Sie sind auf Seite 1von 68

Health & Safety Executive

HSE

Sensitivity of jack-up reliability to wave-in-deck calculation

Prepared by MSL Engineering Ltd for the Health and Safety Executive 2003

RESEARCH REPORT 019

Health & Safety Executive

HSE

Sensitivity of jack-up reliability to wave-in-deck calculation

MSL Engineering Limited Platinum Blue House 1 st Floor 18 The Avenue Egham Surrey TW20 9AB

This document represents a study undertaken by MSL Engineering Limited for the Health and Safety Executive (HSE) to determine the effects of air gap erosion and subsequent hull inundation on a typical Jack-Up structure. A wave-in-deck loading model was developed (The MSL Model) and a sensitivity study was performed to assess the important factors affecting the loads generated during inundation. In addition, a typical Jack-Up structure was modelled using the USFOS structural analysis package in order to determine the dynamic response of a Jack-Up to storm loading where inundation occurs. The effects of response to previous waves, foundation modelling complexity and hull inundation levels were assessed by performing a number of short duration time-domain analyses. This report and the work it describes were funded by the HSE. Its contents, including any opinions and/or conclusions expressed, are those of the authors alone and do not necessarily reflect HSE policy.

HSE BOOKS

Crown copyright 2003 First published 2003 ISBN 0 7176 2177 4 All rights reserved. No part of this publication may be
reproduced, stored in a retrieval system, or transmitted in
any form or by any means (electronic, mechanical,
photocopying, recording or otherwise) without the prior
written permission of the copyright owner.
Applications for reproduction should be made in writing to:
Licensing Division, Her Majesty's Stationery Office,
St Clements House, 2-16 Colegate, Norwich NR3 1BQ
or by e-mail to hmsolicensing@cabinet-office.x.gsi.gov.uk

ii

FOREWORD

This document represents a study undertaken by MSL Engineering Limited for the Health and Safety Executive to determine the effects of air gap erosion and subsequent hull inundation on a typical Jack-Up structure. A wave-in-deck loading model was developed (The MSL Model) and a sensitivity study was performed to assess the important factors affecting the loads generated during inundation. In addition, a typical Jack-Up structure was modelled using the USFOS structural analysis package in order to determine the dynamic response of a Jack-Up to storm loading where inundation occurs. The effects of response to previous waves, foundation modelling complexity and hull inundation levels were assessed by performing a number of short duration time-domain analyses.

iii

iv

CONTENTS
Page No FOREWORD ...............................................................................................................III
CONTENTS ................................................................................................................V
1. 2. 3. SUMMARY..................................................................................................... 1
INTRODUCTION............................................................................................ 3
MSL WAVE-IN-DECK MODEL...................................................................... 5
3.1 Introduction ......................................................................................... 5
3.2 Derivation of Wave-in-Deck Loads........................................................ 5
3.3 Model Implementation .......................................................................... 6
3.4 Wave Aeration ..................................................................................... 7
3.5 Wave Spreading ................................................................................... 7
ENVIRONMENTAL DATA............................................................................. 9
4.1 Site Assessment.................................................................................... 9
4.2 10,000-Year Storm Conditions .............................................................. 9
4.3 Environmental Load Totals ...................................................................10
RESULTS OF WAVE-IN-DECK PARAMETRIC STUDY................................ 11
5.1 Typical Results..................................................................................... 11
5.2 Effect of Increasing Wave Inundation on Wave-in-Deck Loads ............... 11
Effect of Different Wave Theories on Wave-in-Deck Loads .....................12
5.3 5.4 Effect of Wave Aeration on Wave-in-Deck Loads ...................................13
DYNAMIC MODELLING ............................................................................... 15
6.1 Introduction ......................................................................................... 15
6.2 Structural Model................................................................................... 16
6.3 USFOS Hull Model .............................................................................. 18
6.4 Inclusion of Relative Velocity Terms ..................................................... 19
6.5 Foundation Modelling........................................................................... 19
6.6 Application of Pre-Wave Condit ion ....................................................... 21
RESULTS OF DYNAMIC ANALYSES............................................................ 23
7.1 Effect of Inclusion of Wave-in-Deck Loads on Structural Response,
Linear Foundations ............................................................................... 23
7.2 Effect of Modified Morisons Equation to include Relative Velocity ........ 23
7.3 Effect of Increasing Hull Inundation on Structural Response
(Linear Foundations)............................................................................. 24
7.4 Effect of Wave Theory on Structural Response....................................... 24
7.5 Effect of Response to Previous Wave on Extreme Wave Response .......... 24
7.6 Effect of Foundation Fixity on Structural Response ................................ 25
CONCLUSIONS.............................................................................................. 27
REFERENCES................................................................................................. 29

4.

5.

6.

7.

8. 9.

vi

1. SUMMARY

A wave -in-deck model (The MSL model) has been developed to assess the importance of the loads caused by possible wave inundation on the hulls of Jack-Up structures. The MSL method allows the use of higher order wave theories such as Stream Function and Stokes 5th Order theories, and allows an aeration profile for the wave to be inputted. A sensitivity study was performed using the MSL wave-in-deck model (in MathCAD format) to assess the importance of a number of variables on the wave-in-deck loads generated, such as degree of hull inundation, aeration level at the top of the wave and wave theory used. Dynamic pushover analyses were performed on a detailed non-linear model of a typical JackUp structure using the USFOS structural analysis package. The hull structure was modelled such that it attracted the wave-in-deck loads in accordance with the MSL model. The pushover analyses were based on a 10,000-year wave. A sensitivity study was performed to investigate the effects that hull inundation, foundation modelling, wave theory and structural response to a preceding wave have on dynamic response to the extreme wave. It has been found that wave-in-deck loads cause a significant increase in environmental loading once inundation occurs. The horizontal loading at the front face of the hull leads to a large increase in global overturning moment once inundation occurs. A key finding is that the upward wave-in-deck loading (principally caused by buoyancy of the hull) leads to a large decrease in vertical loading on the windward legs, with leg tension possible for inundations of 2m. The reduced vertical foundation load tends to decrease the moment capacity of the spudcan foundation, leading to greater leg/hull moments and global displacements.

2. INTRODUCTION

In April 2001, MSL presented the findings of a study entitled Assessment of the Effect of Wave-in-Deck Loads on a Typical Jack-Up (1,2) at the ISO North Sea Annex Committee Meeting and the SNAME Panel 4 Meeting. This study contained the following: 1. A wave-in-deck model, developed by MSL and referred to as the MSL model, implemented within an Excel Spreadsheet, based on linear wave theory. 2. Static pushover analyses for deck inundation levels ranging from 0.5m to 4m. The wavein-deck loads used in the pushover analyses (performed using the SACS structural analysis package (3)) were calculated using the software developed in item (1). The principal finding of this previous study was that the step change in environmental loading once inundation occurs is likely to have a large consequence on structural integrity. Leg lift off, due primarily to the large buoyancy loads predicted in the MSL model, was identified as a potential failure mechanism. The discussions that took place at the ISO and SNAME meetings, and subsequently at a meeting between HSE, Shell and MSL identified the following areas in which the previous study could be refined: 1. Wave Model One of the principal findings of the previous study was that the vertical loads (mainly buoyancy) generated during wave inundation are considerable. However, this conclusion was reached using Airy waves, which were used to allow the wave-in-deck model to be implemented within an Excel spreadsheet. It was agreed that the effect of higher order waves (for example Stream function and Stokes 5th Order) on wave-in-deck loads should be investigated. 2. Wave Aeration During the ISO/SNAME discussions, it was suggested that the top region of the extreme wave may be highly aerated, hence giving lower buoyancy and horizontal wave loads, and in turn a reduced tendency for leg lift-off to occur. 3. Water Particle Velocity It was suggested that the calculated velocity (from higher order wave theorie s) should be reduced to more accurately capture the wave-in-deck loads, and the sensitivity of the loads to this reduction should be investigated. 4. Foundation Modelling The model used in the previous study used linear foundations. It was suggested that the structure would be more accurately modelled using non-linear foundation springs. 5. Dynamic Response to Wave-in-deck loads The response of a Jack-Up unit to wave loading is comprised of a static and dynamic component. The dynamic component is normally significant as the natural period of the structure is often close to the period of the extreme wave. When a site assessment is

performed using methods recommended in SNAME 5 -5A (4), the usual approach is to apply a Dynamic Amplification Factor (DAF) to the wave loads to represent the worst likely inertia loading on the structure. However, the dynamic response of the structure to wave-in-deck loads may be different to the dynamic response to wave loads on the legs. There are three areas where differences may arise: a) the phasing of wave-in-deck loads may be different to the wave loads on the legs, b) the duration of the wave-in-deck loads will be shorter than the duration of the wave loads on the legs, and c) the vertical buoyancy loads may excite a higher order response mode than the global structural sway mode. The previous MSL study predicted that leg lift-off was a possible failure mode. This conclusion was reached assuming that the structure would react statically to wavein-deck loads, and that the DAFs for wave loading on the legs remain valid when used in conjunction with wave-in-deck loads. It was suggested that the dynamic response of a Jack-Up structure to wave-in-deck loads may more accurately be assessed by performing pushover analyses in the time domain. In order to incorporate the above items the present study has performed the following: 1. The existing Excel wave-in-deck model was converted to MathCAD format. The MathCAD program was updated so that:
rd Wave surface profile and velocity fields could b imported from a 3 party wave e processor. The wave processors used were SACS Seastate and ASAS-WAVE (5).

