You are on page 1of 145

OFFSHORE FOUNDATIONS: TECHNOLOGIES , DESIGN AND

APPLICATION

Pedro Gomes Simões de Abreu


Thesis to obtain the Master of Science Degree in Civil Engineering

Masters in Civil Engineering


Supervisor: Dr. Peter Bourne-Webb

Examination Committee:
Chairperson: Prof. Jaime Santos

Supervisor: Dr. Peter Bourne-Webb

Members of the Committee: Prof. Alexandre Pinto

JULY 2014
Offshore Foundations: Technologies, Design and Application

i
Offshore Foundations: Technologies, Design and Application

ACKNOWLEDGMENTS

First and foremost I would like to take this opportunity to express my sincere gratitude to Dr
Peter Bourne-Webb for encouraging me to pursue this research. Dr Peter was always available to
discuss various aspects of the work and remained an important source of guidance throughout this
project.

It would be remiss of me not to thank my dearest colleague Ana Tavares. Without her I would
not accomplish this entire process, so for that, thank you for your support during these last 5 years.

A special thank you also to my family, particularly my sister and my parents, your guidance,
encouragement, love and understanding, not only over the past year but throughout my life has been
an inspiration. I would like to thank my grandfather Antonio for his effort over the last 70 years to
help all my family to pursue our studies as far as we all wanted.

Finally, all my friends that faced by my side my academic path throughout these last seven
years, thank you for your unwavering support has been a source of strength and encouragement.

This work is dedicated to my parents (Virgílio and Lurdes) and my aunt Ana.

Pedro G. Simões de Abreu

ii
Offshore Foundations: Technologies, Design and Application

iii
Offshore Foundations: Technologies, Design and Application

ABSTRACT

The offshore oil industry started over 60 years ago, since then it evolved immensely. This
evolution was forced by the need of exploiting oil and gas reserves in more challenging regions. The
purpose of this study was to gather information about the foundation structures used in the offshore
industry, and to assess the applicability of two types of foundation in a real scenario. São Tome &
Principe (STP) was selected as the case-study for this paper because it is a member of the Community
of Portuguese Language Countries, and has recently been subjected to several studies in its offshore
region to evaluate its potential as an oil & gas supplier. This paper described the geotechnical
characterisation of the offshore of STP based on investigations performed in the Gulf of Guinea (GoG)
for more than 10 years. The results of the study were that the soil is probably a highly sensitive clay
(St=2 to 4), and the shear strength profile presents a gradient of about 1.5 kPa/m. Another conclusion
is that many sites in the GoG exhibit a greater resistance (up to about 15 kPa) in the first 2 m, this
phenomenon is called a “crust”. This work also describes design principles for two anchoring
systems: the Torpedo Anchors and Suctions Embedded Plate Anchors (SEPLA). For Torpedo, the
results revealed that the pull-out resistance, after reconsolidation, is expected to be 8.7 MN. Whereas,
the results for SEPLA holding capacity is expected to be 10 MN. For both systems the calculations
were made for the largest of their solutions available in the market.

Keywords: Offshore Foundations, ultra-deep water, Torpedo Anchors, SEPLA, São Tomé &
Princípe.

iv
Offshore Foundations: Technologies, Design and Application

v
Offshore Foundations: Technologies, Design and Application

TABLE OF CONTENTS
Acknowledgments ..................................................................................................................................................... ii
Abstract.......................................................................................................................................................................... iv
List of Figures.............................................................................................................................................................. ix
List of Tables .............................................................................................................................................................. xii
List of Symbols ......................................................................................................................................................... xiv
1. Introduction & Motivation ............................................................................................................................1
1.1. Context & Motivation ............................................................................................................................................... 1
1.2. Thesis structure .......................................................................................................................................................... 4
2. Offshore production facilities......................................................................................................................5
2.1. Introduction ........................................................................................................................................................................ 5
2.2. Fixed Platforms ........................................................................................................................................................... 5
2.3. Compliant Tower ........................................................................................................................................................ 7
2.4. Tension Leg Platform ............................................................................................................................................... 8
2.5. Semi-Submersible Floating Production Systems ........................................................................................ 8
....................................................................................................................................................................................................... 9
2.6. SPAR Platform .......................................................................................................................................................... 10
2.7. Floating Production Storage and Offloading Facility ............................................................................. 11
2.8. Subsea System .......................................................................................................................................................... 13
3. Shallow Water Foundations...................................................................................................................... 14
3.1. Spudcans ..................................................................................................................................................................... 14
3.1.1. Geotechnical Calculations ......................................................................................................................... 16
3.2. Pile Foundations ...................................................................................................................................................... 16
3.2.1. Driven Piles ...................................................................................................................................................... 17
3.2.2. Grouted Piles ................................................................................................................................................... 19
3.2.3. Pile Resistance ................................................................................................................................................ 20
3.3. Gravity Base Structures........................................................................................................................................ 22
3.4. Concrete Caissons for Tension Leg Platforms ........................................................................................... 25
3.5. Steel Buckets for Jackets ...................................................................................................................................... 27
3.5.1. Installation in Clay ........................................................................................................................................ 28
3.5.2. Installation in Sand....................................................................................................................................... 28
4. Deep and Ultra-Deep Water Foundations .......................................................................................... 28
4.1. Gravity Anchors ....................................................................................................................................................... 30
4.2. Pile Anchors ............................................................................................................................................................... 31

vi
Offshore Foundations: Technologies, Design and Application

4.3. Suction Caissons ...................................................................................................................................................... 32


4.4. Vertically Loaded Drag Anchor ........................................................................................................................ 34
4.5. Suction Embedded Plate Anchor ..................................................................................................................... 37
4.6. Dynamically Penetrated Anchor ...................................................................................................................... 38
5. Geological Characterization ...................................................................................................................... 43
5.1. Topographical features of ocean floors ........................................................................................................ 44
5.2. Seabed Geology ........................................................................................................................................................ 47
5.2.1. Seabed sediments origin and classification ...................................................................................... 48
5.2.2. Geotechnical characteristics of some offshore regions .............................................................. 52
5.3. Geohazards ................................................................................................................................................................. 53
5.3.1. Triggering Mechanisms for Submarine Slope Failures ............................................................... 55
5.3.2. Geohazard identification ........................................................................................................................... 56
5.3.3. Geotechnical site investigation ............................................................................................................... 59
5.3.4. Submarine slope failures and slides .................................................................................................... 60
6. Case study: Geotechnical considerations for ultra-deep oil fields off São Tome e Principe
63
6.1. Proposed development ........................................................................................................................................ 63
6.2. Geotechnical Site Conditions ............................................................................................................................. 63
6.2.1. Index Properties of Gog Sediments ...................................................................................................... 66
6.2.2. In Situ Stresses and Stress History ....................................................................................................... 69
6.2.3. Shear Strength Profiles ............................................................................................................................... 72
6.3. Design of anchor solutions ................................................................................................................................. 77
6.3.1. Torpedo Anchors........................................................................................................................................... 77
6.3.2. SEPLA .................................................................................................................................................................. 90
6.4. Discussion of Results .......................................................................................................................................... 106
7. Conclusion and Further Research ....................................................................................................... 109
7.1 Conclusion....................................................................................................................................................................... 109
7.2. Further Research ........................................................................................................................................................ 110
Index .......................................................................................................................................................................................... 121
Appendix I – Details of Shallow Foundation Studies. .......................................................................... 123
Appendix II – Calculation to determine Loss in Anchor Embedment. ......................................... 125
Appendix III – Calculations to determine the resistance of the SEPLA. ...................................... 127
Wilde et al. (2001) .............................................................................................................................................................. 127
Merifield et al. (2001) ....................................................................................................................................................... 127
DNV-RP-E302 (2002)........................................................................................................................................................ 129

vii
Offshore Foundations: Technologies, Design and Application

viii
Offshore Foundations: Technologies, Design and Application

LIST OF FIGURES
FIGURE 1 – TREND IN WATER DEPTH FOR EXPLORATION AND DEVELOPMENT DRILLING SINCE 1940. ................................................. 1
FIGURE 2 – MARINE FOUNDATIONS TYPE BY APPLICATION AND WATER DEPTH. FPSO- FLOATING PRODUCTION, STORAGE AND OFFLOADING
VESSEL. VLA- DRAG EMBEDMENT VERTICALLY LOADED ANCHOR. SEPLA- SUCTION EMBEDDED PLATE ANCHOR. DPA- DYNAMIC
PENETRATING ANCHOR. .......................................................................................................................................... 2
FIGURE 3 – MAP OF SÃO TOMÉ & PRINCIPE ISLANDS AND THE ECONOMIC EXCLUSIVE ZONE, COURTESY OF AGENCIA NACIONAL DE
PETRÓLEO DE STP. ............................................................................................................................................... 3
FIGURE 4 – OFFSHORE DEVELOPMENTS SYSTEMS: BOTTOM SUPPORTED, VERTICALLY MOORED STRUCTURES, FLOATING PRODUCTION AND
SUBSEA SYSTEMS, FROM: HTTP://I46.PHOTOBUCKET.COM/ALBUMS/F133/TAMAAA/UPLOAD1/PICTURE1.PNG. ....................... 5
FIGURE 5 – STEEL JACKET PLATFORMS AND CONCRETE GRAVITY STRUCTURE, FROM:
HTTP://PETROAHDAL.WEBS.COM/APPS/PHOTOS/PHOTO?PHOTOID=109394836 AND HTTP:// OFFSHORE-
MAG.COM/ARTICLES/2013/JAN/GO-AHEAD-FOR-14-BILLION-HEBRON-PROJECT-OFFSHORE-EASTERN-CANADA.HTML. ................ 6
FIGURE 6 – OFFSHORE PLATFORM ELEMENTS, FROM:
HTTP://WWW.CONSERVATION.CA.GOV/DOG/PICTURE_A_WELL/PAGES/OFFSHORE_PLATFORM.ASPX. .................................... 6
FIGURE 7 – TRANSPORTATION OF THE BULLWINKLE STEEL SUBSTRUCTURE (JACKET), COURTESY OF SHELL INTL.
HTTP://WWW.ESA.ORG/ESABLOG/WP-CONTENT/UPLOADS/2011/10/STEEL-JACKET-BEING-TOWED-OFFSHORE.JPG. ................. 7
FIGURE 8 – PETRONIUS COMPLIANT TOWER AND EMPIRE STATE BUILDING FOR COMPARISON, FROM:
HTTP://PETROWIKI.ORG/FILE%3AVOL3_PAGE_530_IMAGE_0001.PNG. ....................................................................... 7
FIGURE 9 – WORLDWIDE FLEET OF INSTALLED AND SANCTIONED TLPS, COURTESY OF BRITISH PETROLEUM FROM:
HTTP://PETROWIKI.ORG/FILE%3AVOL3_PAGE_532_IMAGE_0001.PNG. ....................................................................... 8
FIGURE 10 – SEMI-SUBMERSIBLE VESSEL WITH TWIN HULLS (COLUMNS AND PONTOONS).
HTTP://OILANDGASPROCESSING.BLOGSPOT.PT/2009/02/OIL-RIG-OFFHORE-STRUCTURE.HTML ............................................. 9
FIGURE 11 - WORLDWIDE FLEET OF INSTALLED AND SANCTIONED SEMI-SUBMERSIBLE FPS, COURTESY OF BRITISH PETROLEUM FROM:
HTTP://PETROWIKI.ORG/FILE%3AVOL3_PAGE_531_IMAGE_0002.PNG. ....................................................................... 9
FIGURE 12 – THREE DIFFERENT SPAR PROFILES, FROM: HTTP://IMAGES.PENNWELLNET.COM/OGJ/IMAGES/OGJ2/9644JSK02.GIF. ..... 10
FIGURE 13 – WORLDWIDE FLEET OF INSTALLED SPARS, COURTESY OF BRITISH PETROLEUM. ........................................................ 11
FIGURE 14 - FLOATING PRODUCTION STORAGE AND OFFLOADING FACILITY, FROM:
HTTP://OILANDGASPROCESSING.BLOGSPOT.PT/2009/02/OIL-RIG-OFFHORE-STRUCTURE.HTML. .......................................... 12
FIGURE 15 – SUBSEA SYSTEM COMPONENTS. BOP - BLOWOUT PREVENTER. FROM: HTTP://WWW.RYANSRANTINGS.COM/?P=817. .... 13
FIGURE 16 – JACK-UP INSTALLATION PROCEDURE. FROM: HTTP://IGSDELHICHAPTER.COM/IGSDC-
2014_FOUNDATIONSFOROFFSHORE.PDF................................................................................................................. 15
FIGURE 17 – SOME EXAMPLE SPUDCAN SHAPES (RANDOLPH ET AL., 2005) .............................................................................. 15
FIGURE 18 – PILE DRIVING METHOD TROUGH A JACKET LEG (DEAN, 2009) ............................................................................... 18
FIGURE 19 – DETAIL OF THE MUDMATS AT THE BOTTOM OF THE STEEL LATTICE STRUCTURE,
HTTP://WWW.TENSIONTECH.COM/SERVICES/TEXTILE_SOLUTIONS.HTML......................................................................... 18
FIGURE 20 - ARRANGEMENT FOR PILE GROUP AROUND A LEG (DEAN, 2009) ............................................................................ 19
FIGURE 21 – STAGES IN INSTALLATION OF AN OFFSHORE DRILLED AND GROUTED PILE (RANDOLPH ET AL., 2005) ............................... 20
FIGURE 22 – (A) PLANT OF EKOFISK TANK AND (B) CONDEEP GRAVITY BASE DESIGNS (RANDOLPH ET AL., 2005). .............................. 23
FIGURE 23 – CONDEEP STRUCTURES INSTALLED, FROM: HTTP://WWW.INRISK.UBC.CA/FILES/2012/11/CONDEEPCOMPARISON.PNG .. 23
FIGURE 24 –YOLLA A HYBRID GRAVITY BASED STRUCTURE EXAMPLE, BASS STRAIT, AUSTRALIA (WATSON & HUMPHESON,2005) ....... 24
FIGURE 25 – PRINCIPAL DESIGN ISSUES FOR PARALLEL DESCENT INSTALLATION: (A) DOWEL PENETRATION, (B) SKIRT PENETRATION, (C) BASE
SUCTION OR PRESSURE, (D) DOME CONTACT STRESSES, (E) GROUTING PRESSURES AND DENSITY, AND (F) SCOUR PROTECTION (DEAN,
2009). ............................................................................................................................................................ 25
FIGURE 26 – (A) INSET, FOUNDATION TEMPLATE SHOWING CLUSTER OF CONCRETE CAISSONS (B) PREDICTED AND MEASURED CYCLIC LOAD
VS. DISPLACEMENT RESPONSE (CHRISTOPHERSEN, 1993) ............................................................................................. 26
FIGURE 27 – STAGES OF INSTALLATION OF A BUCKET FOUNDATION FOR A JACKET STRUCTURE ....................................................... 27
FIGURE 28 – SEEPAGE PRESSURES SET UP DURING SUCTION INSTALLATION OF BUCKET FOUNDATION (ERBRICH & TJELTA, 1999). ......... 28
FIGURE 29 – LOCATIONS OF SOME CURRENT DEEP-WATER DEVELOPMENTS (DEAN, 2009) .......................................................... 29

