methods aim to make use of the
historical experience of the performance of fixed-jacket structures.
Though they provide criteria for acceptance or rejection, they do not
quantify safety under extreme environmental loading and inherently
contain an element of subjectivity. The desire for quantitative prediction
of the failure rate for ocean structures subjected to severe environments4
has been a continual driving force in the evolution of reliability analysis.~5
The present development (Refs 2, 12) recognises that the physical uncer-
tainties in environmental loading and structural strength should be the
major factor in the formulation of the reliability problem: modelling
uncertainties must be minimised such that they do not dominate. The
method employed here makes use of most realistic models for evaluating
extreme loads, extreme load statistics and structural collapse behaviour to
estimate a failure rate.
This paper describes a methodology for performing structural reliability
analysis and its application to estimate the failure rate of a jack-up
considered for location in a water depth of 91.4m in the central North
Sea. Conventional site assessment practices indicated that this rig would
be critically loaded under severe storm conditions.

OUTLINE OF THE CALCULATION OF STRUCTURAL


RELIABILITY

Figure 1 shows the steps in the calculation of structural reliability. The


procedure begins with the hindcast database of meteorological and ocea-
nographic ('met-ocean') conditions; the database contains the magnitudes
and directions of winds, waves and currents generated by numerical
simulation and matched to measurements. These data are used to produce
a representative combination of extreme environmental conditions,
required to generate a reference load set. In the push-over analysis, this
load set is applied to a finite-element model and increased in increments to
obtain the collapse sequence of the structure and its ultimate resistance.
The collapse analysis is repeated for several directions to produce a failure
surface for the structure. To calculate the probability that the loading will
fall outside the failtlre surface, it is necessary to return to the hindcast and
study the extreme response of a 'generic' structure, i.e. a simplified model
structure typical in configuration and behaviour of the kind under study.
Jack-up under environmental loading 5

I
Reprem~i;ve
O.n' o I Extreme
Load ModelI ~ "~ Condition

Repruentative .
Ioadset o

Pushover
Analysis

ReSistance

Strength Curve

Load or resistance
Probability of failure --~ I
I
Failure Surface

Fig. 1. Structural reliability assessment procedure.

The model of the generic structure may be developed empirically or


analytically (Refs 1, 13). It provides an algebraic relation between global
load and the environmental parameters (crest elevation, wave period,
current, wind speed, directions, etc.). The probability distribution of
extreme load in any direction on the generic structure can be deduced
from the hindcast accurately and efficiently using the methods described
in Ref. 13. The uncertainty in collapse strength can be taken into account
but, in view of the dominant uncertainty in environmental loading, it is
expected to be of secondary importance. 12 The probability of failure is
obtained by convolution of the cumulative directional distribution of
long-term extreme loads with the distribution of the ultimate capacity of
the structure.
The hindcast database of waves, currents and winds, the realistic wave
6 J.W. van de Graafet al.

load models, the generation of extreme long-term load statistics and the
push-over analysis are essential elements; without them a quantitative
reliability analysis is not feasible.

THE JACK-UP RIG

The jack-up, analysed, has an almost triangular shaped pontoon hull and
three triangular lattice structure legs with cross-bracing connecting the
three chords. It is equipped with permanent accommodation, a heli-deck
and a retractable cantilever carrying the drilling derrick. The elevating
system uses racks and pinions on each of the three chords. The chords
consist of a heavy rack section, sandwiched between two half tubular
members, cut along their length. The hull is raised and lowered by 36 jacks
arranged in pairs horizontally opposed across each chord. These jacks are
also used for holding the unit whilst preloading. In the elevated mode, and
also during towing, the fixation (rack-chock) systems are engaged. These
provide a rigid mechanical connection between the legs and the hull.
The analysis is performed for a location in the central North Sea, which
has a water depth of 91-4m. The spudcan-sea bed interaction has been
modelled in different ways, as described below. The met-ocean data is
taken from a grid-point of the NESS (North European Storm Study, see
Shaw7) hindcast typical of the central North Sea.

FINITE-ELEMENT MODEL OF THE JACK-UP

The basic geometry of the jack-up is provided in Table 1. The finite-


element model required by the USFOS program (Soreide et al. s) is shown
in Fig. 2. It consists of 201 nodes and 378 elements. Only six bays of the
portion of the legs which pass through the hull are modelled in detail. The
remaining section of the leg below the hull is r~odelled with an equivalent
tubular member to provide the appropriate lateral stiffness. The deck and
connections to guides and rack-chock are modelled with artificially stiff
and strong tubular elements. The guides between the deck and the leg
chords are modelled such that they provide only lateral restraint. The
lower guide is positioned halfway along a chord-span. Rack-chock stiff-
ness is included as a series of axial springs with the appropriate axial
stiffness.
The leg chords were modelled as tubular members. The dimensions were
chosen to match the cross-sectional area (i.e. the squash load of the chord)
and the weak-axis, plastic-section modulus of the actual chord. For the'
Jack-up under environmental loading 7