A fluid density profile with depth below the free surface may be defined, thereby accounting for wave aeration A reduction in horizontal particle velocity, dependent on depth below the free surface, may be defined. 2. The static model used in the first study (mounted in SACS) was converted for the USFOS analysis package. USFOS is a non-linear analysis program that has the capability of performing time-domain analyses, giving the dynamic response of the structure to the applied loads. The following refinements were made to the structural model: Use of non-linear pinions Use of non-linear foundations (including uncoupled and coupled springs) Wave-in-deck loads calculated within the structural model by using appropriate non structural elements such that the loads were accurately captured in the time domain Relative fluid velocities included in formulation of Morisons equation Sensitivity studies were performed on both the MathCAD wave-in-deck model and the time domain analyses to assess the relative importance of the items discussed above. The remainder of this report is structured as follows: Section 2 describes the MSL wave-indeck loading model and Section 3 details the environmental data used in the project. Section 4 contains the results of the sensitivity study performed on wave-in-deck load generation using the MSL wave-in-deck software. Section 5 contains details of the USFOS structural model used for dynamic simulations of wave-in-deck response, and Section 6 summarises the results of the dynamic analyses. Section 7 contains the conclusions reached from this study.

3. MSL WAVE-IN-DECK MODEL


3.1 Introduction A previously performed literature survey broadly fall within two categories, namely:
(6)

identified that existing wave-in-deck models

A global or silhouette approach, which provides global loads based on an exposed area of deck to the wave A detailed component approach, where loads on i dividual components are summed to n give a global force. The detailed component approach (e.g. Kaplan model (7)) is often used for topsides structures on fixed platforms, where the loads generated on individual members (e.g. beams, columns, items of equipment) may be calculated individually and combined to give a global force. In the case of the solid hull structures found on Jack-Ups, the global/silhouette approach is considered more suitable and has been adapted for use in the MSL wave-in-deck model. It may be noted that API RP 2A Section 17 (8) presents a simple method for predicting the global wave/current forces on platform decks. In the API model the load is evaluated based on calculating the drag force from a wave incident on the projected area of the deck. This method only produces front horizontal loads without explicit consideration of inertial, impact or pressure gradient effects. Wave phasing is also neglected. 3.2 Derivation of Wave-in-Deck Loads The MSL model considers both horizontal and vertical loads generated when deck inundation occurs, and is summarised in Figure 1. The horizontal load is derived assuming that the wave is effectively sliced off by the hull, with its momentum completely destroyed in the process. The vertical load includes the buoyancy of the hull as well as the hydrodynamic force induced by the vertical velocity of the wave particles as the wave passes under the base of the hull. Buoyancy loads are determined by calculating the weight of the displaced volume of water (i.e. the 'shaved off' region of the wave). Reference may be made to Figure 2 for the following derivations. Horizontal load: The horizontal load at the front face of the hull is derived from momentum considerations:

&u Fh = m

(1)

& where m = horizontal mass flow rate at the front face u = horizontal particle velocity including current ( = u + u c )
and hence it may be shown that by integrating the wave load over the wetted height of the front face of the hull that:
h x=0

Fh =

z hull

rb (u + u ) F(h
2 x=0 c

x= 0

- z hull)dz
(2)

where r = fluid density b = the width of the hull h = wave surface profile measured from seabed z = vertical distance from seabed F = Heaviside function; F(a) = 1 for a > 0, 0 otherwise Note that in the case of an aerated wave, the fluid density may be a function of vertical distance from the free surface. Vertical Hydrodynamic Load: The vertical hydrodynamic load on the underside of the hull may be deduced using similar considerations as above, and integrating the vertical load over the submerged area of the hull:

Fv = rbv 2F (h x= 0 - zhull )dx


0

(3)

where v = vertical particle velocity l = hull length in the direction of the wave Buoyancy Load: The buoyancy load is assumed to be equal to the weight of the volume of water displaced by the hull:

Fb = rgb(h - zhull )F(h - z hull )dx


0

(4)

Note that the position of the resultants for both vertical hydrodynamic load and buoyancy load are a function of time. 3.3 Model Implementation The original MSL wave-in-deck model utilised linear waves (i.e. Airy waves). In linear wave theory, the surface profile of the wave, horizontal velocities and vertical velocities may be described using relatively simple equations (9), and are more easily incorporated into a spreadsheet format. Higher order wave theories, such as Stokes 5th order theory and Stream Function waves are much more difficult to directly incorporate within a spreadsheet. Higher order wave theories were included in the updated MSL model b utilising existing y wave generation programs to create surface profiles and velocity fields. Two different programs were used to do this, depending on the wave theory used: (i) (ii) SACS SEASTATE: Used for Stokes 5th order and Stream Function theories ASAS-WAVE: Used for Shell NewWave

The output from these programs was then formatted so that it was in the form of a time history of surface elevations, and velocity fields for horizontal and vertical particle velocities. Increments of 10 degrees were used, giving a total of 36 time steps for the full wave cycle. Interpolation between the values generated by ASAS/SACS was used within the MathCAD program to enable smooth, integrable functions of surface profile with time and horizontal

and vertical velocity with time and depth. This allowed wave-in-deck loads to be calculated over the period of hull inundation. 3.4 Wave Aeration The degree of aeration at the top of the wave will have an effect on the wave -in-deck loads generated, acting to reduce buoyancy loads as well as reducing horizontal and vertical hydrodynamic loads. However, the instantaneous quasi-hydrodynamic load at a given point will only be proportional to the density of the water at that point. The very dramatic decreases in short duration wave slam loads due to an aerated wave front (10) is due to a large increase in water compressibility, and should not be confused with the longer duration quasi-hydrostatic loads being generated in the present situation. Most of the research that has been performed on wave aeration has been concentrated in the surf zone, where the highly turbulent wave action leads to high degrees of aeration (11). In open water situations, wave aeration is likely to be caused by high winds acting on the crest of the wave, as well as some local wave spilling. The aeration levels occurring in open water are likely to be much lower than those measured in the surf zone. In addition, the depth to which this aeration exists is likely to be small, with bubbles probably only penetrating the upper zone of the wave. In order to investigate the sensitivity wave-in-deck loads to aeration at the wave crest, aeration levels of 0, 10%, 20% and 30% at the free surface were used. The distribution of aeration with distance from the free surface was adapted from work presented by Bea (12). Aeration was assumed to act to a depth of one velocity head (using the horizontal crest particle velocity), defined with by following equation:

100gd k + 51 r = r unaerated (1 - k ) + log 2 uc2

(5)

For the extreme waves used in this study, one velocity head is approximately 2.5m. The aeration profiles derived from Equation 5 are shown in Figure 3. Values of the variable k were chosen to give the desired aeration at the free surface. 3.5 Wave Spreading In the SNAME 5-5A Recommended Practice, the reduction in wave loads caused by wave spreading is accounted for by reducing the deterministic wave height from 1.86Hs to 1.60Hs, where Hs is the significant wave height. This reduction may alternatively be modelled by applying a wave kinematics factor to the horizontal particle velocities of the wave. A wave kinematics factor of 0.84 was used in this study whilst keeping the deterministic wave height as 1.86Hs, hence ensuring that a realistic crest elevation was used in the wave-in-deck calculations and that the wave loads on the legs were reduced to a more representative value. The 0.84 factor was used when generating wave-in-deck loads as well as wave loads on the legs.