ix
Offshore Foundations: Technologies, Design and Application
FIGURE 30 – (A) CATENARY MOORING AND DRAG EMBEDMENT ANCHOR. (B) TAUT LEG MOORING AND VLA. FROM:
HTTP://EVENTS.ENERGETICS.COM/DEEPWATER/PDFS/PRESENTATIONS/SESSION5/RODERICKRUINEN.PDF ............................... 30
FIGURE 31 – GRAVITY ANCHORS (RANDOLPH ET AL., 2005) ................................................................................................. 31
FIGURE 32 – (A) ANCHOR PILE AND CHAIN. (B) FORCES ON AN ELEMENT OF THE ANCHOR LINE (RUINEN, 2005) ............................... 32
FIGURE 33 – MOORING APPLIANCE POINT AND VARIATION OF PADEYE DEPTH WITH LOADING ANGLE FOR GIVEN CENTER OF ROTATION
(RANDOLPH & HOUSE, 2002) ............................................................................................................................... 33
FIGURE 34 – SUCTION CAISSON INSTALLATION STEPS, FROM:
HTTP://WWW.EPD.GOV.HK/EIA/REGISTER/REPORT/EIAREPORT/EIA_1672009/HKOWF%20HTML%20EIA/MAIN%20TEXT%2
0FIGURES/FIG%202.36A.PNG. ............................................................................................................................. 34
FIGURE 35 – DRAG ANCHORS, (A) FIXED FLUKE ANCHOR. (B) VRYHOF STEVMANTA VLA. SOURCE: VRYHOF ANCHOR MANUAL (2005) .. 35
FIGURE 36 – INSTALLATION METHODS ............................................................................................................................. 36
FIGURE 37 – (A) ANCHOR DEPTH DETERMINATION. (B) VLA SIMPLE RECOVERY, VRYHOF ANCHOR MANUAL (2005). ......................... 36
FIGURE 38 – COMPONENTS OF A SUCTION EMBEDDED PLATE ANCHOR (GAUDIN ET AL., 2006) ..................................................... 37
FIGURE 39 –SCHEMATIC OF SEPLA INSTALLATION (YANG ET AL., 2012; AND COURTESY OF INTERMOOR). ..................................... 38
FIGURE 40 – RADIUS COMPARISON BETWEEN FLOATING UNITS LINKED TO CONVENTIONAL DRAG ANCHORS AND TORPEDO ANCHORS, FROM:
HTTP://WWW.HINDAWI.COM/JOURNALS/JAM/2012/102618.FIG.002.JPG. ................................................................. 39
FIGURE 41 – DYNAMICALLY PENETRATING ANCHORS (A) TORPEDO ANCHOR WITH FINS AND WITHOUT FINS (MEDEIROS, 2002); (B)
INSTALLATION OF 4 FLUKES TORPEDO ANCHOR (MEDEIROS, 2002; O’LOUGHLIN ET AL., 2004) ............................................ 40
FIGURE 42 – FULL SCALE TORPEDO PILE AND RELEASING SITUATION, LIENG ET AL. (1999) ............................................................ 41
FIGURE 43 – ADVANTAGES AND DISADVANTAGES OF DIFFERENT DEEP WATER ANCHOR TYPES (EHLERS ET AL., 2004). ........................ 42
FIGURE 44 – TOPOGRAPHICAL FEATURES OF OCEAN FLOORS, AFTER
HTTP://WWW.MARINEBIO.NET/MARINESCIENCE/01INTRO/WOIMG/XSECTOPO.JPG. ......................................................... 45
FIGURE 45 – PROFILES OF OCEAN FLOOR TOPOGRAPHY AT SELECTED LOCATIONS, RANDOLPH ET AL. 2011 ....................................... 45
FIGURE 46 – BATHYMETRIC SHADED RELIEF MAP SHOWING A DETAILED EXTENT OF CONTINENTAL MARGINS (NUMBER IN RED RELATES TO
LOCATIONS IN FIGURE 45), US NATIONAL GEOPHYSICAL DATA CENTRE........................................................................... 46
FIGURE 47 – SEDIMENT THICKNESS OF THE WORLD’S OCEANS AND MARGINAL SEAS. SOURCE:
HTTP://WWW.NGDC.NOAA.GOV/MGGLSEDTHICK/SEDTHICK.HTML. ............................................................................... 48
FIGURE 48 – DISTRIBUTION OF SEDIMENT ACROSS A PASSIVE CONTINENTAL MARGIN, FROM:
HTTP://CLASSCONNECTION.S3.AMAZONAWS.COM/927/FLASHCARDS/68927/JPG/4-91305062763712.JPG. .................... 49
FIGURE 49 - SEDIMENTARY PROCESS OF MARINE DEPOSITS (AFTER SILVA 1974) ........................................................................ 49
FIGURE 50 – TURBIDITY CURRENT: (A) SEABED TOPOGRAPHY WHERE SLOPE BREAK MAY START A TURBIDITY CURRENT (B) TURBIDITY
CURRENT PROGRESS (C) SETTLEMENT TURBIDITY PARTICLES. FROM: HTTP://PEOPLE.MATHS.OX.AC.UK/FAY/RESEARCH.HTML....... 50
FIGURE 51 – DISTRIBUTION OF THE DIFFERENT SEDIMENTS DEPOSITS ACROSS THE WORLD, FROM:
HTTP://CLASSCONNECTION.S3.AMAZONAWS.COM/927/FLASHCARDS/68927/JPG/4-91305062763712.JPG. .................... 51
FIGURE 52 – MICROGRAPHS OF (A) CALCAREOUS SAND AND (B) SILICA SAND (DEAN, 2009). ....................................................... 52
FIGURE 53 – WORLDWIDE DISTRIBUTION OF OFFSHORE OIL AND GAS DEVELOPMENTS. DEAN (2009) BASED ON MCCLELLAND (1974) AND
POULOS (1988). ................................................................................................................................................ 53
FIGURE 54 – SCHEMATIC DIAGRAMS SHOWING MAIN OFFSHORE GEOHAZARDS (A) WIDELY USED SUMMARY OF DEEPWATER GEOHAZARDS
(POWER ET AL., 2005) (B) MORE GEOHAZARDS (STROUT AND TJELTA, 2007). ................................................................. 54
FIGURE 55 – SUMMARY OF GENERALISED INVESTIGATION ELEMENTS (CAMPBELL ET AL. 2008) ..................................................... 57
FIGURE 56 – TESTS THAT SHOULD BE PERFORMED DURING A LABORATORIAL TESTING PROGRAM (RANDOLPH ET AL., 2011). ............... 60
FIGURE 57 – DEFINITION OF SLIDE MOBILITY...................................................................................................................... 61
FIGURE 58 – COMPARISON OF VOLUME AND RUN-OUT DISTANCE OF SUBMARINE AND SUB-AERIAL SLIDES (RANDOLPH ET AL. 2011 AFTER
SCHEIDEGGER 1973, EDGERS AND KARLSRUD 1982, HAMPTON ET AL. 1976, DADE AND HUPPERT 1998, DE BLASION ET AL.
2006) ............................................................................................................................................................. 61
FIGURE 59 – SÃO TOMÉ & PRINCIPE LOCATION AND SURROUNDING GEOLOGY, COURTESY OF AGENCIA NACIONAL DO PETROLEO OF STP
FROM: HTTP://WWW.STP-EEZ.COM/DOWNLOADS/POSTERS/3_STP_REGIONALGEOL.PDF . .............................................. 64
FIGURE 60 – CROSS SECTION FROM NIGERIA COST TO PRINCIPE ISLAND, COURTESY OF AGENCIA NACIONAL DO PETROLEO OF STP FROM:
HTTP://WWW.STP-EEZ.COM/DOWNLOADS/POSTERS/3_STP_REGIONALGEOL.PDF . ........................................................ 64
FIGURE 61 – CROSS SECTION FROM PRINCIPE ISLAND TO EQUATORIAL GUINEA COST, COURTESY OF AGENCIA NACIONAL DO PETROLEO OF
STP FROM: HTTP://WWW.STP-EEZ.COM/DOWNLOADS/POSTERS/3_STP_REGIONALGEOL.PDF . ........................................ 65