TABLE 1
Basic Geometry Data of the Jack-up

Water depth (chart datum) 91.4 m


Tidal rise 1.7 m
Storm surge 1.3 m
Finite-element model dimensions Elevation
Bottom of hull above sea bed 110.04 m
Elevation bottom guide above sea bed 117.74 m
Elevation rack-chock location above sea bed 121.41 m
Elevation top guide above sea bed 127.35 m
Distance between stern legs 47.60 m
Distance between stern legs and bow leg 45.75 m
Distance between chords of leg 11.89 m

Fig. 2. Finite element model of the jack-up rig.

stocky chord this ensures correct modelling o f the axial capacity (squash
load) and conservative modelling o f the (plastic) bending capacity. The
tubular braces o f the leg have been given their actual dimensions.
The: steel o f the chord has a nominal yield strength of 689 M P a and the
8 J. IV. van de Graaf et al.

braces of 448 MPa. Young's modulus (E) was assumed to be equal to


2.1 x 105 MPa. A reduced E-value has been used for the equivalent legs;
this is to account for shear deformation which is not included in the beam
element formulation in USFOS.
Two models were used for the foundations. In the first, the spudcans
were modelled as pinned, fixed in translation but free to rotate. Since this
may be a rather unconservative model if the spudcans have not penetrated
the sea bed a second model was also considered, in which the windward
spudcan is allowed to slide horizontally, subject to a frictional restraint
due to the sea bed.

LOADS

Gravity loads

All gravity loads acting on the unit have been applied as distributed point
loads. The total (gravity) load breakdown is a hull dead load = 80.28 M N
and a variable hull load = 33.41 M N (max). Leg-submerged weight,
spudcan weight and the effects of a 0.3% leg inclination are included in
the analysis.

Hydrodynamic loading

Hydrodynamic loads are applied on the legs, see Table 2. 'New Wave'
theory (Tromans et al. ll) has been used to calculate the hydrodynamic
loads on the legs of the jack-up rig. This theory describes the extreme
event occurring when dispersed wave energy focusses in space and time
such that component frequencies come into phase. It is by this process
that extreme waves arise in a real sea. There is much evidence supporting
the theory. It is completely consistent with the well validated random,
directional, time-domain wave simulation described by Rodenbusch 5 and
successfully predicts the measured loads on the Tern platform during the
passage of very high crests. 6
New Wave theory accounts for directional spreading of wave energy.
Spreading reduces the in-line wave kinematics and decreases the loading
on the structure. Data on spreading are provided in the oceanographic
database and have been taken into account in the calculation of the
extreme load statistics by applying a correction factor on the wave kine-
matics (see section on generic load model). Spreading leads to a small
uncertainty in the direction of the extreme load during a sea state.
However, this does not influence the reliability calculations since the long-
Jack-up under environmental loading

TABLE 2
H y d r o d y n a m i c L o a d D i s t r i b u t i o n Over Jack-up Legs

Wave attack Fraction at elevation Fraction at elevation Fraction at elevation


direction above mudline on above mudline on above mudline on
(degrees) leg 1 leg 2 leg 3
0 0.18 at 65.0 m 0.41 at 74.9 m 0.41 at 74.9 m
27.5 0.25 at 68.2 m 0.43 at 75.1 m 0.32 at 71.4 m
62.5 0.40 at 74-4 m 0.40 at 74.4 m 0.20 at 67.0 m
90.0 0-42 at 75-1 m 0.30 at 70.2 m 0.28 at 70-7 m
117.5 0.41 at 74.8 m 0-18 at 65.3 m 0-41 at 74.8 m
152.5 0.25 at 69.0 m 0.32 at 71.0 m 0.43 at 75.0 m
180.0 0.16 at 64.8 m 0.42 at 74-9 m 0-42 at 74-9 m

term statistics of extreme load are almost independent of direction within


the most severe loading sector.
The hydrodynamic model of the structure uses realistic values of drag
and inertia coefficients: a drag coefficient of around 1.2 for rough circular
section members and 0.65 for smooth ones, and an inertia coefficient of
1.5. For the chords the drag coefficient is directionally dependent and has
an average value of 1.8 based on the maximum dimension measured
normal to the rack. Together with New Wave kinematics and a one-
seventh power law current profile, this model yields the distribution of
wave and current loading over the legs of the jack-up required for the
push-over analysis. No account has been taken of any uncertainties in the
hydrodynamic load model or in the force coefficients relevant to extreme
conditions. These uncertainties are small in comparison with the uncer-
tainties in extreme environmental conditions.
The hydrodynamic model of the structure includes phasing effects in the
wave kinematics over the three legs. The hydrodynamic load distributions
over the three legs for the wave attack direction analysed are provided in
Table 12. It gives the fraction of the hydrodynamic load on the jack-up
acting on each leg and the corresponding elevation of application of the
resultant force on the leg. For determining the collapse strength of the rig,
a reference value of 10.0MN (total on the unit) hydrodynamic load has
been used in the push-over analyses. This value is close to that corre-
sponding with conventional site assessment practice.