4. ENVIRONMENTAL DATA
4.1 Site Assessment In the original study (1) , a simple site assessment was performed on the Jack-Up to determine the maximum member utilisation using an environmental load partial safety factor of 1.15 and an LAT of 96.6m. The SACS structural analysis package was used to perform the assessment. It should be noted that the site assessment was limited to the extreme in-place condition only. The following conditions were applied to the structure for the site assessment, and are representative of conditions in the Central North Sea (Table 1):
Table 1 Environmental Conditions for Site Assessment LAT Hs Peak period Hmax Current 1-minute mean wind speed 96.6m 12.4m (50-year value) 15.3sec 23.1m (=1.86Hs) 0.67m/s 40.1m/s

The maximum member utilisation found using the above environmental conditions was 1.0, indicating that the Jack-Up, in the state analysed, is highly utilised. The site assessment was performed using a yield stress of 607N/mm2 . This was increased to 690 N/mm2 for the pushover analyses to represent a mean value. 4.2 10,000-Year Storm Conditions In the present study, the 10,000-year wave was used to generate wave-in-deck loads and perform structural analyses in the time domain. The 10,000-year wave height and period were used as the survival of such a wave is required if a target failure rate of 10-4 is to be achieved. However, the deck inundation levels used in this study are not directly related to a given return period, and have not been determined from site-specific conditions. The 10,000-year wave used is as follows (Table 2):
Table 2 10,000-year Environmental Conditions Hmax Tmax Current 1-minute mean wind speed 31.2m 17.7sec 0.67m/s Note: 50-year values used 40.1m/s Values determined using SNAME 5-5A methodology

It may be noted that individual 10,000-year extreme values of current and wind were not applied in conjunction with the 10,000-year wave conditions.

4.3 Environmental Load Totals The following load totals (see Table 3) resulting from the applied 10,000-year environmental conditions were determined from the original SACS model, and were subsequently replicated using the USFOS model:
Table 3 Environmental Loads (excluding wave -in-deck loads) Item Maximum Base Shear Maximum Overturning Moment Wind Base Shear Wind Overturning Moment Maximum Value 29MN 1590MNm 3.2MN 430MNm Time of Occurrence * +0.4s +0.8s -

* time of zero seconds refers to the moment at which the top of the wave is in-line with the face of the hull

10

5. RESULTS OF WAVE-IN-DECK PARAMETRIC STUDY


5.1 Typical Results A graphical representation of the wave passing through the structure may be found in Figure 4. Figure 5 shows typical results from the MathCAD wave-in-deck software. In this case, a stream function wave was used with a deck inundation of 3m. The following points may be made from Figures 4 and 5: 1. The duration of the horizontal wave-in-deck load is around 2.5 seconds, and is at a maximum when the wave crest is in line with the front face of the hull (t=0 seconds). 2. The peak horizontal wave-in-deck load is 135MN. This is slightly out of phase with the peak base shear from wave loading on the legs (the phase difference is around 0.5 seconds). 3. For an inundation of 3m, the maximum horizontal wave-in-deck load makes up around 30% of the total base shear at t=0. 4. The vertical loading on the hull is dominated by buoyancy loads. The peak buoyancy of 71MN occurs approximately 1 second after the peak horizontal wave-in-deck load. 5. The duration of the vertical loading is around 4 seconds for an inundation of 3m. 6. The maximum wave-in-deck overturning moment (which includes horizontal and vertical loading effects) occurs at t=0 seconds. 7. The overturning wave-in-deck moment becomes negative at around 1.25 seconds after the time of peak horizontal wave-in-deck loading. At this point the centre of buoyancy has passed the plan centroid of the three legs, and the horizontal load is almost zero. 8. The maximum wave-in-deck overturning moment makes up around 55% of the total wave load. Note that the overturning moment was defined as the moment about a horizontal line at the mudline perpendicular to the direction of the applied wave and passing through the plan centroid of the 3 legs. Figure 6 shows base shear and overturning moment time-histories for wave and current on the legs (with no wave-in-deck loads) and wave and current on the structure (including wave-in-deck loads). 5.2 Effect of Increasing Wave Inundation on Wave-in-Deck Loads The effect of increasing wave inundation on the wave-in-deck loads generated (using a Stream Function wave) are shown in Figures 7 to 10. Figure 7 shows horizontal loads for increasing inundation levels. The magnitude of the maximum horizontal wave-in-deck load is approximately proportional to the maximum inundation level, as the velocity profile is roughly constant at the very crest of the wave. The duration of loading increases from 1.4 seconds (1m inundation) to 2.8 seconds (4m inundation). Figure 8 shows the buoyancy loads generated for increasing inundation levels. The maximum buoyancy loads generated dur ing inundation increases from 14MN for 1m inundation (0.5

11

seconds after the peak wave-in-deck base shear) to 102MN for 4m inundation (occurring 1.1 seconds after the peak wave -in-deck base shear). The buoyancy values compare with a total dead load of 161MN for the structure. As the wave passes through the Jack-Up, the centre of action of the buoyancy load moves from providing a positive overturning moment on the hull to a negative overturning moment (Figure 9). Figure 10 shows the effect of inundation leve l on total wave-in-deck overturning moment. The maximum total overturning moment increases from 717MNm (1m inundation) to 2980MNm (4m inundation). The duration of the loading also increases as the maximum inundation increases, with the positive phase of the overturning moment increasing from around 1.8 seconds (1m inundation) to 2.7 seconds (4m inundation). The time period between maximum positive and negative overturning moment is relatively insensitive to inundation level, being primarily a function of wave celerity. 5.3 Effect of Different Wave Theories on Wave-in-Deck Loads As stated previously, the MSL wave-in-deck model was updated so that higher order waves, th such as Stream Function and Stokes 5 Order theories could be incorporated. The surface profile of the wave will affect the magnitude, duration and phasing of buoyancy loads. The horizontal velocities at the top of the wave will directly affect the horizontal loads generated. Figure 11 shows a comparison of wave surface profiles. It can be seen that the Airy wave gives a less peaked profile than the other waves, and hence would be expected to give greater wave-in-deck loading durations. In addition, for a given inundation level, a larger volume of water will be displaced by the hull during the inundation process, and hence the hull will attract larger buoyancy loads for Airy waves. It should be noted that the still water level of the waves were adjusted to give the same extreme crest elevation. The Stream function and Stokes 5th order theory produce very similar profiles. The shape of the NewWave profile is similar to that of the Stream/Stokes 5th profiles at the crest of the wave, and would therefore be expected to provide similar load durations and buoyancy levels. Figure 12 shows the distribution of horizontal particle velocity with elevation for the four wave theories assessed. It may be seen that the horizontal velocities predicted by linear wave theory are between 5% and 10% less than those predicted by both Stokes 5th order and Stream function (this agrees with work presented by Barltrop (9)). The NewWave formulation gives somewhat higher horizontal velocities near the surface of the wave. Figure 13 shows the effect of using different wave theories on wave-in-deck loads. The Airy wave underestimates the horizontal wave-in-deck loading compare to the higher order wave theories by around 12%, as a result of the lower particle velocities predicted. However, the Airy wave overestimates buoyancy loading on the hull by around 20% due to its sinusoidal surface profile. It may be seen from the bottom graph of Figure 13 that the two effects described above have conflicting contributions to overturning moment, with Airy waves being a reasonably good approximation of combined overturning moment for inundations up to 2m. It may be noted that the wave-in-deck loads determined are relatively insensitive to the method of current stretching adopted. Horizontal wave-in-deck loads are dependent on the horizontal particle velocities at the top of the wave, and hence the stretching profile used to extrapolate surface velocities down from the crest to the mean water level will not to have a significant effect on calculated load for inundation levels up to 5m.

12

5.4 Effect of Wave Aeration on Wave-in-Deck Loads Figure 14 shows the effect of aeration at the wave crest on wave-in-deck loads. The aeration profile described in Equation 5 was used. It may be seen that a relatively high surface aeration of 30% reduces the wave-in-deck buoyancy, base shear and overturning moment by around 20%.