x
Offshore Foundations: Technologies, Design and Application
FIGURE 62 – THE RED CIRCLE LIMITS THE EXTENT OF GULF OF GUINEA. .................................................................................... 65
FIGURE 63 – DEEP-WATER SEDIMENTS PHYSICAL PROPERTIES ON GULF OF GUINEA (PUECH, 2004). .............................................. 66
FIGURE 64 – DEEP-WATER SEDIMENTS PHYSICAL PROPERTIES, DATA FROM INTERNATIONAL OCEAN DISCOVERY PROGRAM TESTING
RESULTS, AFTER MASCLE ET AL. (1998). SITES ARE LOCATED OFFSHORE OF IVORY COAST AND WITH DEPTHS RANGING FROM 2090
TO 4637 METERS, FIGURE 66................................................................................................................................ 67
FIGURE 65 – COMPARISON OF THE IODP TESTING RESULTS (OFFSHORE OF IVORY COAST) AND PUECH (2004) SEDIMENT PHYSICAL
PROPERTIES IN GOG. ........................................................................................................................................... 67
FIGURE 66 – MAP SHOWING THE DRILLING LOCATIONS OF IODP IN GULF OF GUINEA, FROM: HTTP://WWW-
ODP.TAMU.EDU/PUBLICATIONS/159_IR/IMAGES/MAP.JPG . ....................................................................................... 68
FIGURE 67 - PLASTICITY INDEX OF GOG SEDIMENTS IN DEPTH (PUECH, 2004). ......................................................................... 68
FIGURE 68 – PLASTICITY CHART OF GOG DEEP-WATER SEDIMENTS, CASAGRANDE DIAGRAM (PUECH, 2004). .................................. 69
FIGURE 69 – CV VERSUS WL CHART (KULHAWY & MAYNE, 1990) ........................................................................................... 69
FIGURE 70 – TYPICAL IN SITU STRESS PROFILES FOR GULF OF GUINEA DEEP-WATER SOILS AND THE COMPARISON WITH GULF OF MEXICO
(PUECH, 2004).................................................................................................................................................. 70
FIGURE 71 – VERTICAL YIELD PRESSURE AND YSR VERSUS PENETRATION DEPTH FOR TYPICAL GOG SITES (J.L. COLLIAT & H. DENDANI ET
AL., 2011). ....................................................................................................................................................... 71
FIGURE 72 - UNDRAINED SHEAR STRENGTH VARIATION WITH DEPTH IN GULF OF GUINEA, DETERMINED BY INTERNATIONAL OCEAN
DISCOVERY PROGRAM USING VANE SHEAR AND PENETROMETER TESTS, AND TEST LOCATIONS............................................... 73
FIGURE 73 – TYPICAL CONE RESISTANCE PROFILE FOR GOG DEEP WATER SEDIMENTS (WD – WATER DEPTH), PUECH (2004). ............. 74
FIGURE 74 – UNDRAINED SHEAR STRENGTH PROFILES OF GULF OF MEXICO, BRAZIL AND GOG PROPOSED BY PUECH (2004). .............. 75
FIGURE 75 – SHEAR STRENGTH PROFILES OF DIVERSE MARINE SITES ........................................................................................ 76
FIGURE 76 –NIGERIAN CONTINENTAL SLOPE STUDY AREA, SULTAN ET AL. (2007). ..................................................................... 76
FIGURE 77 – SHEAR STRENGTH PROFILE OF NIGERIAN CONTINENTAL SLOPE OBTAINED IN LABORATORY GEOTECHNICAL TESTS AND ITS
COMPARISON WITH SHEAR STRENGTH PROFILES PROPOSED BY OTHERS............................................................................. 77
FIGURE 78 – DESIGN PROCESS FOR TORPEDO ANCHORS. ...................................................................................................... 78
FIGURE 79- SCHEMATIC LONGITUDINAL SECTION DRAWING OF THE T-98, BRANDÃO ET AL. (2006). .............................................. 78
FIGURE 80 – PHOTOS OF THE T-98 BODY SECTIONS WELDING AND ITS FINAL ADJUSTMENTS, BRANDÃO ET AL. (2006). ...................... 79
FIGURE 81 – SENSITIVITY OF EMBEDMENT DEPTH PREDICTIONS TO SHAFT ADHESION FACTOR ACCORDING TO RICHARDSON ET AL. (2008)
AND INTERPOLATION FOR ADHESION FACTOR OF 0.8 AND IMPACT VELOCITY OF 40M/S. ...................................................... 83
FIGURE 82 – SENSITIVITY OF EMBEDMENT DEPTH PREDICTIONS TO UNDRAINED SHEAR STRENGTH GRADIENT, RICHARDSON ET AL. (2008).
...................................................................................................................................................................... 84
FIGURE 83 – COMPARISON OF CENTRIFUGE AND FIELD TEST EMBEDMENT DATA, O’LOUGHLIN ET AL. (2013). .................................. 84
FIGURE 84 – CROSS SECTION OF A TORPEDO ANCHOR WITH FOUR FLUKES (AGUIAR EL AL., 2009). ................................................ 87
FIGURE 85 – T-98 HOLDING CAPACITY FOR 4 TILTS AND 3 DIFFERENT LOAD DIRECTIONS IN CAMPOS BASIN, BRANDÃO ET AL. (2006). .. 87
FIGURE 86 – TYPICAL SEPLA WITH KEYING FLAP, WANG ET AL. (2012) .................................................................................. 91
FIGURE 87 – PLATE ANCHOR SETUP BEFORE KEYING PROCESS. ............................................................................................... 93
FIGURE 88 – GRAPH SHOWING THE RESULTS ACHIEVED BY O’LOUGHLIN ET AL. (2006) FOR ECCENTRICITY RATIOS BETWEEN 0.17AND 1.0.
...................................................................................................................................................................... 94
FIGURE 89 – COMBINED LOADING PATHS FOR HIGH AND LOW ECCENTRICITY PLATE ANCHORS, O’LOUGHLIN ET AL. (2006). ................ 94
FIGURE 90 – LOSS IN ANCHOR EMBEDMENT DURING KEYING VERSUS ANCHOR GEOMETRY FACTOR, SONG ET AL. (2009). .................... 97
FIGURE 91 – PROBLEM NOTATION ................................................................................................................................ 100
FIGURE 92 – SHALLOW AND DEEP ANCHOR BEHAVIOUR, MERIFIELD ET AL. (2001) .................................................................. 101
FIGURE 93 – DESIGN CHART FOR RECTANGULAR ANCHORS IN CLAY, ALLOWS DETERMINING ANCHOR BREAK-OUT FACTOR NCO AT VARIOUS
EMBEDMENT RATIOS (MERIFIELD ET AL., 2003). ...................................................................................................... 102
FIGURE 94 – EQUATION AND CURVE DESCRIBING THE VARIATION OF NC VALUE IN SHALLOW FAILURE ZONE, DAHLBERG ET AL. (2004). . 104

xi
Offshore Foundations: Technologies, Design and Application

LIST OF TABLES

TABLE 1 –DESIGN PARAMETERS FOR STEEL PILES IN SILICEOUS SANDS IN ISO (2004) .................................................................. 21
TABLE 2 – PILE DESIGN MULTIPLIER PARAMETER (𝜶) FOR CLAYS BASED ON 𝒔𝒖/𝝈𝒗′ OR OCR (SEMPLE AND GEMEINHARDT, 1983). .... 21
TABLE 3 – SHAFT FRICTION AND END BEARING CAPACITY ON CARBONATE SANDS BASED ON KOLK (2000). 𝝉𝒔, 𝒔𝒊 AND 𝒒𝒔𝒊 ARE
RESPECTIVELY THE UNIT SHAFT FRICTION AND UNIT END BEARING OF SILICEOUS SAND, WHILE 𝝉𝒔, 𝒄𝒂 AND 𝒒𝒄𝒂 ARE RESPECTIVELY
THE UNIT SHAFT FRICTION AND UNIT END BEARING OF CALCAREOUS SAND......................................................................... 22
TABLE 4 – SURVEYS AND THEIR PURPOSE (DEAN, 2009) ...................................................................................................... 43
TABLE 5 – DRAG COEFFICIENTS ACCORDING TO DIFFERENT PENETROMETER SHAPE AND REFERENCE, ON FREEMAN AND HOLLISTER
FORMULA L IS THE PENETROMETER LENGTH AND D IS THE DIAMETER. .............................................................................. 80
TABLE 6 – ADOPTED VALUES FOR DIFFERENT SOIL CHARACTERISTICS AND REFERENCES FROM WHICH THEY WERE BASED ON. ................. 81
TABLE 7 – MODEL AND PROTOTYPE ANCHOR DIMENSIONS AND SOIL PROPERTY. ........................................................................ 83
TABLE 8 – SOIL PROPERTIES, PENETRATION AND VERTICAL RESISTANCE OF CAMPOS BASIN IN BRAZIL AND STP OFFSHORE. AVERAGE
UNDRAINED SHEAR STRENGTH GRADIENT WAS PROPOSED BY MEDEIROS (2001)................................................................ 89
TABLE 9 - ADVANTAGES AND DRAWBACKS OF SUCTION PILES AND DRAG EMBEDDED PLATE ANCHORS (VLAS), WILDE ET AL. (2001) ..... 90
TABLE 10 – PARTIAL SAFETY FACTOR FOR ANCHOR RESISTANCE, DAHLBERG ET AL. (2004). ........................................................ 105
TABLE 11 – PLATE ANCHOR HOLDING CAPACITIES IN THE IDEALIZED STP OFFSHORE CONDITIONS ACCORDING TO THREE DIFFERENT
AUTHORS. ....................................................................................................................................................... 106
TABLE 12 – SEPLA AND TORPEDO ANCHOR ADVANTAGES. ................................................................................................. 108
TABLE 13 – SEPLA AND TORPEDO ANCHOR DISADVANTAGES. ............................................................................................. 108

xii
Offshore Foundations: Technologies, Design and Application

xiii
Offshore Foundations: Technologies, Design and Application

LIST OF SYMBOLS
𝜃 angle
𝜂 Empirical reduction factor
𝐾 coefficient of lateral earth pressure
𝐷𝑟 Relative density
σ’vy yield vertical effective stress
α Adhesion factor
ze , H Embedment depth
Wsub submerged weight of the anchor in the water
wL Liquid limit
w Water content
vT Terminal velocity
vimpact Impact velocity
v velocity
t thickness
su,o undrained shear strength at the ground surface
St sensitivity
Rs friction resistance
Rb End bearing resistance
PI Plasticity index
p’c Maximum effective stress previously applied
p’0 effective stress currently applied
Nco break-out factor
NC end bearing capacity factor
ma Mass of ballast
M0 Initial moment
m’ Effective mass
m mass
M Moment
L lenght
k,ρ Undrained shear strength gradient
h heigth
g Acceleration of gravity
Fs,max Maximum Parallel force
Fs Parallel force
Fn,max Maximum normal force
Fn Normal force
Fd drag force
F Force
f Shank resistance
Etotal Total energy
e Loading eccentricity
deff Effective diameter
D* Optimal depth
d diameter
cv Coefficient of consolidation
CD Drag coefficient
B breadth
Ap Sectional area
a acceleration
∆𝑧𝑒 Loss in anchor embedment
𝜏𝑠−𝑚𝑎𝑥 limiting unit skin friction
𝜏𝑠 Unit shaft friction

xiv
Offshore Foundations: Technologies, Design and Application

𝜏𝑠,𝑠𝑖 unit shaft friction of siliceous sands


𝜏𝑠,𝑐𝑎 unit shaft friction of calcareous sand
𝜏𝑓,𝑐𝑦 Cyclic shear strength
𝜎′𝑣0 actual vertical effective stress
𝜎𝑣′ vertical effective stress
𝜌𝑤 Water density
𝛿′ soil-pile friction angle
𝛾𝑠 unit weight of the soil
𝛾𝑚 Partial safety factor
𝑧𝑖 Installation penetration depth
𝑠𝑢,𝑚𝑒𝑎𝑛 Mean static undrained shear strength
𝑠𝑐 Shape factor
𝑞𝑠𝑖 unit end bearing of siliceous sand
𝑞𝑐𝑎 unit end bearing of calcareous sand
𝑞𝑏𝑢−𝑚𝑎𝑥 limiting bearing pressure on the base
𝑞𝑏𝑢 bearing pressure on the base
𝑘𝜃 Gradient for estimating anchor loss in anchor embedment
𝑒𝑤 Loading eccentricity for anchor weight
𝑒𝑓 Loading eccentricity for friction resistance
𝑊𝐹 Equivalent plate width
𝑈𝑐𝑦 Cyclic loading factor
𝑁𝑞 Bearing capacity factor
𝐿𝐹 Equivalent plate length
𝐸𝑝 Potential energy
𝐸𝑐 Kinetic energy
𝐴𝑡 Transversal area
𝐴𝑠 Shaft area
𝑠̅𝑢 , su,avg average shaft undrained shear strength
𝑊′ 𝑎 Overall submerge anchor weight
𝑠𝑢 undrained shear strength

xv
1. INTRODUCTION & MOTIVATION
1.1. CONTEXT & MOTIVATION
The offshore oil industry started in 1947 with the installation of the first oil rig in just 6
metres of water, off the coast of Louisiana in the United States. Nowadays, there are over 7000
offshore platforms around the world, located in a large range of water depths which since the late-
1990s, are starting to exceed 2000 m. This progression forced a change in the concept of “deep
water”; in the 1970s deep-water meant depths of 50 m to 100 m while now deep-water refers to
water depths around 800 m. Further to this, when referring to water depths greater than about 1000
m, the phrase “ultra-deep” water is now used. Figure 1 shows the evolution of exploration water
depths through time, from the mid-20th century on.