Wind loading

Wind loads on the hull and the legs are applied at the hull location and are
calculated to have a magnitude of 30% of the hydrodynamic loading, that
is 3.0 MN.
10 J. IV. van de Graaf et al.

Structural dynamic amplification

Under the action of the time-dependent loads, compliant structures deflect


sufficiently for the mass inertia forces to amplify the structural responses.
For most jack-ups the mass inertia is predominantly associated with the
deck. In a static push-over analysis the dynamic amplification is simulated
by adding a load set which represents the mass inertia forces) This load
set was determined in a separate dynamic spectral analysis. A dynamic
amplification of 1.53 on hydrodynamic load is predicted for the 100-year
environment. This DAF is simulated by adding 5-3 MN lateral load at
deck elevation.
The wave and current load, 10.0 MN, and the wind load on the legs and
the hull, 3.0 MN, together with the mass inertial force of 5.3 M N on the
hull constitute the reference load set Sref = 18.3 MN. This value is repre-
sentative of the one used in conventional design practice.

CALCULATION OF THE FAILURE SURFACE

The failure surface was calculated for two spudcan support conditions:
first with all spudcans treated as pinned and secondly with the windward
spudcan allowed to slide along the sea bed.

Failure surface with pinned spudcans

Static push-over analyses were performed to obtain the failure surfaces of


the jack-up rig. In these analyses first gravity loads were applied, followed
by the environmental and mass inertia loading S; these were gradually
increased until the unit collapsed. Table 3 provides the load cases consid-
ered and the load factor 2 = S/Sref obtained for the various failure modes.
Also provided is the load factor 2100 = S/S~oo, which relates to the 'true'
100-year load Sl00 of 9.15 MN. the load S100 is obtained by considering the
'true' 100-year environmental load of 6.5 MN (see section on extreme load
statistics) with an unchanged structural dynamic amplification factor of
1.53 on the wave and current loading. The failure surface is presented in
Fig. 3 in terms of the components ;tx = Sx/Sloo and 2y = Sy/Sloo of the
load factor 2100.
The collapse of a jack-up rig can result from a number of mechanisms;
overturning, structural collapse of a leg, foundation failure or a combina-
tion of these. The capacity of the rig can also be limited by a localised
failure of the jacking mechanism and/or the rack-chock mechanism.
Which failure mechanism dominates depends on water depth and sea bed
Jack-up under environmental loading 11

TABLE 3
Load Factor at Ultimate Strength

Wave attack Variable X location Failure Load factor Load factor


direction hull load c.o.g. (m) mechanism 2 21oo
(degrees)
0.0 100% 0.0 Leg collapse 1.51 3.02
27.5 100% 0-0 Leg collapse 1.13 2.26
62.:5 50% 0-0 Overturning 0.89 1.78
62.:5 75% 0.0 Overturning 0.93 1.86
Leg collapse 1.02" 2-04
90.0 100% -3.9 Leg collapse 1.07 2.14
117.5 100% -3.9 Leg collapse 1.35 2.70
152.5 100% -3.9 Leg collapse 1.07" 2.14
180.0 50% -3-9 Overturning 0-86 1.72
180.0 75% -3.9 Overturning 0.92 1.84
Leg collapse 1.03" 2-06
152.5 50% -3.9 Windward start 0.75 1.50
of spudcan
sliding
180.0 50% -3.9 0.68 1-36
152.5 50% -3-9 Leg collapse 0.79 1.58
after windward
spudcan sliding
180.0 50% -3.9 0.76 1.52
Note: 0 degrees direction corresponds to bow leg leeward.
*These analyses were continued beyond the point of lifting of the windward leg, producing
artificial tension in the windoward leg during the analysis. This allowed the leg failure
surface to be obtained.

conditions, some analyses were continued beyond the point of lifting of


the windward leg, thereby producing artificial tension in the windward leg
during the analysis. This allowed the leg failure surface in Fig. 3 to be
obtained.