13

14

6. DYNAMIC MODELLING
6.1 Introduction The dynamic response of a Jack-Up unit to wave loading, even when no inundation of the hull occurs, is complex. The effects of random waves, dynamic amplification, drag dominated (i.e. non-linear) loading, non-linear structural response, global P-delta moment and relative fluid velocities mean that the calculation of extreme structural response must be performed with care. The step change in loading that occurs with hull inundation makes the response even more complex, with the nature of the true response being far from obvious. Where no inundation occurs, the dynamic amplification of static response to extreme loading may be assessed by following the methods described within SNAME 5 -5A. A simple and commonly used method is to determine a DAF based on a Single Degree of Freedom (SDOF) model. The DAF is calculated by comparing the global sway period of the structure with the period of the extreme wave. A dynamic loadset is then applied statically to the model to represent this inertia loading. The SDOF method assumes that structural response is dependent on one loading regime with a given period, and on one modal response to that loading. However, when inundation of the hull occurs, the loading and response regime become much more complex. The items listed below will affect the true extreme response: (i) The duration of horizontal wave-in-deck loads: It was seen in Section 5.1 that the duration of horizontal wave loads is generally in the region of 1 to 3 seconds. This compares with a natural period in the sway mode of between 5 seconds (fixed foundations) and 12 seconds (pinned foundations) for the present structure. It is likely that the dynamic amplification factor for this type of loading alone will be less than 1.0 for pinned foundations, but may be greater than 1.0 for stiff foundations. (ii) The duration of vertical wave-in-deck loads: The duration of the vertical wave loads (mainly buoyancy loads) is in the region of 1 to 4 seconds. However, the period of the structural mode likely to be excited by these loads is less than 2.0 seconds for this structure (including vertical foundation flexibility), suggesting that the structure may react statically to vertical loading. (iii) Load phasing It was seen in Section 5.1 that the greatest wave-in-deck overturning moment generally occurs at the time when the top of the wave is in line with the front face of the hull. This is in the region of 0.5 seconds before the peak wave-on-leg loading occurs. In addition, the wave-in-deck overturning moment becomes negative at a time when the wave load on the legs is near or at its maximum value. This negative overturning moment is likely to have a moderating effect on the maximum horizontal structural response. (iv) Relative particle velocity: The predicted horizontal displacement of the Jack-Up under extreme load conditions is considerable, with displacements in the order of 3 -5m being possible. For example, a wave period of 17.7 seconds causing a 5m displacement gives a peak horizontal velocity of around 1.8 m/s (assuming sinusoidal response). Given that the particle velocity at the crest is around 7 m/s, the incorporation of relative velocities will have a significant effect for extreme wave conditions.

15

(v) Foundation conditions: The degree of foundation fixity assumed at the spudcan has a profound effect on the predicted structural response of a Jack-Up unit. Increased foundation fixity leads to lower bending moments at the leg/hull connection as well as a reduced sway natural period. A reduction in the natural period will reduce the DAF for the wave loads on the legs and probably increase the DAF for horizontal wave-in-deck loads. The inclusion of uncoupled non-linear foundations (i.e. separate non-linear springs in the horizontal, vertical and rotational directions) makes the structural response more complex, as the bending moment distribution down the legs and the natural period of the structure will vary during the wave cycle. Perhaps the best method for modelling foundation fixity during a wave cycle is to apply coupled non-linear springs to the base of the legs. This means, for example, that the rotational fixity provided by the spudcan is dependent on the vertical load applied to it at that time. The USFOS analysis package (13) contains a SPUDCAN element to specifically include these effects, whose formulation is based on work by Van Langen (14). (vi) Initial conditions prior to the extreme wave: The structural displacements, velocities and accelerations prior to the extreme wave loading on the structure will significantly affect the response to the extreme wave itself. The most rigorous analysis method to assess their effects would be to run a number of 3 hour simulations of the storm conditions causing the 10,000-year wave in the time domain. Statistical methods (such as those detailed in SNAME 5-5A) could then be used to determine the worst likely response in a 3-hour period. It should be noted that it is this extreme response that a dynamic amplification factor applied to quasi-static wave loads is meant to represent. Such lengthy time history analyses is beyond the scope of the present study, even if pertinent data for such storms existed. However, the importance of the initial structural conditions prior to the main wave may be made by applying a pre wave to the structure, with the main wave occurring at some time later. The pre-wave is applied to give the structure a set of initial displacements, velocities and accelerations. This approach is described in more detail in Section 6.6. A detailed structural model of a typical Jack-Up was created in USFOS by converting and updating the SACS model used in the original MSL wave-in-deck study. The model was created to allow the investigation of the above effects and to model dynamic structural response to wave-in-deck loading as accurately as possible. 6.2 Structural Model 6.2.1 Introduction

The modelled Jack-Up is a 3-legged structure with a maximum elevated weight of 12500t (Figure 15). The total weight of the modelled platform is 16500t. The legs are triangular, consisting of 3 tubular chords spaced at 12.2m and braced at bays of 6.96m with a K-bracing arrangement. The combined weight of the legs and the spudcans (excluding buoyancy) is 1275t. At the bottom of each leg is a spudcan, with a plated tank section immediately above extending to a height of 23m. The overall height of each leg is 146m, and are spaced approximately 55m apart. The connection between the leg and hull consists of a set of pinions and rigid horizontal guides at the bottom of the hull and at the top of the yoke frame. In plan, the hull structure is approximately triangular, with a side length of approximately 72m (Figure 2). The four pyramid structures shown in Figure 2 are used only to apply the vertical topsides loads to give the correct vertical centre of gravity.

16

The following comparisons of key results were performed to ensure the conversion from the original SACS model to USFOS format was successful: Dead loading Base shear and overturning moment for the extreme wave Mode shapes and their associated periods Static pushover response.

A comparison of the results from the two models is shown in Table 4. Static pushover analyses were performed on both models. Ultimate structural capacity and global force displacement curves were found to agree closely.
Table 4 Comparison of SACS and USFOS finite element models Results for: Dead weight Base Shear Overturning moment Sway natural period SACS 16500t 29.5MN 1660MNm 11.1 seconds USFOS 16500t 29.6MN 1720MNm 11.4 seconds

USFOS is a dedicated non-linear, dynamic analysis package, specifically written for the offshore industry. The following enhancements to the original SACS model were made in converting to USFOS: Use of uncoupled and coupled non-linear foundations Modelling of non-linear pinions, allowing load redistribution at the leg/hull connection Analyses performed in the time domain Relative fluid particle velocities included in Morisons equation. Leg modelling

6.2.2

All major structural members in the legs were explicitly modelled. The leg model was sufficiently detailed to ensure representative loads resulting from wave and current would be generated. The main chord members of a Jack-Up may have different section properties across the major and minor axes, caused by rack teeth and any additional strengthening to the tubular that may be present. For the purposes of the present study, the chord members were modelled as equivalent tubulars, whose diameter and wall thickness were chosen to give the average plastic modulus and the cross-sectional area of the combined section. It is these properties that have the most effect on the global structural stiffness and the plastic axial and bending capacity of the legs. The tank plating at the base of each leg was represented by truss members to give a suitable stiffness. Brace offsets were not modelled. The hydrodynamic loads for rack post members were generated assuming the original chord diameter of 850mm and modified values of Cd and Cm based on recommendations contained in SNAME 5-5A. Averaged values of Cd for the three chords in each leg were used for simplicity. The following values of Cd and Cm were used (Table 5).

17

Table 5 Hydrodynamic Coefficients Applied to Chord Members Element Rough rack post Smooth rack post Rough brace member Smooth brace member Tank region at base of leg Cd 1.66 1.45 1.0 0.65 8.9* Cm 1.8 2.0 1.8 2.0 89.1*

* coefficients applied to individual members of 850mm diameter

6.2.2

Leg/Hull Connection

The guides and pinions forming the connection between the legs and the hull were represented by spring elements within USFOS. The guides and the pinions were attached to the yoke frame, which was modelled to give a representative structural stiffness. The following guide and pinion stiffnesses were modelled (Table 6):
Table 6 Guide and Pinion Stiffnesses Element Upper guide Lower guide Pinions Stiffness 86.0 x 103 t/m 500 x 103 t/m 75.0 x 103 t/m (vertical) 35.1 x 103 t/m (horizontal)

The gaps that exist between the guides and the chord members and in the pinion mechanism were not modelled in this study. The pinion members were modelled as bi-linear springs. Pinion slippage (yielding) was assumed to occur at an ultimate capacity of 12.3MN, with additional displacement assumed to give no extra load capacity. It may be noted that during the short duration time-domain analyses performed in this study, although some pinion slippage was noted, the displacements of the pinion springs were not excessive. 6.3 USFOS Hull Model The hull was modelled using plate elements. The plate thicknesses were chosen to reflect the true stiffness of the bulkhead stiffnesses including local stiffeners. Plate thicknesses and stiffener geometries were taken directly from available drawings of the jack-up structure. The plate elements were assumed to behave elastically under all applied loads. Non-structural elements were added to the plated hull model to allow the automatic generation of wave-in-deck loads during time-domain analyses, as illustrated in Figure 16. Two sets of elements were defined: (i) Buoyancy elements: These were defined as vertical members extending from the base to the top of the hull. Each member was partitioned into 99 sections to allow the accurate calculation of hull buoyancy as the wave crest passed through the hull. A cross-sectional area was assigned