Figure 1 – Trend in water depth for exploration and development drilling since 1940.

Over the past century, many advances have been made in the development of offshore
technology, including the foundations of the structures that support the working platforms. There
are currently, several potential foundation solutions for any given offshore site conditions. The
circumstances of a particular site may be either fair or harsh depending on: the weather conditions
such as wind, currents and waves; or the local conditions such as the water-depth, seabed
geotechnical quality and topography, among other particular circumstances.

Nowadays offshore platforms can be positioned either in shallow, deep or ultra-deep water
columns, which results in very different foundation solutions. These definitions will be advanced in
the following chapters. Not only the foundation systems vary with water depth, but also the offshore
platform selected to explore the oil & gas reserve may differ. Figure 2 shows a diagram where the
different foundation solutions are classified by water depth.

1
Offshore Foundations: Technologies, Design and Application

Marine Foundations

Shallow (<500 m) Deep & Ultra-Deep

Gravity Based Steel Concrete Embedded


Spudcans Piles Gravity
Structures Buckets Caissons Anchors

Driven Grouted Grillage Suction


Tank Condeep Piles Piles
Box Piles
Caissons
VLA SEPLA DPA
and berm

Torpedo
Anchors

Figure 2 – Marine Foundations type by application and water depth. FPSO- floating production, storage and offloading vessel. VLA- drag embedment
vertically loaded anchor. SEPLA- suction embedded plate anchor. DPA- dynamic penetrating anchor.

2
Offshore Foundations: Technologies, Design and Application

As near-shore and thus, shallow water resources have largely been exploited, the
industries´ attention has turned to the exploration of more oil and gas fields that are further
offshore and in deeper water. As a consequence, some countries where Portuguese is the official
language have gained attention, and as a result some are currently under investigation (e.g.
Mozambique, Guinea Bissau and São Tomé & Principe), while in others investments have already
been made (e.g. Brazil and Angola).

São Tome & Principe is a member of the group of countries which belong to the Community
of Portuguese Language Countries (CPLP) and over the past five years has been subject to many
field tests in order to quantify the potential oil & gas reserves, and to evaluate the quality of the
potential extractable product (oil) of those reserves as well. Therefore, the chosen scenario to
assess the feasibility of employing two different foundation systems for possible future
production developments is off the coast of Sao Tome & Principe (STP). The exclusive economic
zone (EEZ) of STP is now divided into several blocks which can be seen in Figure 3. Those blocks
are licensed to Oil & Gas companies, so they can develop investigation work and evaluate the
potential of the reserves.

Figure 3 – Map of São Tomé & Principe islands and the economic exclusive zone, courtesy of Agencia
Nacional de Petróleo de STP.

3
Offshore Foundations: Technologies, Design and Application

As a final course thesis of Civil Engineering specializing in geotechnics, the work presented
in this thesis has the purpose of gathering information about the foundation structures used in
the offshore petroleum industry and the assessment of the application of two types of foundation
in a realistic scenario – in this instance the developments proposed for STP described above.

1.2. THESIS STRUCTURE


Following this introduction, Chapters 2 through 4 present a review of the literature
relating to existing offshore foundation systems. In Chapter 2, offshore platforms are discussed; it
is from these facilities where the operations for oil and gas extraction are developed. They can
either be fixed platforms that stand above the water and are directly fixed to the seabed or floating
platforms that are fixed to the seabed by means of mooring systems. The review continues with
the presentation of the two main foundation groups, i.e. “shallow” water (Chapter 3) and “deep &
ultra-deep” water (Chapter 4) foundations. Within each grouping the various foundation concepts
are presented, and the design of these foundation systems is explained.

Chapter 5 summarizes the different aspects involved in marine geotechnical


characterization. First, the topographical features of ocean floors are explained and then the
genesis of the seabed geology is described, i.e. the origin of the sediments that form the seabed.
Lastly, geohazards that are commonly found in the seabed are described, as well as their triggering
mechanisms.

Chapter 6 presents a case-study developed for the offshore region of the Western African
country São Tomé & Principe. The chapter starts by explaining which foundation systems will be
assessed for their feasibility of use in that region; as this region typically has water depths greater
than 1000 m, the proposed solutions will be based on anchor type systems. Then, the geotechnical
site conditions are characterized based on the findings from the Gulf of Guinea published by other
authors. Finally, a conceptual design of the foundations is made and the results obtained are
discussed in terms of which might be more suitable for use.

4
Offshore Foundations: Technologies, Design and Application

2. OFFSHORE PRODUCTION FACILITIES

2.1. INTRODUCTION
Many factors may influence the selection of an appropriate development strategy for
offshore hydrocarbon deposits. Among those factors considered are (French et al., 2006):

 Water depth
 Reservoir size
 Proximity to existing infrastructure
 Number of wells
 Operating considerations
 Economic factors
 Anticipated well intervention frequency (e.g. frequency of well replacement)

Offshore structures can be broadly categorised as fixed platforms, compliant towers,


floating structures and subsea systems. The various types of offshore development systems
currently in use are shown in Figure 4.

Figure 4 – Offshore developments systems: Bottom supported, vertically moored structures,


floating production and subsea systems, from:
http://i46.photobucket.com/albums/f133/tamaaa/upload1/Picture1.png.

2.2. FIXED PLATFORMS


Fixed platforms are working decks supported by legs directly connected onto the seabed
foundations. They can be tubular steel jackets, concrete or hybrid gravity structures. Steel jackets

5
Offshore Foundations: Technologies, Design and Application

are primarily pile supported, while gravity structures achieve stability by virtue of their immense
structural weight and large diameter base. Additional stability may be provided by use of base
skirts which penetrate several meters into the seabed. The limitation of this kind of structures is
related with economic issues; therefore they are not used in water depths greater than 500 m.
Figure 7 shows the jacket of the Bullwinkle platform located in Gulf of Mexico, the platform has
total height of 529 m and it extends 412 m below the waterline.

Figure 5 – Steel jacket platforms and concrete gravity structure, from:


http://petroahdal.webs.com/apps/photos/photo?photoid=109394836 and http:// offshore-
mag.com/articles/2013/jan/go-ahead-for-14-billion-hebron-project-offshore-eastern-
canada.html.

Figure 6 – Offshore Platform elements, from:


http://www.conservation.ca.gov/dog/picture_a_well/Pages/offshore_platform.aspx.

6
Offshore Foundations: Technologies, Design and Application

Figure 7 – Transportation of the Bullwinkle steel substructure (jacket), courtesy of Shell Intl.
http://www.esa.org/esablog/wp-content/uploads/2011/10/Steel-jacket-being-towed-
offshore.jpg.

2.3. COMPLIANT TOWER


A compliant tower is a slender steel space-frame tower with a piled foundation. These
towers are designed to be more flexible in bending than fixed structures and, therefore, more
“compliant” to the environment. This flexibility means that the platform can withstand significant
lateral loads by sustaining large lateral deflections. Compliant towers are typically applicable in
water depths ranging from 500 m to 600 m. The tallest in the world is the Petronius tower,
illustrated in Figure 8, which stands in 535 m of water and has a total height of 609.9 m, making
it one of the highest freestanding structures ever built (Chevron 2000).

Figure 8 – Petronius compliant tower and Empire State Building for comparison, from:
http://petrowiki.org/File%3AVol3_Page_530_Image_0001.png.

7
Offshore Foundations: Technologies, Design and Application

2.4. TENSION LEG PLATFORM


A Tension Leg Platform (TLP) consists of a semi-submersible platform moored by vertical
tendons connected to the seafloor. The excess buoyancy provided by various hull components
maintains the tension in the mooring system even during storm loading conditions. TLPs are
capable of being used in water depths up to 2000 m. Figure 9 illustrates the worldwide fleet of
installed TLPs, with special attention to Kizomba A platform (see arrow) which is located in
Angola (country member of CPLP), in nearly 1200 m of water.

Figure 9 – Worldwide fleet of installed and sanctioned TLPs, courtesy of British Petroleum from:
http://petrowiki.org/File%3AVol3_Page_532_Image_0001.png.

2.5. SEMI-SUBMERSIBLE FLOATING PRODUCTION SYSTEMS


A semi-submersible floating production system (FPS) typically comprises parallel
pontoons connected to the topside by numerous vertical columns, Figure 10. Semi-submersible
platforms can be moved from place to place, and the pontoons and columns can be filled with
water to alter the buoyancy of the system for improved stability under wave and wind loading.
Semi-submersibles can be deployed in a wide range of water depths for both temporary and
permanent operations. Figure 11 shows the FPSs used worldwide and the wide range of water
depths (300 m to 2000 m) where the system has been deployed.

8
Offshore Foundations: Technologies, Design and Application

Figure 10 – Semi-Submersible Vessel with twin hulls (columns and pontoons).


http://oilandgasprocessing.blogspot.pt/2009/02/oil-rig-offhore-structure.html

Figure 11 - Worldwide fleet of installed and sanctioned semi-submersible FPS, courtesy of British Petroleum from:
http://petrowiki.org/File%3AVol3_Page_531_Image_0002.png.

9
Offshore Foundations: Technologies, Design and Application

2.6. SPAR PLATFORM


A SPAR consists of a large diameter, vertical, cylindrical hull which supports the platform
by means of excess buoyancy. Buoyancy chambers located near the top of the hull enable the
buoyancy of the structure to be controlled thereby maintaining platform stability. In addition,
spiral strakes fitted to the hull minimize lateral movement due to vortex shedding, improving
lateral stability. The SPAR can be anchored to the seabed by vertical tethers but catenary or taut
mooring lines are more common. These details can be seen in Figure 12, where two different SPAR
profiles are illustrated, i.e. a classic SPAR with soilid hull and truss SPAR. SPARs are theoretically
capable of being deployed in water depths up to 3000 m (Figure 13), and currently the deepest
platform in the world is the Perdido SPAR in the Gulf of Mexico, which floats in 2438 meters of
water.

Figure 12 – Three different SPAR profiles, from:


http://images.pennwellnet.com/ogj/images/ogj2/9644jsk02.gif.

10
Offshore Foundations: Technologies, Design and Application

Figure 13 – Worldwide fleet of installed SPARs, courtesy of British Petroleum.

2.7. FLOATING PRODUCTION STORAGE AND OFFLOADING FACILITY


Floating Production Storage and Offloading (FPSO) facilities comprise a large tanker type
vessel fitted with production and storage facilities, Figure 14. The storage capabilities of FPSO
mean that it may be suitable for economically marginal fields located in remote areas in which
pipeline infrastructure does not exist. Smaller shuttle tankers may be used to transport the
hydrocarbons to an onshore processing facility. FPSOs can be fixed in position or comprise
multiple mooring lines meeting at a single point. The single point mooring allows the tanker to
weathervane to achieve an optimal orientation with regard to the prevailing environmental
conditions. The key advantages of FPSOs relate to their ability to operate on short term or
permanent developments in water depths up to and exceeding 3000 m. For example, the P-50
FPSO is moored in approximately 1240 m of water in the Albacora Leste field in the deepwater
Campos Basin, Brazil (Brandão et al. 2006).

11
Offshore Foundations: Technologies, Design and Application

Figure 14 - Floating production storage and offloading facility, from:


http://oilandgasprocessing.blogspot.pt/2009/02/oil-rig-offhore-structure.html.

12
Offshore Foundations: Technologies, Design and Application

2.8. SUBSEA SYSTEM


Subsea systems typically comprise either a single subsea well, whose production is piped to a
nearby platform, or multiple wells producing through a manifold and pipeline system to a distant
production facility. Multi-component seabed facilities such as subsea wells, manifolds, control
umbilicals and flowlines allow subsea systems to recover hydrocarbons in water depths and
conditions that would normally preclude the installation of a conventional fixed or floating
platform. Subsea systems are capable of operating in any water depth (Richardson et al., 2008).
Figure 15 illustrates several subsea system components.

Figure 15 – Subsea system components. BOP - Blowout preventer. From:


http://www.ryansrantings.com/?p=817.