Overturning
For a jack-up unit with three legs there are three possible overturning
mode,;. In each mode environmental loading overturns the jack-up unit
about an axis passing through two of the three spudcans. Overturning is
initiated when the moment of environmental loading about the axis
exceeds the stabilising moment resulting from the weight of the jack-up.
The failure surface for jack-up overturning is shown in Fig. 3. In analysing
overturning, the most onerous combinations of variable hull load and
centre of gravity (c.o.g.) shift due to cantilevered drilling were assumed;
12 J. W. van de Graaf et al.

I ..... Overturning 100%variable load I


Overturning 50% variable load
Leg collapse
~'. 3.0

leg collapse following . • ....~.

(A) il
(B) ,; ..'2"" '°

'. - " -3.0

Fig. 3. Plan view (A) and structure failure surface (B).

these are the main causes of deviation from 'triangular' symmetry of the
failure surface. Push-over analyses showed that overturning would imme-
diately result in leg failure of the leeward legs just below the hull.

Structural collapse of a leg

A jack-up unit resists the lateral environmental loading as a portal frame.


The spudcan foundations have been conservatively modelled as pin join-
ted. As a result, the portal frame collapse mechanism involves structural
failure of the legs just below the hull. The load factor-hull deflection
curves for three characteristic wave attack directions, 0 °, 90 ° and 180° are
shown in Fig. 4. Leg collapse occurs through global leg bending, with one
or more chord-spans of the leg failing in compression or tension. The
failing chord-spans are always in the bay below that in which the lower
guide is positioned halfway. The "global leg hinges' which form behave
essentially as asymmetric cross-sections. When there are vertical loads the
hinges have more global bending capacity in one direction than in the
other (see Fig. 5). As a result the jack-up is significantly stronger in the 0
degree wave attack direction, when two chord spans on the bow leg fail in
compression, than in the other wave attack directions, when one chord-
span fails in compression on each stern leg. The failure surface for leg
collapse resulting from chord-span failure is also shown in Fig. 3.
The ultimate collapse strength (see Fig. 4) is achieved at failure of the first
Jack-up under environmental loading 13

'°F ..... :i
~ . 4-1 ................ 7..................F..................;.................i--;; i:r ................

. 4 ~ i i i ",~
1 2J i ., .....i~..................!..-..T..:...t
i t =i ..............
_! ^ ~ i'n,i
~.o-~ .................it~~,,: ............... i..................i//ri ................

il ~\ ;f l , i

Wave attack
0.4 ............./z..~.......\..i~O~i ...............
direction
i \ ~.i."~.I............... . . . . . 0.0 o
0.2 ..................... ~ : - ~ ,............~ i .................
............ .... 90 °
• 180 °
o Y i i.,,
-6 -4 -2 0 2 4 6
Deflection (m)

Fig. 4. Load factor - hull deflection curves.

P/PY j

, I.M/PYh

~%. -"v 1 ~ Rotation axis


/ t.rou0.
,oohor
s
Rotation axis / ~-'" J /
through one chord ~ /

PY = chOrd squash I°adl~ I I ""

Fig. 5. Moment - axial load interaction surface for jack-up leg with three chords.

chord: there is no system redundancy. Structural collapse immediately


follows failure o f the first leg. The load f a c t o r - deck deflection curves reflect
the brittle nature o f the structural collapse of this jack-up rig.

Failure surface with windward spudcan sliding

F o u n d a t i o n failure is site specific. F o r central N o r t h Sea bed conditions


punch-through o f the leeward spudcans was assumed not to be critical.
14 J. W. van de Graaf et al.

The influence of sliding of the windward spudcan was addressed for the
most critical wave attack directions: around the bow leg, windward.
Figure 6 shows the horizontal and vertical spudcan reactions of the
leeward and windward spudcans for 180° wave attack direction. (This
particular analysis does not model sliding and was continued beyond the
point of lifting of the windward leg, producing artificial tension in the
windward leg during the analysis, but allowing leg collapse to be
analysed.) The resistance to sliding was assumed to be equal to a lower
bound estimate of the friction between the bottom sediments and the
windward spudcan (valid for hard sand sediments). Hence, the limiting
value on the horizontal spudcan reaction H for the jack-up rig can then be
related to the vertical spudcan reaction V by the friction law:

H = (0.275 + 0.441/) MN

This equation is also shown in Fig. 6. To model sliding, the push-over


analysis is repeated with the horizontal restraint of the windward spudcan
removed and the horizontal spudcan reaction H is modelled as an applied
load. Before sliding, this spudcan reaction H increases nearly linearly with
applied lateral loading. After sliding, the spudcan reaction is proportional
to the vertical reaction V in the above equation, and decreases nearly
linearly with applied lateral loading. The procedure is schematically
shown in Fig. 7. The resulting horizontal and vertical spudcan reactions of
the leeward and windward spudcans are shown in Fig. 8. The corre-
sponding load factor deflection curves for both the hull and the sliding
spudcan are shown in Fig. 9. After the start of spudcan sliding, at a load

80

Z
:[
>

,,i i
2ol. ...............
"ki...~,~;_...ti..................
~;":~ ...... .~...!..................