18

to each member appropriate to the hull plan area around it. The sum of these member areas equalled the plan area of the hull, taking due account of the voids in the hull around the le g sections. The dummy members were assigned zero drag and inertia coefficients. The areas of the members were then increased by 10% to account for vertical hydrodynamic loads predicted by the MathCAD model. (ii) Drag elements: Additional vertical dummy elements were added to the side of the hull facing the incoming waves. Each member extended over the full height of the hull. Appropriate values of the product of Cd and member diameter were chosen to give a suitable value of effective frontal area. Member buoyancy and the inertia coefficients were set to zero. The USFOS wave -in-deck model was validated by applying a Stokes 5th order wave statically to the USFOS model and comparing the reported loads with those predicted from the MathCAD analyses reported in Section 5. A comparison of the two methods is shown in Figure 17. There is excellent agreement between the two methods (both in terms of the magnitude of the loads generated and the relative phasing of vertical and horizontal load components), with no need for correction factors to be applied. It should be noted that for the validation, the hull leg was prevented from moving vertically and horizontally, hence removing the effects of relative fluid velocities and displacements. 6.4 Inclusion of Relative Velocity Terms The potentially large structural displacements of a Jack-Up in extreme storm loadings suggests that the velocity of the structure during the wave cycle may be significant when compared to the wave particle velocities. Morisons equation may be extended to include a relative velocity term:

& & Fdrag = 0.5 rCD D un - rn (un - rn )

(6)

where u n is the combined particle velocity from wave and current normal to the member considered & rn is the velocity of the considered member normal to its axis in the direction of the combined particle velocity The inclusion of a relative velocity term introduces a fluid damping effect. Relative velocities were included within the load formulation in the USFOS model except where noted. 6.5 Foundation Modelling 6.5.1 Linear Foundations

Three sets of linear foundations were used in the present study (Table 7):

19

Table 7 Linear spring stiffnesses Spring stiffness Rotational (kNm/rad) SNAME 5-5A extreme Fatigue spring SPUDCAN* initial stiffness 4.96 x 106 124 x 106 20 x 106 Horizontal (kN/m) 75.5 x 103 579 x 103 316 x 103 Vertical (kN/m) 130 x 103 2.48 x 106 412 x 103

* The SPUDCAN linear springs were derived using the USFOS in-built spudcan model. Section 6.5.2 below contains further details of the spring derivation.

The SNAME 5-5A extreme fixity was determined using the SNAME recipe. The fatigue spring fixity is close to the measured values inferred from measured natural periods of a similar Jack-Up (15). 6.5.2 Coupled Non-Linear Springs

A SPUDCAN element is included within USFOS which was specifically developed to represent the foundation fixity conditions provided by preloaded spudcans. The element is derived from the stiffness formulation presented by van Langen, which is consistent with SNAME 5-5A Recommended Practice and is based on hardening plasticity theory. The USFOS SPUDCAN element creates a set of coupled springs that represent foundation stiffnesses in the horizontal, vertical and rotational directions. The rotational and horizontal stiffness are dependent on the vertical load on the element at that time, and hence the foundation stiffness is modified continuously throughout the dynamic analyses. The parameters used for the SPUDCAN element are shown in Table 8, and are typical of data in the Central North Sea.
Table 8 Input Parameters for USFOS spudcan element Soil type Effective spudcan diameter Undrained shear strength Shear modulus Embedment depth Preload Clay 16.19m 50kN/m2 5.39 x 103 kN/m2 5m 175 x 103 kN*

* artificially high value, chosen to allow the analysis to continue beyond the 50-year preload value of 101.3 x 103 kN.

The spudcan element provides a linear section to the load-displacement relationship, beyond which a plastic displacement region exists. Typical load-displacement curves (for horizontal and moment loading) are shown in Figures 18 and 19. The relationship between vertical load and moment capacity (both ultimate moment capacity and the moment at first yield) and vertical load and sliding capacity are shown in Figures 20 and 21. It may be seen that the ultimate moment capacity is broadly in line with the average ultimate moment capacity determined by Noble Denton (16) following the SNAME 5 -5A 20

methodology using similar soil parameters. The moment capacity reduces from a maximum of 285MNm at a vertical load of 100MN to zero when no vertical load is applied. The sliding capacity reduces from a maximum of 12.3MN at a vertical load of 100MN to zero when no vertical load is applied. Table 7 above shows the initial foundation stiffnesses derived from the USFOS model. Comparison with the extreme and fatigue stiffnesses derived by Noble Denton show good agreement. As was noted in Table 8, an artificially high value of preload was applied to the SPUDCAN element (175MN instead of around 100MN). This was required as the element is only formulated up to the preload vertical load, with larger loads causing the analysis to stop. The increased preload gave larger moment capacities than would otherwise have been predicted. 6.5.3 Non-Linear Foundations

In order to investigate the importance of the coupling effect of the foundation springs (see Section 6.5.2) on structural response, and to determine whether uncoupled non-linear springs would give simila r results, non-linear springs were developed based on the foundation stiffnesses derived from the USFOS SPUDCAN element. The force-displacement curves were based on the results from the SPUDCAN element at a vertical load of 80MN. The linear springs were based on the initial stiffness of the SPUDCAN element. 6.6 Application of Pre-Wave Condition In order to model the possible contribution of the structural response to a preceding wave on the response to the main wave, the following methodology was used: 1. Apply 50-year deterministic wave statically to the structure to create a wave loadset 2. In the dynamic model, apply the 50-year loadset gradually to the structure 3. Suddenly release the load, which then makes the structure oscillate at its natural (i.e. sway) period 4. Apply the 10,000-year wave at a time Dt after releasing the 50-year load. By varying the value of Dt over a range equal to the sway natural period of the structure the full envelope of maximum response to the extreme wave may be determined.

21

22

7. RESULTS OF DYNAMIC ANALYSES


7.1 Effect of Inclusion of Wave-in-Deck Loads on Structural Response, Linear Foundations Three analyses were performed to assess the significance of the following wave-in-deck loads for global response:
Run number 1 2 3 Vertical wave-in-deck loads x x Horizontal wave-in-deck loads x

The above runs were performed with an inundation of 3m and linear extreme foundation springs. The horizontal displacements at deck level are shown in Figure 22. It can be seen that the inclusion of wave-in-deck loads has slightly increased the maximum horizontal displacement. It may also be seen that, in this case, the inclusion of buoyancy loads has acted to reduce the maximum displacement as compared to the wave-in-deck model with no buoyancy taken into account. The buoyancy reduces structural displacement (for the linear foundation case) for two reasons: firstly, the large upward load acts to stiffen the structure, and secondly, the buoyancy creates a negative overturning moment at the same time as the peak wave-on-legs base shear. The effect of including wave-in-deck loads on vertical foundation reactions is shown in Figure 23. For the windward leg, the inclusion of wave-in-deck loads, including buoyancy, reduced the vertical reaction significantly between the times of 6 and 10 seconds (i.e. during the time when the hull is inundated). The decrease in vertical load on the two windward legs is approximately equal to the buoyancy predicted by the MSL MathCAD model (69MN compared to 71MN predicted in Section 5) and hence indicates that the structure reacts statically to vertical loading. The large decrease in vertical load does not occur for the wavein-deck model with buoyancy effects ignored, although there is a slight decrease in loading due to the increased overturning moment caused by the horizontal wave-in-deck component. In the case of the leeward leg, it may be seen that the wave-in-deck buoyancy has a moderating effect on the maximum vertical reaction, especially when the wave has passed the centroid of the hull and is hence causing a negative overturning moment. The wave-in-deck case with buoyancy loads ignored leads to an overestimate of vertical loading on the leeward leg, as the additional overturning moment from horizontal wave-in-deck loads is included but the relieving moment due to eccentric buoyancy is not. 7.2 Effect of Modified Morisons Equation to include Relative Velocity Figures 24 and 25 shows the effect of ignoring relative velocity terms on structural response, using linear extreme foundation springs and an inundation of 3m. The inclusion of relative velocity effectively introduces a damping term into the response. It should be noted that the reductions in maximum response and maximum/minimum leg reactions may not be as marked for cases where stiffer foundation conditions are assumed, as the structural velocities will be smaller.