13
Offshore Foundations: Technologies, Design and Application

3. SHALLOW WATER FOUNDATIONS


Water depth is considered shallow if the seabed depth does not exceed about 500 meters.
The principal types of foundation used in this situation are:

 Spudcans
 Piles
 Gravity base structures (GBS)
 Concrete caissons
 Steel Buckets

All these solutions have advantages and disadvantages, some cannot be applied in all seabed soil
conditions and some are more favourable than others, e.g. GBS in soft clay soils induce higher
settlements than spudcans do, due to its enormous self-weight. Hence, the next sections will
describe their principal properties as well as their pros and cons.

3.1. SPUDCANS
In the offshore industry an important role is played by self-elevating mobile drilling units,
commonly known as jack-ups, due to their flexibility and cost-effectiveness (Randolph et al.,
2005). It has proved to be a very useful construction “tool”, especially when working in turbulent
sea areas, or breaking waves such as shoal or coastal waters, and in swift currents (Gerwick Jr. et
al., 2007).

These structures consist of a buoyant triangular unit resting on three or more retractable
legs. This unit supports drilling and other topside equipment; it moves onto the intended location
with legs retracted, then releases the legs onto the seabed, and raises the hull out of the water, as
shown on Figure 16. On jack-ups, the foundation legs operate independently of each other, and
their foundations are usually known as “spudcans”.

14
Offshore Foundations: Technologies, Design and Application

Figure 16 – Jack-up installation procedure. From: http://igsdelhichapter.com/IGSDC-


2014_FoundationsforOffshore.pdf.

These foundations have a unique geometry, since they are installed relying only on the
structure’s self-weight plus an additional designed preload, which is intended to minimize
settlement and improve resistance to environmental solicitations. Spudcans are roughly circular
in plan, typically they have a shallow conical underside (in the order of 15 to 30 degrees to the
horizontal) with a sharp protruding spigot. In the larger jack-ups operating today, the spudcan
diameter can exceed 20 meters, with the shapes varying with the manufacturer and rig (Randolph
et al., 2005). Figure 17 illustrates two different spudcan shapes. The usual height of a jack-up
structure is over 160 m.

Figure 17 – Some example spudcan shapes (Randolph et al., 2005)

Since jack-ups started to be applied in deeper water and harsher environments,


preloading also started to play a more important role in design. Preloading induces bearing
capacity failure in the soil beneath and around each spudcan, causing the spudcan to penetrate
into the seabed until the soil resistance equals the applied load. Using the jacks on one leg at a

15
Offshore Foundations: Technologies, Design and Application

time, the barge acting as the reaction, the legs are forced into the soil. After these a pile hammer
may be used on the top of the legs to gain even greater penetration. If the same penetration was
given to all legs a punch-through failure or a damage on the foundation might occur during
preloading, which could result in the jackup toppling over and one or more legs being bent or
broken. So, to avoid these risks, modern jackups are able to preload the spudcans individually
(Dean, 2009).

3.1.1. GEOTECHNICAL CALCULATIONS


Jackups are mobile units, and when designed the foundation conditions where it will be
applied are not usually known. So, for each new site, SNAME (2002) recommends that a site-
specific assessment be done. This may include:

 Assessment of geohazards
 Foundation assessment for installation, commonly including a preload check
 Foundation assessment for operations, including a sliding check and an overturning check
based on assessed vertical and horizontal actions
 Assessment of effects of the jackup on nearby structures
 Leg extraction assessment, for when the jackup is moved off site to another location

These analyses are preformed to predict the footing penetration during installation and
preloading, and the capacity to withstand a design storm.

3.2. PILE FOUNDATIONS


Piles are slender columnar elements in a foundation which have the function of
transferring load from the superstructure through weak compressible strata or through water,
onto stiffer or more compact and less compressible soils or onto rock. They may be required to
carry uplift loads when used to support tall structures subjected to overturning forces from winds
or waves. Piles used in marine structures are subjected to lateral loads from the impact of berthing
ships and from waves (M.J. Tomlinson, 1977).

Pile foundations can be used either in shallow or deep water, the link to the working
platform is what differs. In shallow water, the connection is typically made by a steel lattice
structure commonly called a jacket. This is the most used structure for fixed offshore platforms.
Piles can also be used as anchors in moored floating facilities, and this application will be
discussed further in Chapter 4.

There are two construction methods used for piles that are constructed offshore: driven
and grouted. The most commonly used are metallic driven piles because they are the most reliable

16
Offshore Foundations: Technologies, Design and Application

and have the easiest construction path. Although we are in a marine environment, there is no
problem with corrosion as the steel pile embedded in the seabed, has no contact with oxygen.

3.2.1. DRIVEN PILES


Offshore, the most frequently used pile type is the open-ended steel pipe, which is driven
into the seabed by a hammer. Pile diameters range from 0.76 m up to 2.5 m, but exceptionally a
diameter of 5.1 m has been successfully used on offshore wind turbines. The wall thickness may
vary along the pile length, so it will be thicker where moment is greater (near the pile head).
Typical diameter to wall thickness ratios (d/t) are between 20 and 60. The lower value represents
the greatest curvature that can normally be achieved in a steel rolling machine. The highest value
represents a curvature beyond which wall-buckling or section ovalisation effects can be common
(Barbour and Erbrich, 1994; MSL, 2001; Aldridge et al., 2005; Randolph et al., 2005).

The process of installing an offshore driven pile through a steel jacket leg is illustrated in
Figure 18. First, the steel jacket is released onto the required location, where it will be supported
by mudmats. Mudmats are templates used in the bottom of the steel jacket to avoid its undesirable
penetration into soft soils, Figure 7 and Figure 19 shows four mudmats at the bottom of a steel
jacket structure. Then the first section of pile is lowered trough the leg. A hammer is installed on
the head of the pile, and is used to drive the first segment until its limit, the pilling equipment is
removed and another pile segment is lifted on and welded in place. This weld is normally
subjected to non-destructive testing, after which the hammer is lifted back on, and the whole
procedure starts again until the designed pile penetration is achieved. Unless the jacket confines
the pile, grout is injected into the annular space to provide the structural connection between
them (Dean, 2009).

17
Offshore Foundations: Technologies, Design and Application

Figure 18 – Pile driving method trough a jacket leg (Dean, 2009)

Figure 19 – Detail of the mudmats at the bottom of the steel lattice structure,
http://www.tensiontech.com/services/textile_solutions.html

When more than one pile is required per leg, sleeves may be attached to the jacket around
the base of its legs. These sleeves operate both as guides and pile pooler. Piling is done in the same
way as for a leg pile, with each pile installed in several segments, if necessary. Figure 20 shows
the sleeve arrangement for pile group.

18
Offshore Foundations: Technologies, Design and Application

Figure 20 - arrangement for pile group around a leg (Dean, 2009)

3.2.2. GROUTED PILES


Even though driven piles are the most common type used in the offshore environment,
there is also the equivalent of a bored pile. It involves the grouting of a steel section, which is
inserted in a previously drilled hole. Figure 21 shows the stages in construction of a drilled and
grouted pile. In order to avoid collapse of loose uncemented material near the seabed, it is often
necessary to drive a primary pile first; alternatively stabilizing mud can also be used. This solution
is only used if an adequate drilling barge is already on the site, since it is more expensive to install
and has longer construction period than driven piles (Randolph et al, 2005).

Whenever the seabed is composed of calcareous sediments, and other potentially


crushable material, where the shaft friction obtained with driven piles has been found to be very
low, drilled and grouted piles are more reliable. The low shaft friction is associated with very low
radial effective stresses around the pile, a situation remedied by drilled and grouted pile
construction, where the original horizontal effective stresses in the ground can be restored by
appropriate grouting design (Randolph et al., 2005).

19
Offshore Foundations: Technologies, Design and Application

Figure 21 – stages in installation of an offshore drilled and grouted pile (Randolph et al., 2005)

3.2.3. PILE RESISTANCE


Although many studies have been made into the understanding of end-bearing and shaft
friction resistance of piles, design methods still rely on empirical methods. The determination of
the soil resistance can be made applying current offshore guidelines, e.g. API RP2A and ISO 19902,
or CPT- based methods. The latter’s advantage is that it takes account of the detailed stress history
of the soil around the pile (Randolph et al., 2005; Dean, 2009).

Unit parameters
Table 1 summarizes the unit parameters recommended by API RP2A and ISO 19902, and
also gives some data for carbonate sands. For granular material, drained conditions are assumed
to apply. The parameters are based on relative density and silt content.

The design approach for shaft friction is expressed as,

𝜏𝑠 = 𝐾𝜎𝑣′ 𝑡𝑎𝑛 𝛿 ′ ≤ 𝜏𝑠−𝑚𝑎𝑥

where K is a coefficient of lateral earth pressure, and the recommended values are between 0.7
and 0.8 for open ended piles loaded in compression and 0.5 to 0.7 for tension piles, 𝜎𝑣′ is the
vertical effective stress, 𝛿 ′ is the soil-pile friction angle, and 𝜏𝑠−𝑚𝑎𝑥 is a limiting unit skin friction,
which varies with soil type and density.

The limiting bearing pressure on the base of the pile is expressed as,

20
Offshore Foundations: Technologies, Design and Application


𝑞𝑏𝑢 = 𝑁𝑞 𝜎𝑣0 ≤ 𝑞𝑏𝑢−𝑚𝑎𝑥

Where 𝑁𝑞 ranges from 12 to 50 according to the grain size and relative density of the material. All
parameters needed to determine pile resistance in sand are given in Table 1.

Siliceous Sands
Soil 𝒒𝒃𝒖−𝒎𝒂𝒙
Soil density Dr(%) 𝜹′ (˚) 𝝉𝒔−𝒎𝒂𝒙 (kPa) 𝑵𝒒
description (MPa)
Loose 15-35 20 65 12 3
Medium 35-65 25 80 20 5
Sand
Dense 65-85 30 95 40 10
Very Dense 85-100 35 115 50 12
Silty Sand Loose, Med 15-65 20 65 12 3
Dense 65-85 25 80 20 5
Clayey Sand
Very Dense 85-100 30 95 40 10
Loose 15-35 15 45 8 2
Sandy Silt Med. Dense 35-85 20 65 12 3
Very dense 85-100 25 80 20 5
Table 1 –Design parameters for steel piles in siliceous sands in ISO (2004)

For cohesive material, axial pile failure is assumed to occur in undrained conditions. Shaft
friction is calculated as multiple, α of the undrained shear strength 𝑠𝑢 . The multiplier depends on
the ratio 𝑠𝑢 /𝜎𝑣′ or in terms of the over-consolidation ratio (OCR), using Semple and Gemeinhardt’s
(1983) relation. This last interpretation also expresses, 𝑠𝑢 = 0.2𝜎𝑣′ × 𝑂𝐶𝑅 0.85 . API RP2A
recommends that the unit end bearing be taken as 9 𝑠𝑢 in clay. Table 2 provides the way to
determine the adhesion multiplier, α.

ISO 19902 and API RP2A Interpreted in terms of OCR


Range of 𝒔𝒖 /𝝈′𝒗 𝜶 Approximate range of OCR 𝜶
𝒔𝒖 /𝝈′𝒗 ≤ 𝟎. 𝟐𝟓 1 𝑂𝐶𝑅 ≤ 1.3 1
𝟎. 𝟐𝟓 ≤ 𝒔𝒖 /𝝈′𝒗 ≤ 𝟏 0.5/ (𝑠𝑢 /𝜎𝑣′ ) 0.5 1.3 ≤ 𝑂𝐶𝑅 ≤ 6.6 1.11/𝑂𝐶𝑅 0.425
𝟏 ≤ 𝒔𝒖 /𝝈′𝒗 0.5/ (𝑠𝑢 /𝜎𝑣′ ) 0.25 6.6 ≤ 𝑂𝐶𝑅 0.74/𝑂𝐶𝑅 0.213
Table 2 – Pile design multiplier parameter ( 𝜶 ) for clays based on 𝒔𝒖 /𝝈′𝒗 or OCR (Semple and
Gemeinhardt, 1983).

It is now recommended that grouted piles are better suited to soil profiles consisting
primarily of calcareous and carbonate materials (Kolk, 2000). For soil profiles that contain thin

21
Offshore Foundations: Technologies, Design and Application

calcareous and carbonate sand layers, driven piles may still be feasible. Kolk’s recommendations
for open-ended driven piles in calcareous soils are summarized in Table 3.