> 0[-- ..~ .... !


"
i i t ......
....
Sliding surface
Windward spudcan
Leeward spudcan
-20
2 4 6 8 10 12
Hodzonlal spuclcan reaclion H MN

Fig. 6. Horizontal spudcan reaction H - vertical reaction V.


Jack-up under environmental loading 15

o<x<x, Post collapse behaviour


XaFw ~F. when ),aF w
----Im-

XFw XF. XF.


-----Im- -.---ira.

• ;LR= ~.R
, , . ) ~H';--HH , / . ' H

H .-------ID,. H ------~ H -----..~


Spudcan reactions Spudcan reactions Spudcan reactions
a) Ioadcase 1 b) Ioadcase 2 c) Ioadcase 3

F h = hydrodynamic load R = reaction force before R s = reaction force after


a = dynamic amplification factor sliding sliding
;Ls = Ioaclfactor at start of sliding ~f = Ioadfactor at collapse

Fig. 7. Simulation of spudcan sliding.

8O

Z o0 ....................................
! ..................
>

40 ........................................................................................ ". .................

",j i i i J
20- i \\i i.- -'~ '

o
>
0 f'°" ...... Sliding surface
i
I
.... Windward spudcan

-20
0 2 4 8 10 12
Leeward spudcan
I
Horizontal spudcan reaction H MN

Fig. 8. Horizontal spudcan reaction H - vertical reaction V.


16 J. IV. van de Graaf et al.

1.6

1.4 ......................................................................................................................

1.2 ....................................................................................................................

E ', i
= 1.0-
o
m t~

.................. ~..,p~..........~. . . . . . . . . . . . . . . . . . . ~. . . . . . . . . . . . . . . . . . , . . . . . . . . . . . . . . . . . . . ..................


~ 0.8- ~

!, , ,~!
o
~ 0.6-
,i'-N'I Hull
02 ii .............i...................i...................................................................... .... Windward spudcan
...... No sliding
0 n
-6 -4 -2 0 4 6
Deflection (m)

Fig. 9. Load factor - hull deflectioncurves.

factor 2 = 0.68, lateral load is shed from the sliding windward spudcan to
the leeward spudcans (see Figs 8 and 9). This causes increased leg bending
in the leeward legs just below the hull. Structural collapse occurs earlier,
for the 180° wave attack direction, at a load factor 2 = 0.76, instead of at
2 = 1.02, when the windward spudcan does not slide and overturning is
not permitted by allowing artificial tension in the windward spudcan to
occur. One facet of the failure surface for leg collapse following spudcan
sliding is also shown in Fig. 3. The non-linear analysis identifies 11%
reserve capacity beyond the start of spudcan sliding.

UNCERTAINTY IN COLLAPSE STRENGTH

The collapse strength has been treated as a deterministic quantity. It is


based on the 'as designed' condition of the jack-up; this ignores gross
imperfections due to damage or deterioration (e.g. fatigue, corrosion,
etc.). The uncertainty in system strength has not been included in the
reliability assessment. This is acceptable since the physical uncertainty in
extreme global fluid loading is the dominant source of uncertainty) 2
For structural members in offshore design, even when buckling, the
yield strength is the predominant source of variability in the strength of
the member. The jack-up rig uses high strength steel for the chords of the
legs. Mill test data 9 suggest similar values for bias (ratio between nominal
yield strength and actual mean value of the distribution of measured yield
strengths) and coefficients of variation (COV) for the distribution of yield
Jack-up under environmental loading 17

strength of high strength steel as for those of normal structural steel.


However, in the push-over analyses the nominal value for yield strength
has been used. This is effectively a lower bound on the true yield strength.
Material that does not fulfill the specified strength is generally rejected in
routine tests.
The evaluation of structural reliability ideally requires the use of the
actual distribution of strength. If the strength is modelled as deterministic,
an appropriate value should be used to account for the uncertainty in
strengtlh. Modelling a system, which has a distribution of strength with a
COV less than 10%, as behaving deterministically using the mean value is
only slightly unconservative in the estimation of reliability 12 and it
provides a good estimate of the failure strength. This simplifying assump-
tion of deterministic behaviour is easily outweighted in the present analy-
sis by the use of the nominal yield strength in the calculation of the failure
surface.