23

7.3 Effect of Increasing Hull Inundation on Structural Response (Linear Foundations) The effect of increasing hull inundation on horizontal structural displacement is shown in Figure 26. Inundation levels up to 3m were investigated. SNAME 5-5A extreme fixity levels (Section 6.5) were used. No pre-wave was applied. It may be seen that the amount of hull inundation does not greatly affect the horizontal displacement of the structure. Although the horizontal wave-in-deck loads are significant (approximately 30% of the total base shear), they are slightly out of phase with the wave loading on the legs. In addition, the presence of large upward vertical loads due to hull buoyancy acts to stiffen the structure in the global sway mode, and the negative overturning moment identified in Section 5.2 also reduces the extreme response. The effect of increasing hull inundation on the vertical reaction on the leeward leg is shown in Figure 27. Increasing inundation has little effect on the maximum reaction, as at the time at which it occurs (t=10 seconds) the hull is no longer inundated. Figure 28 shows the vertical load on one of the windward legs for differing inundation levels. Between 6 and 10 seconds, the time during which hull inundation occurs, the load reduces significantly due to hull buoyancy. At an inundation level of 3m, the minimum vertical load is reduced to 3.8MN as compared to 24.8MN with no inundation. Once the inundation phase is complete, the vertical load on the windward legs returns to a similar value as the no inundation case. It should be noted that the simulations described above did not take the effects of a preceding wave into account. 7.4 Effect of Wave Theory on Structural Response In Section 5, the MSL wave-in-deck model predicted that Airy waves overestimate vertical wave-in-deck loads by around 20% as compared to higher order wave theories such as Stream Function and Stokes 5th order theory. In order to assess the implications of the use of Airy waves on structural response, an Airy wave was applied to the USFOS model and the results compared with those of a Stream Function wave for 3m inundation. Figure 29 shows the vertical loads on the windward and leeward legs for the two wave types. It may be seen from the upper graph that the Airy wave produces a longer duration of reduced vertical reaction as well as a lower minimum value. For this case (i.e. 3m inundation), negative vertical spudcan loads are predicted. The lower part of Figure 29 demonstrates that the maximum vertical load on the leeward leg is around 10% higher for the Airy wave case as compared to the Stream Function wave. 7.5 Effect of Response to Previous Wave on Extreme Wave Response Twelve separate analyses were performed using the methodology outlined in Section 6.6, using a range of values of Dt between 1 and 12 seconds. This represents the full duration of the global sway mode, implying that the full enve lope of structural response, including the maximum response, has been captured. An inundation level of 3m and extreme linear foundation springs were used. Figure 30 shows the horizontal displacements for the twelve analyses performed. The difference between the highest and lowest maximum displacements is relatively small, indicating that the pre-wave conditions have a relatively small effect on the final maximum horizontal response. The maximum displacement case occurred where the main wave struck the hull when the structure had a large displacement in the opposite sense than the direction of

24

wave attack - i.e. at this phase angle the structure has its largest acceleration in the direction of wave attack. Figure 31 shows the envelope of vertical reactions on the windward leg for a range of Dt values. It may be seen that the minimum foundation reaction is 1.9MN, as compared to the mean minimum vertical load value of +1.8MN. It appears from this result that an assessment of the likely conditions before the extreme wave strikes the hull may be important, especially where leg tension/lift-off is a possibility. 7.6 Effect of Foundation Fixity on Structural Response 7.6.1 Introduction

In order to assess the importance of the method of foundation modelling on structural response the following modelling approaches were used and the results compared: Linear springs (3 different rotational stiffnesses used) Non-linear uncoupled springs Non-linear coupled springs

Details of the above are contained in the following sections. 7.6.2 Linear Foundations

The results of analyses using the three linear foundation stiffnesses detailed in Table 5 above are shown in Figure 32. Each analysis was performed with a hull inundation of 2.8m. It is clear that the foundation stiffness has a large effect on structural response the maximum horizontal displacement is approximately three times greater for the extreme foundation stiffness compared to the fatigue stiffness. However, foundation stiffness does not appear to have a very large effect on the minimum vertical load in the case of the windward legs, with the sudden decrease in reaction due to buoyancy effects being largely independent of the foundation conditions. The differences in minimum vertical reaction are principally due to additional P-delta moment due to differences in horizontal displacement. 7.6.3 Non-Linear Foundation Springs

The effect of the foundation modelling method on structural response was investigated by using three modelling methods: linear, non-linear and coupled non-linear springs. All three foundation models had the same initial stiffnesses in the horizontal, vertical and rotational directions. The effect of the type of foundation modelling used on horizontal displacement is illustrated in Figure 33. It may be seen that the maximum displacement is highly dependent on the foundation modelling method used (Table 9):
Table 9 Effect of Foundation Modelling Method on Structural Response Foundation model Linear springs Non-linear uncoupled springs Coupled springs (USFOS spudcan element) Maximum displacement (m) 3.35 3.98 5.01 Increase on linear case +19% +50%

25

Modelling the foundations as coupled non-linear springs represents the most accurate approach available for representing the true response of the foundations. It may be seen that it is this modelling approach that gives the largest response to extreme wave loading. Figure 34 shows the vertical loads and moments at the base of a windward leg during a wave cycle. The maximum ultimate moment capacity occurs at a vertical load of around 100MN. It may be seen that in all cases, the vertical load on the windward leg foundation is below 100MN, and hence the foundation moment capacity reduces as the applied vertical load reduces. In the case of the linear spring model, the applied moment exceeds the capacity of the foundation, with a maximum moment of 508MNm. The point of maximum moment occurs at a time where the vertical load is relatively low, hence giving a small moment capacity of 120MNm. The low vertical load on the windward legs is due to the global overturning moment on the structure in combination with large buoyancy loads caused by hull inundation. For non-linear coupled springs, the moment at the foundation is limited by the ultimate capacity of the foundation (in this case about 260MNm based on an axial load of 80MN). As the springs are uncoupled, the moment capacity is independent of vertical load, and hence during the period of low vertical leg loading (i.e. during hull inundation) the foundation is outside the predicted failure surface. In the case of the coupled non-linear springs, it may be seen from Figure 34 that the loads on the foundation remain within the moment failure surface. This has a particularly large effect when the vertical loads are at a minimum (i.e. during hull inundation) where the moment is limited to less than 80MNm. The effect the method of foundation modelling has on the moment at the base of the windward leg is demonstrated in Figure 35. The moment is limited to below 150MNm between 5 and 10 seconds for the spudcan element case, whereas the moment in the linear and the uncoupled non-linear foundation models exceed this value during this time. This leads to a reduction in global stiffness for the spudcan element case compared to the linear and uncouple d non-linear foundation models, hence leading to greater displacements (Figure 33). Figure 36 shows the phasing of vertical foundation loads for the leeward and windward legs (spudcan foundation case). The upper and lower limits for the vertical load are taken from Figure 21, and signify where the horizontal foundation capacity is significantly reduced from the maximum capacity of 12MN. It may be seen from Figure 36 that the windward leg reaction is almost always greater than the lower threshold value. In this case, there is no significant shear redistribution from the windward legs to the leeward leg. However, if the vertical load was reduced by a slightly larger amount, leg sliding would be predicted, which would in turn lead to a significantly higher shear load on the leeward leg. Figure 37 shows the phasing of the horizontal and vertical loads on the leeward leg. It may be seen that the maximum vertical load occurs around 2 seconds after the peak horizontal load, and hence in this case the loads on the leeward leg foundation remain well within the failure envelope shown in Figure 21. This implies that leg sliding caused by excessively large vertical loads is not predicted for the leeward leg.

26

8. CONCLUSIONS

The following conclusions have been made as a result of the analyses presented in this report: Wave inundation causes a significant step change in environmental loading. The vertical loads generated during hull inundation are considerable, and are principally due to buoyancy effects. For example, an inundation of 3m caused a vertical load equivalent to 40% of the structural dead weight. The use of Airy waves (as compared to higher order wave theories such as Stream Function and Stokes 5th Order) is likely to underestimate horizontal loading and peak overturning moment whilst overestimating vertical loads. Aeration at the top of the wave is unlikely to significantly affect wave-in-deck loading. The Jack-Up unit analysed reacted statically to vertical wave-in-deck loading. This is likely to be the case for most types of Jack-Up. The vertical loads on the windward legs reduce to almost zero for inundations of between 2m and 3m. The response to the extreme wave is not highly dependent on the effects of the preceding wave, hence suggesting that the DAF to be used in conjunction with an extreme wave (with hull inundation) should be relatively low. Where linear foundations are used, the horizontal displacement is highly dependent on the level of fixity assumed. The use of coupled non-linear foundations increases horizontal response and reduces the moments at the bottom of the windward legs. This will in turn increase the moments at the leg/hull connections.