Range of 𝑼𝒏𝒊𝒕 𝒔𝒉𝒂𝒇𝒕 𝒇𝒓𝒊𝒄𝒕𝒊𝒐𝒏, Unit end bearing, Notes


carbonate 𝝉𝒔−𝒎𝒂𝒙 (𝒌𝑷𝒂) 𝒒 (𝑴𝑷𝒂)
contents: CC
𝑪𝑪 ≤ 𝟐𝟎% As siliceous sand As siliceous sand -
𝟐𝟎% ≤ 𝑪𝑪 𝜏𝑠,𝑠𝑖 − 𝐾𝑐𝑐 (𝜏𝑠,𝑠𝑖 − 𝜏𝑠,𝑐𝑎 ) 𝑞𝑠𝑖 − 𝐾𝑐𝑐 (𝑞𝑠𝑖 − 𝑞𝑐𝑎 ) 𝐶𝐶
log( )
𝐾𝑐𝑐 = 20
≤ 𝟖𝟎% log(4)
𝟖𝟎% ≥ 𝑪𝑪 min(0.14𝜎𝑣′ ,15𝑘𝑃𝑎) 0.7𝑞𝑐 for coring, 𝑁𝑞 = 10 and
0.3𝑞𝑐 for plugged. 𝑞𝑙𝑖𝑚 = 3 𝑀𝑃𝑎 if
𝑞𝑐 determined from CPT no CPT data
cone (MPa). available
Table 3 – Shaft friction and end bearing capacity on carbonate sands based on Kolk (2000). 𝝉𝒔,𝒔𝒊 and
𝒒𝒔𝒊 are respectively the unit shaft friction and unit end bearing of siliceous sand, while 𝝉𝒔,𝒄𝒂 and 𝒒𝒄𝒂
are respectively the unit shaft friction and unit end bearing of calcareous sand.

3.3. GRAVITY BASE STRUCTURES


Gravity base structures (GBS) are designed to be founded at or just below the seafloor,
transferring their loads to the soil by means of shallow footings. Usually these structures are made
of reinforced and prestressed concrete, but some were built of steel or a hybrid of concrete and
steel. These structures have a large base “footprint” with purpose of minimizing soil-bearing
loads. An important advantage of these solutions is the possibility of oil storage within the base
structure, i.e. the base operates both as foundation and storage facility.

GBS are also used for offshore wind power plants. By the end of 2010, 14 of the world's
offshore wind farms were supported by gravity-base structures.

Design loads for an offshore GBS are superior to onshore conditions. Due to its large
volume, inertial forces under waves, earthquake, and impact from vessel or icebergs are much
greater than usual. Thus, sliding tends to become the dominant mode of failure. So, to prevent this
possibility, concrete or steel skirts and dowels are employed; these are designed to penetrate the
seabed and thus force the failure surface deeper below the seafloor. Skirts also provide protection
against scour and piping (Gerwick Jr. et al., 2007).

These structures have evolved over time, the first of its kind was the Ekofisk tank, which
was installed in the North Sea in 1973 (Clausen et al., 1975). The experience gained on this first
project lead to conceptualization of a better, and now common, concrete deep water structure

22
Offshore Foundations: Technologies, Design and Application

called Condeep. A Condeep gravity base comprises a number of cylindrical cells usually displayed
in a hexagonal arrangement, the underside of the cells has a convex profile and half a metre inside
the concrete skirts the top of the dome touches down on the seabed, as illustrated in Figure 22(b).
In Figure 22 it is visible the more complex design of Condeep foundation relative to the Ekofisk
tank is apparent. While the Ekofisk tank had short (40cm) concrete skirts, the Condeep has steel
skirts (to 3.5m), that project from concrete skirts. The condeep design has much smaller wave
forces acting on the structure as the major volume is located below the water surface. In Figure
23, it is possible to see this characteristic on some different condeep platforms.

Figure 22 – (a) Plant of Ekofisk tank and (b) Condeep gravity base designs (Randolph et al., 2005).

Figure 23 – Condeep Structures installed, from:


http://www.inrisk.ubc.ca/files/2012/11/CondeepComparison.png

Another unusual group of GBS has been developed, and it is a hybrid concrete-steel
solution, i.e. a mixed structure which has a concrete base and a steel lattice structure, Figure 24.
They have been applied at sites where calcareous muddy silts and sands dominate, because of
their lighter weight when compared to concrete GBS. The rapid construction time is a very
important reason for choosing this kind of solution (Dean, 2009).

23
Offshore Foundations: Technologies, Design and Application

Figure 24 –Yolla A Hybrid gravity based structure example, Bass Strait, Australia (Watson &
Humpheson,2005)

Gravity foundations are often considered to be more complicated than jackets because the
soil behaviour must be considered in a three-dimensional volume that stretches one or more base
diameters below, and several diameters to either side of the base. The main design requirement
is to determine the foundation footprint, and the skirt depth and spacing. There are three design
codes for gravity platforms, ISO 19903:2006, API RP2A and DNV-OS-C502:2012.

Besides structural design another really important aspect of the implementation of these
structures is the foundation construction method. The foundation is constructed in a dry dock, but
the most difficult phase of the construction is when fixing it to the sea ground. Most gravity
platforms are kept level as they are lowered to the seabed. This allows the dowels and then the
skirts to penetrate the seabed almost simultaneously across the footprint of the base. Dowels are
steel pipes up to about 2 m in diameter, they contact the seabed first and pin the platform to the
seabed and penetrate a few meters below all other components. As for the skirts, they can be of
steel (a few centimeters thick) or concrete (about a meter thick). Their height is determined by
the need to transfer vertical load to competent layers and provide shear against horizontal forces,
and design is similar to piles. In Figure 25, all the issues that must be considered while designing
and installing the platform are illustrated (Dean, 2009).

24
Offshore Foundations: Technologies, Design and Application

Figure 25 – Principal design issues for parallel descent installation: (a) dowel penetration, (b) skirt
penetration, (c) base suction or pressure, (d) dome contact stresses, (e) grouting pressures and
density, and (f) scour protection (Dean, 2009).

3.4. CONCRETE CAISSONS FOR TENSION LEG PLATFORMS


Concrete caissons evolved from deep skirted concrete base foundations, and comprise
individual or clusters of small concrete caissons or “bucket” foundations. Figure 26(a) illustrates
a TLP and its foundation system. This foundation system has the particularity that the resistance
is provided by a combination of concrete self-weight and the interaction between the caisson and
seabed (Randolph et al., 2005). The average static tension is counteracted by the weight of
concrete foundation, while load originating from cyclic waves and wind (design storm) are
transferred to the soil by skirt friction and suction under the top cap (Stove et al., 1992). During a
storm, the TLP will drift out of alignment with the foundations introducing a moment action into
them, which gives the most critical load situation (Christophersen, 1993). Figure 26(b) presents
the limit equilibrium (upper solid line) and three dimensional finite element (lower solid line)
analysis predictions, as well as measured values of bearing capacity and displacement for the
concrete caisson cluster with tension loading (Christophersen, 1993).

25
Offshore Foundations: Technologies, Design and Application

Figure 26 – (a) Inset, Foundation template showing cluster of concrete caissons (b) Predicted and
measured cyclic load vs. displacement response (Christophersen, 1993)

The concrete caisson installation procedure, involves:

 The free release of the concrete caisson(s) onto the seabed, this step is very important to
ensure a minimum penetration by means of the caisson self-weight, so a water tight seal
is provided to withstand the differential pressures applied during the subsequent
application of suction (Stove et al., 1992).
 Water evacuation from the inside of the caisson is commenced by means of pumps located
on its top. As the soil responds under undrained conditions, the caisson skirts are forced
to penetrate the soil due to the internal negative pressure (suction).
 Afterwards the top of the foundation that has not penetrated into the seabed can be
covered with ballast to increase the weight and confinement.
 Finally, the connection and tensioning to the TLP can be made.

The resistance calculation for the skirts is similar to that for piles; from observation the
adhesion coefficient α is in the range of 0.15 and 0.30 depending on the material (Randolph et al.,
2005). The resistance available due to suction has to be analyzed on a site-by-site basis, since it
depends on the soil characteristics, such as permeability, and load magnitude. The greater the
cyclic tension loads, the greater the suctions that will be developed.

Caissons have certain advantages over piles as anchors for deeper water moorings, if they
can provide enough tensile capacity. For example, the pumps used for caisson installation do not
have the same problems as piling hammers at great working depths (even though new systems
are being developed for the latter to allow operations in water depths of 3 km). Also, the larger

26
Offshore Foundations: Technologies, Design and Application

diameter of caisson foundations provides a larger area for ballast and can also mobilize greater
reverse end bearing or passive suction during uplift compared to a pile foundation (Clukey et al.,
1995).

3.5. STEEL BUCKETS FOR JACKETS


Steel buckets (also known as suction cans) are used as an alternative to pile foundations
for jackets. They have also been used extensively for offshore wind turbine foundations. These
suction foundations are steel cylindrical structures, closed on one end and open on the other.
Bucket foundations often exceed 5 m diameter, some reach 10 or 20 meters. The jackets, to which
the buckets are attached, have much larger dimensions, so the installation of these structures can
prove to be quite difficult to perform with the use of only a crane. Therefore, pontoons are used
to bring the structure to site, from which it is launched into the water. Then, the structure dives
by slowly filling the hollow sections with water and the entire structure is aligned with the help
of a crane so it is correctly placed on the seabed.

In the same manner as for concrete caissons, for installation, the open end of the bucket is
placed on the seabed and the water contained within the cylinder and the floor is pumped out.
This creates a vertical load on the structure, pulling it into the ground. Figure 27 illustrates the
three steps of the installation, first the structure touches down on the seabed, then the ballast
containers are filled so the structure penetrates the soil under its own weight, and finally, water
is pumped from the caissons producing suction penetration.

Suction foundations can be applied in sands as well as in clays, and in softer clays they
work very well (Randolph et al., 2005). Depending on which soil the steel buckets are installed,
different failures in the soil around the bucket can occur as discussed below.

Figure 27 – Stages of installation of a bucket Foundation for a jacket structure

27
Offshore Foundations: Technologies, Design and Application

3.5.1. INSTALLATION IN CLAY


One of the failure mechanisms is that when there is a lot of friction on the outside of the
can, too high a suction is developed and plastic failure of the soil occurs. In this case, instead of
penetration of the can, the soil can be sucked up into the can instead. This phenomenon is called
reverse end bearing failure. Normal end bearing failure occurs when after installation, too large a
load is placed on the soil inside and right underneath the can. The reverse can happen if either the
suction applied or the tension load on the can is too large.

3.5.2. INSTALLATION IN SAND


In sands a different failure mechanism can occur, and it is related to the inflow of water.
In Figure 28 the flow net into the suction can has been sketched; during the suction penetration
phase, pumping creates an under-pressure across the foundation baseplate and, and more
importantly, sets up seepage flow that reduces tip resistance and internal skirt friction. The inflow
will always be present because of the pressure differential between the sea and the inside of the
can and the fact that sands are relatively permeable. If the suction produces a very fast flow, a
large flow gradient can occur, which will decrease effective stresses and ultimately liquefaction
can occur. If this does not occur in all the cans at the same time, accidents may happen.

Figure 28 – Seepage pressures set up during suction installation of bucket foundation (Erbrich &
Tjelta, 1999).

4. DEEP AND ULTRA-DEEP WATER FOUNDATIONS


The demand for oil products and natural gas, has forced companies to search for resources
in increasingly remote sites. Many of these sites are offshore and have water depths in excess of
1000 m, Figure 29. Water depths in excess of 500 m are considered to be “deep”, and “ultra-deep”
when greater than 1000 m. The economic investment associated with developing these sites is
huge, so it is vital to produce solutions with an optimal balance between reliability and economy.

28
Offshore Foundations: Technologies, Design and Application

Clearly, at these great depths it becomes increasingly impracticable to build load transfer
structures such as jackets or gravity based structures. Therefore, different foundation solutions
have been adopted, i.e. anchors with mooring systems.

There is a vast range of solutions which are sub-divided between gravity anchors and
embedded anchors.

Gravity anchor types include:

 Boxes
 Grillage and Berm

Embedded anchor types include:

 Anchor piles
 Suction caissons
 Drag anchors (fixed fluke)
 Vertically loaded drag anchors (VLA)
 Suction embedded plate anchors (SEPLA)
 Dynamically penetrated anchors (DPA)

Figure 29 – Locations of some current deep-water developments (Dean, 2009)

Many current deepwater developments are close to the continental rise, and so are subject
to additional potential geohazards associated with possible land sliding (Dean, 2009). These
aspects are presented and discussed in Section 5.