G E N E R I C LOAD MODEL A N D EXTREME LOAD STATISTICS

A generic load model of the jack-up rig is required to estimate long-term


extreme load statistics. This model is an algebraic relation between a
global load and the environmental variables. Both the base shear force F
and the overturning moment M on the structure may be described by an
expression of the form:

F, M = A1U 2 + A2UaTdp cos0c + A3~b2a2 + A4Ua2dpcos Oc/T


+ A5~p2a3/T 2 + A 6 W 2 c o s 0 w

where
a crest elevation;
T = zero crossing period of the sea state;
U = depth-integrated current reduced to account for current blockage;
W= wind speed sustained for 1 minute at 10 metres above mean sea
level;
OC = angle between mean wave and current directions;
Ow = angle between mean wave and wind directions;
correction factor on wave kinematics, accounting for directional
spreading, (see Ref. 13)

and Ai depends on the configuration of the structure, the attack direction


and weakly on the period of the waves - and on whether one wishes to
determine a force or a moment. The wave period dependence of Ai
18 J. IV. van de Graaf et al.

accounts for phasing effects in the wave kinematics. Current blockage


effects are taken into account in the reduced current speed U. l° These
expressions have been tested against numerical simulations of fluid load-
ing on more detailed models of the structure described in the previous
section on hydrodynamic loading.
The long-term statistics of extreme loads on the jack-up rig were
deduced from the NESS 7 met-ocean database using the methods described
by Tromans et al. 14 These authors have described the method in terms of
base shear force. However, the theoretical basis is valid for a wide range of
met-ocean and structural response variables.
The complete description of long-term load statistics in a particular
direction requires a combination of long- and short-term scale models.
The short-term statistics describe the probabilistic behaviour of the
extreme base shear force within a whole storm. These statistics are
completely defined by the most probable value of the extreme force, Fmp,
characterising the storm. The cumulative probability of an extreme value,
F, occurring in a storm characterised by a most probable extreme value,
Fmv, may be written e(FIFmp).
The long-term statistics describe the probability density p(Fmv) of
storms in terms of their Fmp values. The probability distribution of the
extreme F of any random storm (of unknown Fmp) is then
P(FIr.s.) = f P(FIFmp).p(Fmp).dFmp.
Storms arrive randomly at an average rate of v per year. Thus, the
probability distribution of the extreme base shear force F over some long
interval T, say 25 or 100 years, is
P(FIvT) = [P(FIr.s.)] vr for large vT.
P(FIFmp), p(Fmp) and v can be deduced from an extensive time series of
met-ocean variables such as NESS.
The 'true' 100-year return period loads generated on the jack-up at the
location in the central North Sea are found to be 5.0 M N for wave and
current loading from the stern direction. Combined with a corresponding
wind load of 1.5 MN, this results in a total 100-year return period load of
6.5MN. These values are significantly lower than those obtained by
conventional design practice (see Table 4). The difference arises mainly
from taking the non-coincidence of extremes of wind, wave and current
into account, and to a lesser extent from the combined use of better kine-
matic models and more realistic drag coefficients.
The extreme load on thejack-up rig at the location in the central North
Sea, normalised on the true 100-year return period value, is plotted as a
function of the return period in Fig. 10. It should be noted that the true
Jack-up under environmental loading 19

TABLE 4
Comparison of 100-year and Site Assessment Environ-
mental Loads (Wind, Wave and Current) for the Jack-
up Unit

Conventional site lO0-year return loads


assessment loads
Base shear 13.0 MN 6-5 MN
OTM 1083 MNm 604 MNm

2.0-

1.5-

1.0 I I I
102 103 104 10s
Retum pedod (years)

Fig. 10. Extreme load statistics for a central North sea location.

100-year return period environmental load used as a basis for Fig. 10 is


very different from the value used in conventional design practice (see
Table 4).

LONG-TERM RELIABILITY

The structural collapse of a jack-up rig involves overturning or leg failure.


Since the most critical failure mode is leg failure following windward leg
sliding and this is largely determined by base shear force on the jack-up,
total horizontal force is an appropriate global load variable to use in the
reliability calculations. The long-term reliability has been calculated using
the long-term distribution of environmental loading. The jack-up is
assumed to be positioned such that it has the worst heading with respect
20 J. W. van de Graaf et al.

to the extreme weather conditions. Figure 11 shows contour lines of the


cumulant of extreme long-term loads superposed on the failure surface for
overturning and leg collapse. The relevant portion of the failure surface
for leg collapse following sliding of the windward spudcan is also shown.
The prediction of the probability of failure begins with the calculation
of the probability of survival Po under extreme loading in a very narrow
sector centred on the wave attack direction 0. This is the probability that
the load will not exceed the value required to cause structural failure, i.e.
will not fall outside the failure surface, in that direction:
P0(survival) = Po(Lo < 2o'Sref)
where
Z 0 ~-environmental load in direction 0
20 = collapse load factor in direction 0
Sre = reference base shear force.
When the uncertainty in system strength is included in the calculation,
the above equation becomes a convolution of the probability density of
system collapse strength and the cumulative distribution of extreme load-
ing, both for the direction O.
Assuming independence of rare events, we can write the probability of
long-term survival under extreme loading as the product of the prob-

..... Overturning 100% vadable load


....... Overturning 50% variable load
Leg collapse

?'- 3.0"
Failure surface for ; "~."-~
leg Gollapee following ;
spudcansUdmg \ ~. / Q2.0~

..
," ,i"."/'7:- ~ 1.o-
i '#

..