27

28

9. REFERENCES

1. MSL Engineering Ltd. 'Assessment of the Effect of Wave-in-Deck Loads on a Typic al Jack-Up'. HSE Offshore Technology Report OTO 2001 034, June 2001. 2. Howarth, M., Dier, A. and Jones, W. A Study of Jack-Up Hull Inundation under Extreme Waves, Eighth International Conference on the Jack-Up Platform, City University, 2001. 3. SACS Release 5 User Manual. Engineering Dynamics, Inc. 1999. 4. Site Specific Assessment of Mobile Jack-Up Units, TR 5-5A, Society of Naval Architects and Marine Engineers (SNAME), Jersey City, 1997. 5. ASAS-WAVE User Manual - Wave Loader for Offshore Structures, Version 13, 2001. 6. BOMEL and Offshore Design. Review of Wave-in-Deck Load Assessment Procedures, Offshore Technology Report No. OTO 97 073. 7. Kaplan, P., Murray, J.J. and Y, W.C. Theoretical Analysis of Wave Impact Forces on Platform Deck Structures. Offshore Mechanics and Arctic Engineering Conference, Copenhagen, June 1995. 8. API Recommended Practice 2A-WSD, Twenty-first Edition, December 2000. 9. Barltrop N. and Adams A. Dynamics of Fixed Marine Structures, MTD / ButterworthHeinemann, 1991 10. Howarth M et. al. "Scale effects of wave impact pressures on Cob armour units" 25th International Conference on Coastal Engineering, September 1996, Orlando, publn. ASCE, New York 11. Crawford, A. J., Bullock, G. N., Hewson, P. J. and Bird, P. A. D. 1998. Wave impact pressures and aeration at a breakwater. Ocean Wave Measurement and Analysis. American Society of Civil Engineers, Vol.2, pp 1366 -1379. 12. Bea R, Iverson R and Xu T. "Wave-in-deck Forces on Offshore Platforms", Journal of Offshore Mechanics and Arctic Engineering. Volume 123, Number 1, February 2001. 13. USFOS Version 7-7 Users Manual, Sintef Group, 2000. 14. Van Langen, H and Hospers, B. Theoretical Model for Determining Rotational Behaviour of Spud Cans, Offshore Technology Conference 7302, 1993. 15. MSL Engineering Limited. Interpretation of Full-Scale Monitoring Data from a Jack-Up Rig, HSE Offshore Technology Report No. 2001 035. 16. Maersk Endurer Mobile Jack-Up Drilling Unit. Shearwater Site-Specific Assessment Studies. Noble Denton Report No. L17726/NDE/GAH, 1997.

29

wave direction

top of wave assumed to be sliced off

horizontal wavein-deck load air gap SWL

wave + current load on legs

buoyancy + vertical wave loads

50 year crest

sea bed
Figure 1 MSL Wave-in-Deck Model

30

80m

72m

b(x)

PLAN

surface profile h(x,t) v(z,x,t)

PARTICLE VELOCITY z hull

u(z,x,t)

ELEVATION

Figure 2
MSL Wave-in-Deck Model - Nomenclature

31

aeration,% 0 0 5 10 15 20 25 30 35

0.2 depth / velocity head

0.4

0.6

0.8

k=0.685 k=1 (Bob Bea) k=1.369 k=2.054

1.2

Figure 3 Aeration Profiles for Wave-in-Deck Load Calculation

32

Stage 1: Just prior to wave in deck loading - t = -1.25s Structure has initial displacement, velocity and accelerations from response to previous waves

Stage 2: Top of crest at hull front face - t = 0 sec 1. Maximum horizontal wave-indeck loading 2. Buoyancy loads acts to give an additional positive overturning moment 3. Wave loading on legs yet to reach a maximum value

Stage 3: t = 1.25s to 3.25s 1. Maximum wave loading on legs has occurred shortly before stage 2. Zero horizontal wave-in-deck loading 3. Buoyancy load providing a negative overturning moment (i.e. a 'righting' moment)

Figure 4 Graphical Representation of Wave Cycle

33

Horizontal Load 14E+6 12E+6 10E+6


Force (N)

8E+6 6E+6 4E+6 2E+6 000E+0 -2 -1 0 1


time (s)

Vertical Load 80E+6 70E+6 60E+6


Force (N)

buoyancy hydrodynamic

50E+6 40E+6 30E+6 20E+6 10E+6 000E+0 -2 -1 0 1


time (s)

Total Overturning Moment 3E+9 2E+9 2E+9


Moment (Nm)

1E+9 500E+6 000E+0 -2 -1 -500E+6 -1E+9


time (s)

Figure 5
Typical Wave-in-Deck Loads Generated using MSL Model

34

50E+6

legs + wave-in-deck legs only


40E+6

base shear (N)

30E+6

20E+6

10E+6

000E+0 -2 -1 -10E+6 time (s) 0 1 2 3 4 5

5E+9 legs + wave-in-deck legs only overturning moment (Nm) 4E+9

3E+9

2E+9

1E+9

000E+0 -2 -1 0 1 time (s) 2 3 4 5

Figure 6
Wave-in-Deck Loads + Wave Loads on Legs, Inundation = 3m

35

20E+6 1m 2m 15E+6 3m 4m

Load (N)

10E+6

5E+6

000E+0 -2 -1 0 1 time (s) 2 3 4

Figure 7
Effect of Inundation on Horizontal Wave-in-Deck Load

120E+6 1m 2m 3m 80E+6 Load (N) 4m

100E+6

60E+6

40E+6

20E+6

000E+0 -2 -1 0 1 time (s) 2 3 4

Figure 8
Effect of Inundation on Buoyancy

36

3.0E+09 1m 2m 3m 4m

Overturning moment (Nm)

2.0E+09

1.0E+09

0.0E+00 -2 -1 0 1 2 3 4

-1.0E+09 time (s)

Figure 9 Effect of Inundation on Overturning Moment due to Vertical Wave-in-Deck Loads

3.0E+09 1m 2m 3m 4m

Overturning moment (Nm)

2.0E+09

1.0E+09

0.0E+00 -2 -1 0 1 2 3 4

-1.0E+09 time (s)

Figure 10 Effect of Inundation on Total Overturning Moment due to Wave-in-Deck Loads

37

130 120 110 elevation (m) 100 90 80 70 60 50 40 0 2 4 6 8 10 time (s) 12 14 16 18


Airy (linear) Stream function Stokes 5th order Shell NewWave

Figure 11
Surface profiles of 4 types of deterministic wave

120

100

80
elevation (m)

60

40
Airy (linear) Stokes 5th order Stream function Shell NewWave 0
0 1 2 3 4 5 6 7 8 9 10
horizontal particle velocity (m/s)

20

Figure 12
Horizontal particle velocity profiles of 4 types of deterministic wave

38

Maximum Horizontal Loads 30E+6 Stream 25E+6 Load (N) 20E+6 15E+6 10E+6 5E+6 000E+0 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 inundation (m) Stokes Airy NewWave

140E+6 120E+6 100E+6 Load (N) 80E+6 60E+6 40E+6 20E+6 000E+0 0 0.5 1
Stream Stokes Airy NewWave

Maximum Vertical Loads

1.5

2.5

3.5

4.5

inundation (m)

Maximum Combined Overturning Moment 5E+9 overturning moment (Nm) Stream 4E+9 3E+9 2E+9 1E+9 000E+0 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 inundation (m) Stokes Airy NewWave

Figure 13
Effect of Wave The ory on Wave-in-Deck Loads

39

0.8

0.6 Load 0.4

0.2

Buoyancy Horizontal Peak OTM

0 0 5 10 15 20 25 30 35
aeration at free surface (%)

Figure 14
Effect of Wave Aeration on Wave-in-Deck Loads

40

Figure 15 USFOS Structural Model

41

Wave direction

Elevation of Front Face of Hull 1a 8a 9a 4a 10a 16a 17a 11a 5a 12a 6a 21a 22a 18a 19a 20a 13a 14a 7a 15 21b 22b 18b 19b 20b 13b 14b 6b 17b 16b 10b 11b 12b 5b Each member split into 99 subdivisions 7b Drag dummy member 2a 3a Vertical drag dummy members at the hull front face Each member split into 99 subdivisions Drag Members at front face: Cd x D chosen to give correct horizontal load. Buoyancy set to zero.