The mooring system plays an important role in deep and ultra-deep waters, since it is not
viable to build and install load transfer structures from the waterline onto the seabed in such deep

29
Offshore Foundations: Technologies, Design and Application

waters. To overcome this issue, several types of anchors were created to be attached to mooring
systems, there are two main types of mooring:

1. Catenary mooring systems are generally used in shallow to deep water (up to 1000
metre). Chain or wire rope mooring lines are used, of which a significant length lies on the
seabed, and the anchor is loaded in a horizontal direction. Typically conventional drag
embedment anchors are used in these systems. Figure 30(a) shows the geometry of a
catenary mooring system and a common drag anchor in profile.
2. Taut leg mooring systems are typically used in deep and ultra-deep water (greater than
1000 metre). Mooring lines used are light weight (synthetic rope or wire rope), and they
enter the seabed at a significant angle. The anchor is loaded in the horizontal and vertical
direction. Vertical loaded anchors (VLA) are commonly used, although there are diverse
solutions. Figure 30(b) illustrates the straight alignment of the taut leg mooring system
and an exemplar of a VLA.

Figure 30 – (a) Catenary mooring and drag embedment anchor. (b) Taut leg mooring and VLA.
From: http://events.energetics.com/deepwater/pdfs/presentations/session5/roderickruinen.pdf

4.1. GRAVITY ANCHORS


An anchoring system is normally required to provide resistance forces that are primarily
horizontal, with cyclic as well as static components. Gravity anchors consist of heavy weight steel
structures (box, or grillage), filled or covered with granular fill (either rock-fill, or heavier material

30
Offshore Foundations: Technologies, Design and Application

such as iron ore), and placed on the seafloor. Simply, the structural element is placed first, and
then the bulk fill is added (Randolph et al., 2005; Dean, 2009).

Two different structures are represented in Figure 31, on the left is a conventional box
anchor filled with iron ore, which provides ballast, and on the right a covered grillage. The latter
is considerably more efficient in terms of quantity of steel for a given holding capacity, but is much
less efficient in terms of the quantity of ballast required. Design of this type of anchor is also more
complex since a variety of failures mode must be considered, ranging from sliding of the complete
berm, pulling out of the grillage, or combinations involving asymmetric mechanisms (Randolph et
al., 2005).

Figure 31 – Gravity anchors (Randolph et al., 2005)

4.2. PILE ANCHORS


Pile anchors have a similar behaviour to pile foundations (i.e. little skin friction developed
in calcareous soil), but the construction method and the forces they need to withstand are
different. They are very effective in many soils, and can either be drilled in and grouted using an
offshore mobile drilling rig, or driven in with an underwater hammer. Advances have been made
to allow hydraulic hammers to work in deep waters and with greater power, so the driving of the
piles does not become an issue (Gerwick et al., 2007).

The anchor pile system consists of a mooring chain or cable, and the pile. Figure 32(a)
shows a simple system of an anchor pile and chain. In a catenary mooring system, the chain is laid
along the seabed describing a smooth curve, and the anchoring force that it provides includes the
weight of the line, the friction on the seabed, and the frictional resistance from the soil on the
buried part of the anchor line, as well as the pull-out resistance of the pile itself. The soil resistance
along the length of a buried chain or cable can be a significant proportion of the overall anchoring

31
Offshore Foundations: Technologies, Design and Application

resistance provided by the system. Figure 32(b) shows the forces on an element of the anchor line
in a catenary mooring system. In a taut mooring system, the line is tensioned, and rises from the
seabed without passing along the seafloor.

The concept of an anchor pile starts with a pad-eye that is attached to a pile and a line is
attached to the pad-eye, Figure 32(a). The location of the pad-eye is designed with the purpose of
reducing potential rotation of the pile when loaded. Then the pile is driven into the seabed, and
may be left protruding slightly above the seafloor, so as to be retrieved later. After installation, the
chain is attached to the floating platform and tightened. The most difficult anchoring soil of all is
a soft mud, silt, or loose sand overlying a hard material such as conglomerate or very dense sand
and silt. So, a conventional pile may be placed in holes excavated by clamshell bucket and then
back filled with dumped rock (Gerwick et al., 2007). However, this operation gets increasingly
difficult with larger water depths.

Figure 32 – (a) Anchor pile and chain. (b) Forces on an element of the anchor line (Ruinen, 2005)

4.3. SUCTION CAISSONS


Although concrete caissons have been used, the majority of suction caissons are fabricated
from steel, which have a similar concept to steel buckets on shallow foundations. Suction caissons
operate as anchors, and vertical capacity is granted by the weight of the plug of soil inside and the

32
Offshore Foundations: Technologies, Design and Application

friction on the outer surfaces, and in addition, the characteristic negative end-bearing resistance.
The latter, as in a steel bucket, is the force required to separate the lower end of the soil plug from
the undisturbed soil.

Typically, suction anchors are open at the bottom and closed at the top. They have large
diameters, typically more than 5 meters in diameter and are 20 to 30 meters in length, with a
length to diameter ratio (L/d) in the range of 3 to 6. Normally the cylinders have very high ratios
of diameter to wall thickness (d/t ~100 to 250), that require internal stiffeners to prevent
structural buckling during installation, and due to the large lateral loads imposed by taut
moorings. Mooring loads are applied by an anchor line attached to the side of the caisson at a
depth that optimises the holding capacity. Usually this requires the line of action of the load to
pass through a point on the axis at a depth of 60% to 70% of the embedded depth. Figure 33
illustrates the optimal depth (D*) of the padeye, from which it is possible to realise that a taut wire
does not require as deep a padeye as a catenary mooring.

Figure 33 – Mooring appliance point and variation of padeye depth with loading angle for given
center of rotation (Randolph & House, 2002)

Suction caissons have an identical installation process to buckets and concrete caissons.
The four major installation steps are illustrated in Figure 34. Firstly, the caisson is delivered to
the required location, and then it is initially penetrated into the seabed under self-weight with the
top-vent opened. Afterwards, the remaining penetration is completed by pumping water from
inside the caisson, using demountable pumps connected to a valve in the lid and operated by ROVs.
When the desired skirt penetration is achieved, the pumps are stopped and the vent is closed to
stop water flow into the interior. This allows internal suctions to develop under vertical loading
(uplifting), hence maximising the end-bearing resistance (Randolph et al., 2005).

33
Offshore Foundations: Technologies, Design and Application

Figure 34 – Suction Caisson Installation Steps, from:


http://www.epd.gov.hk/eia/register/report/eiareport/eia_1672009/HKOWF%20HTML%20EIA/
Main%20Text%20figures/Fig%202.36a.png.

4.4. VERTICALLY LOADED DRAG ANCHOR


High capacity drag anchors evolved from conventional ship anchors. Traditionally, drag
anchors comprise a broad fluke rigidly connected to a shank, as shown in Figure 35(a). The angle
between shank and fluke is pre-determined, though may be adjusted prior to anchor placement
on the seabed. This angle is typically around 50° for clay conditions and 30° in sand or where clay
of high strength occurs at the seabed. For installation, the anchor is positioned on the seabed
correctly orientated and it is then embedded by pre-tensioning the chain to an appropriate proof
load. Depending on soil conditions, penetration depths usually range from 1 to 5 fluke lengths
(typical fluke lengths being 1 to 8 m), and anchors can be dragged through a distance of 10 to 20
times the fluke length, typically a holding capacity of 20 to 50 times the anchor weight is mobilized.

34
Offshore Foundations: Technologies, Design and Application

Figure 35 – Drag anchors, (a) fixed fluke anchor. (b) Vryhof Stevmanta VLA. Source: Vryhof anchor
manual (2005)

Vertically loaded drag anchors (VLA), also known as drag-in plate anchors, were
developed to overcome the existing limitations on fixed-fluke anchors, which could not withstand
significant vertical load components at the seabed. Actually, fixed-fluke anchors are removed by
applying vertical load to the anchor chain. Therefore, common drag anchors cannot be used for
deep-water foundations using taut or semi-taut polyester rope moorings.

The VLA is similar to the conventional drag anchor except the fluke is a plate with a much
slender profile and the shank is replaced by either a much thinner shank or a wire harness, Figure
35(b). It is installed like a conventional drag anchor with a horizontal chain load at the mudline
and then different mechanisms are used to allow the fluke to rotate until it is perpendicular to the
applied load [angle adjuster in Figure 35(b)], mobilizing the maximum possible soil resistance,
and enabling the anchor to withstand both horizontal and vertical loading. There are two different
installation methods; it can be either single line or double line installation. Figure 36 illustrates
both of the installation methods. For each installation method a recovery method for the VLA is
available, typically by detaching the front chains or wires and pulling the anchor in the opposite
direction to the installation direction with a fraction of the installation load. Figure 37(b) shows
the VLA being pulled out of the seabed after detaching the front mooring.

35
Offshore Foundations: Technologies, Design and Application

Figure 36 – Installation methods

Figure 37 – (a) Anchor depth determination. (b) VLA simple recovery, Vryhof anchor manual
(2005).

VLAs achieve the desired effect of installing a plate anchor at a sufficient depth below the
seabed in order to resist the mooring loads, but there is an inherent problem with these kinds of
anchors in knowing exactly where they are in the soil. Usually, the depth is determined by
measuring the angle of the anchor line makes with the seabed, and the length of wire in the soil,
but for many reasons this angle may not correspond to reality, as it may have a differential angle
along its embedded length. In Figure 37(a), a scheme to determine the approximate anchor depth
is indicated.

Initially, VLA utilization was predominantly for semi-permanent moorings, for example,
for mobile offshore drilling units (MODUs). Thus, it can also be considered a temporary
foundation. Nowadays, VLAs are commonly used to moor permanent floating facilities, such as a
FPSO in the Campos Basin, Brazil, which is located in 1600 m of water with a taut-line mooring
system secured to the seabed by nine 14 m² Vryhof Stevmanta VLAs.

36
Offshore Foundations: Technologies, Design and Application

4.5. SUCTION EMBEDDED PLATE ANCHOR


A new system, called a suction embedded plate anchor (SEPLA), was developed to
overcome the problems of the conventional plate anchor (e.g. VLA), achieving greater and more
precise depth location below the seabed (Dove et al., 1998; Wilde et al., 2001).

The SEPLA uses a suction caisson (or “follower”) to embed a rectangular plate anchor,
providing a known initial penetration depth for the anchor, at a specified geographical location.
The components of a SEPLA are illustrated in Figure 38.

Figure 38 – Components of a suction embedded plate anchor (Gaudin et al., 2006)

SEPLA installation consists of 3 steps: caisson penetration, caisson retraction, and anchor
keying. These steps are shown schematically in Figure 39. First, the caisson with a plate anchor
slotted vertically in its base is lowered to the seafloor and penetrated into the soil under its dead
weight until the skin friction and end-bearing resistance of the soil on the caisson equal the
caisson’s dead weight. The vent valve on the top of caisson is then closed and the water trapped
inside is pumped out. The ensuing differential pressure at the top drives the caisson to the design
depth. The plate anchor is then released and the water is pumped back into the caisson, causing
the caisson to move upward, leaving the plate anchor in place in a vertical orientation. The caisson
is retracted from the seabed and prepared for the next installation. As the anchor chain is
tensioned, it cuts into the soil. Simultaneously, the anchor line applies a load to the anchor’s offset
padeye causing it to rotate or “key”. In order to achieve the maximum mobilized capacity, the plate
must be as close to perpendicular to the direction of loading as possible (Yang et al., 2012).

37
Offshore Foundations: Technologies, Design and Application

Figure 39 –Schematic of SEPLA installation (Yang et al., 2012; and courtesy of InterMoor).

SEPLA installation accuracy represents a great improvement over that for drag
embedment anchors, however two questions emerge (these questions are applied to all offshore
plate anchors such as VLAs). Firstly, the caisson penetration and anchor keying provokes a
disturbance in the soil mass around the SEPLA, which leads to a decrease of the soil strength in
the region. Secondly, when keying is being initiated, a loss of embedment depth occurs. While, the
first question can be solved as the soil strength is largely recovered over time by soil
reconsolidation, the second problem cannot because loss of embedment depth is permanent. This
is a very important issue, since SEPLA capacity significantly depends on its embedment depth
when the soil has increasing strength with depth (which is a typical in the offshore environment).
Therefore, it becomes very important to accurately estimate the loss of embedment depth during
the keying process. This estimate can then be factored into the design; Wilde et al. (2001) report
upward movements ranging between 0.5 and 1.7 times the plate height, which is a wide range
when plate heights of 2.5 m to 4.5m are used in practice.