10"s,10"4 per year


-3,o
Probability contours for
1041,10-7,10-8,
, ,, ,,
.; • '~,,, ,,
,,. ~i,' '~' •
~ . . . . ~ ' , ~ , "~%"-.

(B)
,,~.'~
,';'\ I
', ; \
', i
-,o

"
-1.0:
I

-2.o,
y O ">'~

i'.;'" -3.o!
Fig. 11. Failure surface with probability contours for exceeding extreme load.
Jack-up under environmental loading 21

TABLE $
Return Periods and Failure Rates for Jack-up in the Central North
Sea

Failure mode )~0min -- failure load Failurerate/year


lO0-year load
Overturning 1.78 1.8 E-6
Leg collapse 2.04 1.3 E-7
Failure due to sliding 1.52 2.5 E-5

abilities of surviving the extreme loading predicted for each separate wave
attack direction.
P (survival) = IlauoPo(Lo < 20"Sref)
where
Ilauo =: product over all wave directions 0 of the severe sector.
Finally, the probability of failure is obtained as:
P (failure) = 1 - P (survival).
The results, considering different failure modes, are shown in Table 5.
The critical failure modes of the jack-up unit are collapse of leeward legs
resulting from sliding of the windward leg or overturning when the sea bed
conditions do not permit sliding.

DISCUSSION

The only factors controlling the true failure rate are the inherent physical
uncertainties that reflect the variable nature of weather, the sea, structural
behaviour and steel strength. An essential requirement in a reliability
assessment is the use of proven models, minimising conservatisms and
modelling uncertainties inherent in a standard design procedure. The
methods applied here involve realistic, highly validated models for predict-
ing extreme loading and the calculation of system strength. Thus, though
some conservative elements are included, modelling ~uncertainty has been
s~'gnif'mantly reduced in order to approach the predietion.ofa ~true' failure
rate. Various authors 2' 12.15 have argued that the prediction of failure rates
has advanced beyond the notional level, and can now be used in compar-
isons with failure rates obtained from historical data for other events. For
the jack-up considered in this study, the risk of structural collapse under
extreme storm loading at the location considered is of the order of 10-5 .
22 J. IV. van de Graaf et al.

Although the intent of a reliability assessment is to use well proven or


calibrated models throughout, it has not been possible to avoid a number
of uncertain elements that have generally been treated conservatively.
Overall, the assessment should be viewed as conservative. Some of the
conservatisms are site dependent. Others require further work to evaluate
them properly.
The following points are worth noting:

- - structural strength has been calculated based on the designer's


information of the jack-up. Where information was incomplete, a
best estimate has been made based on the available data. No
attempt has been made to assess and quantify the effect of the rig's
condition, i.e. the structural strength is based on the 'as-designed'
condition;
-- the 'true' 100-year return period loads predicted are a factor of
two lower than the loads used for a conventional specific site
assessment. In the latter, individual extreme values for waves,
wind and current are combined for evaluating the loading.
However, these extreme values do not occur simultaneously. This
non-coincidence has been accounted for in the present reliability
study by calculating statistics of extreme load, rather than of
environmental variables. This has led to significantly lower
extreme loads. Directional spreading of wave energy also has a
considerable effect, leading here to a 22% reduction in total
hydrodynamic load, when compared with a uni-directional wave.
It is noted that use of the New Wave theory with the present
spreading produces loads about 10% below those obtained from a
conventional Stokes wave practice with the same environmental
conditions;
- - structural dynamic amplification has been included in the analysis.
Its magnitude has been evaluated using linearised spectral meth-
ods. 3 The use of this procedure to define an inertial load set is a
first pragmatic approach, which enables quasi-static procedures to
be maintained. A conservative aspect is that the push-over analy-
ses evaluate collapse behaviour using structural dynamic amplifi-
cation for 100-year return period conditions. However, the return
period for collapse is greater than 10000 years, for which the
dynamic amplification is significantly smaller. A more rigorous
assessment of the reliability of dynamically responding structures
requires further study;
-- rotational spudcan fixity has not been considered. The spudcans
(footings) have been modelled as pinned, but with no resistance to
Jack-up under environmental loading 23

rotation. For evaluating rig collapse with windward spudcan slid-


ing this is not unreasonable. For this water depth and in the limit
state of rig collapse, when sliding of the windward spudcan occurs,
the vertical loading on the leeward spudcans is so high that the
moment resistance is reduced to practically zero.