8b 1b 2b 3b

4b

Vertical buoyancy dummy member at the centre of each zone Buoyancy dummy member

Hull Plan

Buoyancy members:
Diameter chosen to give correct
area and hence buoyancy load Cd set to zero.

Figure 16 USFOS Wave-in-Deck Model

42

1.2E+07

Horizontal Load
USFOS MathCAD

1.0E+07

8.0E+06

Force (N)

6.0E+06

4.0E+06

2.0E+06

0.0E+00 -2 -1 0 1 2 3 4 5

time (s)

5.0E+07

Buoyancy Load

USFOS

4.0E+07

MathCad

Force (N)

3.0E+07

2.0E+07

1.0E+07

0.0E+00 -2 -1 0 1
time (s)

Figure 17
Validation of USFOS Wave-in-Deck Model

43

300E+6

250E+6

200E+6

moment (Nm)

150E+6

100E+6

USFOS element 50E+6 Yield Moment Ultimate Moment 000E+0 0 0.005 0.01 0.015 0.02 0.025 0.03

rotation (radians)

Figure 18 Typical SPUDCAN Moment-Rotation Relationship


140E+6 120E+6 100E+6 horizontal load (N) 80E+6 60E+6 40E+6 20E+6 000E+0 0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1 horizontal displacement (m)

Figure 19
Typical USFOS SPUDCAN Horizontal Load-Displacement Relationship

44

200 180 160


yield moment ultimate moment ND ultimate capacity

vertical load (MN)

140 120 100 80 60 40 20 0 0 50 100 150 200

250

300

moment (MNm)

Figure 20
Vertical Load vs. Moment Capacity, USFOS SPUDCAN Element

200 180 160 140 vertical load (MN) 120 100 80 60 40 20 0 0 2 4 6 8 horizontal load (MN) 10 12 14
yield load ND ultimate capacity

Figure 21
Vertical Load vs. Sliding Capacity, USFOS SPUDCAN Element

45

'Extreme' Linear Foundation Springs


8
full wave-in-deck model wave-in-deck model (no buoyancy) no wave-in-deck loads

hull displacement (m)

4
2

0
0 -2
-4
2 4 6 8 10 12 14 16 18 20

-6
time (s)

Figure 22 Effect of Inclusion of Wave-in-Deck Loads on Displacement

46

Windward Leg Load


100E+6

80E+6 vertical leg load (N)

60E+6

40E+6

20E+6

full wave-in-deck model wave-in-deck model (no buoyancy) no wave-in-deck loads

000E+0 0 2 4 6 8 10 time (s) 12 14 16 18 20

160E+6
140E+6
120E+6
vertical leg load (N) 100E+6
80E+6
60E+6
40E+6
20E+6
000E+0
0 -20E+6
2 4

Leeward Leg Load


full wave-in-deck model wave-in-deck model (no buoyancy) no wave-in-deck loads

10 time (s)

12

14

16

18

20

Figure 23 Effect of the Inclusion of Wave-in-Deck loads on Vertical Foundation Loads

47

3m Inundation, "Extreme" Springs

8
6

horizontal displacement (m)

4
2
0
0 -2
-4
-6
-8

time (s) Figure 24 Effect of Relative Velocity on Structural Response


relative velocity terms included relative velocity terms ignored

10

15

20

48

150E+6

Windward Leg
relative velocity ignored

125E+6

relative velocity included

vertical load (N)

100E+6

75E+6

50E+6

25E+6

000E+0 0 2 4 6 8 10 12 14 16 18 20

time (s)

200E+6

Leeward Leg

relative velocity ignored 150E+6 relative velocity included

vertical load (N)

100E+6

50E+6

000E+0 0 5 10 15 20

-50E+6

time (s)

Figure 25 Effect of Relative Velocity on Vertical Foundation Reactions

49

"Extreme" Linear Foundation Springs


6 5 horizontal displacement (m) 4 3 2 1 0 -1 -2 -3 -4 -5 time (s) 0 2 4 6 8 10 12 14 16 18 20
no inundation 1m inundation 2m inundation 3m inundation

Figure 26 Effect of Increasing Inundation on Horizontal Displacement

"Extreme" Linear Foundation Springs


140E+6

120E+6
no inundation

100E+6 vertical load (N)

1m inundation 2m inundation 3m inundation

80E+6

60E+6

40E+6

20E+6

000E+0 0 2 4 6 8 10 time (s) 12 14 16 18 20

Figure 27 Effect of Increasing Inundation on Leeward Leg Reaction

50

90E+6 80E+6 70E+6 60E+6 vertical load (N) 50E+6 40E+6 30E+6
Hull Buoyancy

20E+6 10E+6 00E+0 0 2 4 6 8 10 time (s) 12 14

0m inundation 1m inundation 2m inundation 3m inundation

16

18

20

Figure 28 Effect of Increasing Inundation on Reaction of Windward Legs

51

windward leg

100E+6

80E+6

vertical load (N)

60E+6

40E+6

20E+6

stream function Airy

000E+0 0 -20E+6
time (s)

10

15

20

leeward leg

140E+6 120E+6 100E+6

vertical load (N) stream function Airy

80E+6
60E+6
40E+6
20E+6
000E+0
0 5 10
time (s)

15

20

Figure 29 Comparison of Foundation Loads, Stream Function and Airy waves

52

"Extreme" Linear Springs, Inundation = 3m


8
envelope of maximum response

4 horizontal displacement (m)

0 0 -2 5 10 15 20 25

-4

-6 time (s)

Figure 30 Effect of Preceding Wave on Horizontal Displacement

53

90E+6
80E+6
70E+6
60E+6
vertical reaction (N) 50E+6
40E+6
30E+6
20E+6
10E+6
000E+0
2 -10E+6
4

"Extreme " Linear Foundation Springs, Inundation = 3m

10

12 time (s)

14

16

18

20

22

24

Figure 31 Effect of Preceding Wave on Windward Leg Vertical Reaction

54

Horizontal Displacement

'Extreme' foundation springs USFOS initial stiffness 'Fatigue' foundation springs

4 displacement (m)

0 0 -2 2 4 6 8 10 12 14 16 18 20

-4

-6 time (s)

90E+6 80E+6 70E+6 vertical reaction (N) 60E+6 50E+6 40E+6 30E+6 20E+6 10E+6 000E+0 0 2 4 6

Windward Leg

'Fatigue' foundation springs 'USFOS' initial stiffness 'Extreme' foundation springs

10 time (s)

12

14

16

18

20

Figure 32 Effect of Different Linear Foundation Stiffnesses on Response

55

5
4

horizontal displacement (m)

linear springs non-linear springs spudcan elements

3
2
1
0
0 -1 -2
time (s)

10

15

20

Figure 33 Comparison of Horizontal Displacement for Different Foundation Models

56

200E+6

foundation failure surface linear springs

150E+6

non-linear uncoupled springs spudcan element

vertical load (N)

100E+6

initial position

50E+6

-400E+6

-200E+6

000E+0 000E+0

200E+6

400E+6

600E+6

moment (Nm)
Figure 34 Comparison of Windward Leg Foundation Behaviour for Different Foundation Models

57

600E+6
SPUDCAN elements linear springs non-linear uncoupled springs

400E+6

foundation moment (Nm)

200E+6

000E+0 0 2 4 6 8 10 12 14 16 18 20

-200E+6

-400E+6
time (s) Figure 35 Windward Leg Foundation Moments for Different Foundation Models

58

140E+6 windward leg leeward leg 120E+6

100E+6

horizontal foundation capacity signifcantly reduced above this line

vertical load (N)

80E+6

60E+6

40E+6

20E+6

horizontal foundation capacity signifcantly reduced below this line

000E+0 0 2 4 6 8 time (s) 10 12 14 16 18

Figure 36 Phasing of Vertical Foundation Loads

59

140E+6

25E+6

vertical load horizontal load


120E+6 vertical load (N) 20E+6 100E+6

15E+6 80E+6

60E+6 10E+6

40E+6 5E+6 20E+6

000E+0 0 2 4 6 8 time (s) 10 12 14 16 18

000E+0

Figure 37 Phasing of Foundation Loads, Leeward Leg

Printed and published by the Health and Safety Executive C1.25 04/03

horizontal load (N)

ISBN 0-7176-2177-4

CRR 019

25.00

9 780717
621774

Das könnte Ihnen auch gefallen