CONCLUSIONS

The evaluation of structural reliability as presented here depends on a few


essential elements:
(i) the availability of a hindcast database (or equivalent met-ocean
information);
(ii) accurate, proven and calibrated load models for predicting extreme
loading on the structure;
(iii) a method for calculating the cumulative distribution of long-term
extreme loads using a hindcast database;
(iv) the calculation of the collapse strength of thc structure.
While conventional reasscssmcnt procedures indicated that the jack-up
rig was critically loaded when subjected to extreme storm loading, the
present, more realistic reliability assessment shows it to bc reliable. For
this type of rig the probability of failure under long-term extreme loading
is of the order of 10-5 per annum.
The low failure rate predicted here for a jack-up rig that marginally
satisfied conventional site assessment criteria and practices gives confi-
dence in the general level of reliability implied by these practices. This
conclusion does not necessarily apply to other jack-ups, structures or
locations. Undoubtedly, this example demonstrates the value of reliability
methods. The predicted failure rates facilitate rational decision concerning
the use of certain jack-up rigs at specific locations and the discussion of
the safety levels embodied in offshore design codes. However, application
of the method demands an extensive met-ocean database, including joint
directional wave, wind and current data, and a non-linear structural
analysis package.

ACKNOWLEDGEMENTS

W e wish to thank Shell Expro, Aberdeen who initiatedand sponsored this


work. W e are particularly indebted to A.W. van Beck and M. Efthymiou
for their invaluable help and the information they provided.
24 J. w. van de Graafet al.

REFERENCES

1. Forristall, G. Z., Larrabee, R. D. & Mercier, R. S., Combined oceanographic


criteria for deep water structures in the Gulf of Mexico. Offshore Technology
Conference, 1991, OTC 6541.
2. van de Graaf, J. W. & Tromans, P. S., Statistical verification of predicted
loading and ultimate strength against observed storm damage for an offshore
structure. Offshore Technology Conference, 1991, OTC 6573.
3. Greeves, E. J., Jukui, B. H. & Sliggers, P. G. F., Representing jack-up
dynamic response and the inertial loadset technique. Fourth International
Jack-Up Conference, City University, London, 1993.
4. Marshall, P. W. & Bea, R. G., Failure modes of offshore platforms. Beha-
viour of Offshore Structures Conference (BOSS) Trondheim, 1976.
5. Rodenbusch, G., Random directional wave forces on template offshore
platforms. Offshore Technology Conference, 1986, OTC 5098.
6. Rozario, J. B., Tromans, P. S., Taylor, P. H. & Efthymiou, M., Comparison
of loads predicted using 'Newwave' and other wave models with measure-
ments on the Tern structure. Society of Underwater Technology Conference,
London, 1993.
7. Shaw, C. J., The North European storm study. Presented at The Impact of
Technical Developments on Safety Cases, London, March 1992.
8. Soreide, T. H., Amdahl, J., Granli, T. & Astrup, O. C., Collapse analysis of
framed offshore structures. Offshore Technology Conference, 1986, OTC
5302.
9. Tall, L. & Alpsten, G. A., On the scatter in yield strength and residual stres-
ses in steel members. Symposium on Concepts of Safety of Structures and
Methods of Design, London, 1969.
10. Taylor, P. H., Current blockage - reduced forces on offshore space frame
structures. Offshore Technology Conference, 1991, OTC 6519.
11. Tromans, P. S., Anaturk, A. R. & Hagemeijer, P., A new model for the
kinematics of large ocean waves - application as a design wave. ISOPE
Conference, Edinburgh, 1991.
12. Tromans, P. S., & van de Graaf, J. W., A substantiated risk assessment of a
jacket structure. Offshore Technology Conference, 1992, OTC 7075.
13. Tromans, P. S., van de Graaf, J. W., Efthymiou, M., Vanderschuren, L. &
Taylor, P. H., Extreme storm loading on fixed offshore platforms. Behaviour
of Offshore Structures Conference (BOSS), Vol. 1, London, 1992, 325-336.
14. Tromans, P. S., Hagemeijer, P. M. & Wassink, H. R., The statistics of the
extreme response of offshore structures. Ocean Engng, 19 (1992) 161-181.
15. Vugts, J. H. & Edwards, G. Offshore structural reliability assessment - - from
research to reality? Behaviour of Offshore Structures Conference (BOSS)
London, 1992.

You might also like