Even though the undrained capacity of plate anchors has been extensively investigated by
means of analytical and experimental methods; for SEPLA, there are a limited number of reported
studies and therefore the keying process is not yet well understood. However, Yang el al. (2012)
present a theoretical model to predict the trajectory and corresponding capacities of SEPLA
during the keying process based on plastic limit analysis.

4.6. DYNAMICALLY PENETRATED ANCHOR


As offshore exploitation moves to water depths of around 3000 m, new technologies have
had to be developed in order to reduce installation costs, and facilitate construction. Moreover,
the high number of floating production and drilling units in operation may provoke the congestion
of the sea bottom due to the high number of risers and mooring lines employed. In this scenario,
dynamically penetrated anchors (DPA), and in particular Torpedo anchors, have proven to be a

38
Offshore Foundations: Technologies, Design and Application

reliable alternative used in Brazilian offshore fields (Aguiar et al., 2009). The reduced mooring
line radius employed on torpedo anchors relative to catenary mooring systems with drag anchors,
reduces sea bottom congestion, Figure 40.

Figure 40 – Radius comparison between floating units linked to conventional drag anchors and
torpedo anchors, from: http://www.hindawi.com/journals/jam/2012/102618.fig.002.jpg.

Torpedo anchors (TA) are the most applied type of DPA and they have been developed by
the Brazilian oil company Petrobras. TAs are cone-tipped, cylindrical steel pipes filled with
concrete and scrap metal. They penetrate the seabed relying on the kinetic energy they acquire
while free falling from heights of between 30 m and 150m above the seabed. Torpedo anchors
come in various sizes from 0.76 m to 1.07 m in diameter, 12 m to 17 m in length, and 241 kN to
961 kN in weight. The inside of the anchor shaft is filled with ballast to increase the weight and
maintain the centre of gravity below the centre of buoyancy for stability. Some versions of the TA
have been fitted with 4 flukes at the trailing edge, ranging in width from 0.45 m to 0.9 m, and 9 m
to 10m long (Raie, 2009; Medeiros et al.,1997, 2001, 2002). Two different DPAs, with and without
fins are pictured in Figure 41(a).

Torpedoes can easily reach velocities of 25 m/s to 35 m/s at the seabed after being
released from a height of 20 m to 40 m above the seabed, allowing tip penetrations up to 3 times

39
Offshore Foundations: Technologies, Design and Application

the anchor length and holding capacities after consolidation that are expected to be in the range
of 5 to 10 times the weight of the anchor (Randolph et al., 2005).

Figure 41 – Dynamically penetrating anchors (a) Torpedo anchor with fins and without fins
(Medeiros, 2002); (b) installation of 4 flukes torpedo anchor (Medeiros, 2002; O’Loughlin et al.,
2004)

The installation procedure for DPA has developed from its original method. Instead of
using only one anchor-handling vessel (AHV) to lower the anchor to a predetermined height above
the seabed, using the permanent mooring line, now two AHV are used. The installation process
was modified to minimize the effect of drag force on the mooring line on the free falling motion of
the anchor. Accordingly, the anchor is lowered using an installation wire from the first AHV while
the second AHV holds the permanent mooring line that is attached to the anchor and forms a loop.
A remote release system is used at the end of installation wire to release the anchor (Araujo et al.,
2004). A chain segment is recommended for the lower portion of the mooring line because model
tests of the anchor installation (Lieng et al., 2000) have shown that chain drag does not reduce the
velocity of the anchor during free fall. Figure 41(b) demonstrates the lowering of two model scale
torpedo anchors to position them before free-fall releasing. A full scale torpedo pile and the
situation immediately prior to TA release, in which it is possible to see the loop between the
permanent mooring line and the installation line, is illustrated in Figure 42.

40
Offshore Foundations: Technologies, Design and Application

Figure 42 – Full scale torpedo pile and releasing situation, Lieng et al. (1999)

The main reason for using this type of mooring solution is its simplicity and speed of
installation. With regard to the equipment required for installation, the torpedo anchor
installation is depth-independent. Moreover, torpedo piles are cost-effective throughout
fabrication, transportation, and installation. Fabrication is easy and inexpensive due to the simple
design of the torpedo anchors. The cost of transportation is low because the compact design of the
torpedo anchor allows more anchors to be transported per voyage of the AHV than, for example,
suction caissons. Also, the installation is economical because an external source of energy is not
required for installation and a quick installation is possible using one or two AHVs and limited use
of ROVs. Finally, the predicted holding capacity is less dependent on the precise evaluation of the
soil shear-strength profile. Since higher strength profiles reduce the penetration and lower
strength profiles increase penetration, therefore the holding capacity is mainly a function of the
kinetic energy obtained during free falling. Nevertheless, torpedo anchors have the disadvantage
of the uncertainty in verticality of the anchor, which affects the holding capacity (O’Loughlin et al.,
2013; Raie, 2009).

Figure 43 presents a table, where the advantages and disadvantages of the suction caisson,
VLA, SEPLA and torpedo anchor systems discussed in this work are summarised.

41
Offshore Foundations: Technologies, Design and Application

Figure 43 – Advantages and disadvantages of different deep water anchor types (Ehlers et al., 2004).

42
Offshore Foundations: Technologies, Design and Application

5. GEOLOGICAL CHARACTERIZATION
Before exploration, governments, energy companies and academic institutions study what
seem to be the most likely areas to look for evidence of energy sources. When the objective is to
find hydrocarbon resources, the geology of the area gives indications of the likelihood of the
existence of hydrocarbon traps and other oil and gas sources. But if the objective is renewable
energy then metocean data gives information of certain locations where the appropriate
environmental conditions occur, e.g. areas where strong winds constantly blow will be
appropriate to install wind turbines. If potentially commercial resources are found, the
government will determine the boundaries of offshore exploration lots, and tender exploration
licenses.

In exploration for hydrocarbon sources, geophysical surveys of licensed areas are run to
determine whether and where the oil or gas is likely to be. These surveys typically penetrate
several kilometers into the seabed. For renewable energy, the focus is on wind, wave, current,
or tidal characteristics. Once a resource has been found and government consents have been
obtained, further survey work is carried out to determine engineering and other design
conditions and parameters (Dean, 2009). Table 4 lists some of the surveys that may be done.

Table 4 – Surveys and their purpose (Dean, 2009)

Survey Purpose
To measure water depths, map the seafloor, identify seafloor hazards
Bathymetry survey
such as unevenness, slopes, fluid expulsion features, and collapses
To determine wind, wave, and current characteristics at the platform
Metocean study sites; these will form the basis of estimates of platform loading and
seabed scour
To identify environmental issues and measure environmental flora and
Environmental
fauna populations, so that the site can be returned to its original
baseline survey
condition after the production operations have finished
Geohazards To identify and plan the mitigation of geological and geotechnical
assessment hazards
Shallow geophysical To identify soil layering and sub-bottom hazards such as geological
survey faults, in-filled ancient riverbeds, voids, shallow gas, and rock
To verify soil layering and determine soil properties at the platform
Geotechnical survey
sites

43
Offshore Foundations: Technologies, Design and Application

To identify bathymetry, soil conditions, and hazards along pipeline


Pipeline or cable
routes between different offshore platforms, and between platforms
route surveys
and onshore
Seismic risk To identify the seismicity of the area and determine spectra and/or
assessment acceleration time histories to be used in seismic design
Done just before installing a platform at a site, to check that the seafloor
Seafloor survey has not changed and hazards such as dropped objects, sand dunes,
shipwrecks have not occurred there

In order to evaluate, which foundation solution is best for each site, the geological
characterization must answer important questions such as:

 What is the water depth?


 What is the composition of the mudline?
 What soils do we find?
 What are the layer thicknesses?
 What geohazards do we find there?

5.1. TOPOGRAPHICAL FEATURES OF OCEAN FLOORS


Figure 44 illustrates a typical marine topography of the ocean bottom, the features
common to all oceans: the continental margin, the continental rise and the abyssal plain of the
deep ocean. Trenches and seamounts (oceanic ridge or rise) may also be present in the deep
ocean. The continental margin comprises the continental shelf and slope. The shelf is the
submerged continuation of the adjacent land. The seaward extent of the continental shelf is the
shelf break, or the continental ridge, leading into the continental slope. The toe of the continental
slope is marked by the continental rise leading into the abyssal plain. The continental margins
(covering approximately 20% of the total ocean floor) are extremely important as oil reservoirs
and as such, are of most interest to engineers concerned with harnessing offshore oil and gas
resources.

44
Offshore Foundations: Technologies, Design and Application

Figure 44 – Topographical features of ocean floors, after


http://www.marinebio.net/marinescience/01intro/woimg/xsectopo.jpg.

While ocean topography can be described in general terms, regional variations naturally
exist. Figure 46 shows a shaded relief map of land topography and ocean bathymetry compiled by
the United States National Geophysical Data Centre, it also shows the amplification of Southern
Africa, where is possible to distinguish the continental shelf and slope, as well as the abyssal plain.
In addition, Figure 45 shows the topographical profile of the continental margin in a selection of
offshore locations, illustrating the potential diversity of shelf and slope bathymetry. These
locations are indicated on the world map in Figure 46.

Figure 45 – Profiles of ocean floor topography at selected locations, Randolph et al. 2011

45
Offshore Foundations: Technologies, Design and Application

Figure 46 – Bathymetric shaded relief map showing a detailed extent of continental margins
(number in red relates to locations in Figure 45), US National Geophysical Data Centre

46
Offshore Foundations: Technologies, Design and Application

The continental shelf can be virtually non-existent or extend several hundred kilometers
offshore with the shelf break occurring between depths of 10 m and 500 m. Shelf breaks are
deepest off glaciated areas and shallowest in areas with extensive coral growth. The average slope
of the continental shelf is approximately 1:500 (0˚07’). Beyond the shelf break, the continental
slope has a steeper gradient than the shelf, ranging from an average 1:40 (1.2˚) in delta regions
but up to 1:10 (6˚) in faulted areas, and reaches water depths of 2000 m to 3000 m. The
continental rise, at the foot of the continental slope, has gradients ranging from 1:1000 and 1:700.
The abyssal plain adjacent to the continental rise is smooth with gradients between 1:1000 and
1:10000, and occurs from depths of 2500 m to 6000 m.

5.2. SEABED GEOLOGY


In the majority of the ocean bottom, the first layers are thick compositions of marine
sediments; Figure 47 shows a digital model compiled by the U.S. National Geophysical Data Centre
showing the distribution of sediment in the world’s oceans and marginal seas. These sediments
are clearly thicker near continents, and thinner on newly formed mid-ocean ridges. In some areas,
strong bottom currents are responsible for cleaning away any sediment, avoiding its settlement
and consolidation. Even though, the continental margins cover only 20% of the seabed area, they
contain nearly 75% of the total marine sediments. In many places, the continental rise is a
depositional feature, formed mainly of sediment slurry, and reaching up to 1.6 km thickness.
Canyons often cut across the rise and act as channels for the seaward transport of sediment.
Abyssal plains are connected by canyons or other channels to landward sources of sediments,
which are transported as dense slurries to the plains (Poulos, 1988; Randolph et al., 2011).

47
Offshore Foundations: Technologies, Design and Application

Figure 47 – Sediment thickness of the world’s oceans and marginal seas. Source:
http://www.ngdc.noaa.gov/mgglsedthick/sedthick.html.

5.2.1. SEABED SEDIMENTS ORIGIN AND CLASSIFICATION


Marine sediments are composed of detrital material either from land or the remains of
marine organisms which, now, leads to the principal classification of sediment as terrigenous
(transported from land) or pelagic (sediments that settle through the water column). Pelagic
sediments are deposited so slowly that near shore and coastal areas are overwhelmed by
terrigenous deposits.

On one hand, terrigenous material comes from sediments carried by the flow of rivers,
coastal erosion, aeolian or glacial activities. The size of the soil particles is used as the basis for
describing terrigenous sediments, and then sub-categorizing them. These sediments tend to be
grains of silicate-based minerals such as quartz and feldspar, and are principally formed from the
erosion of rocks – leading to the term lithogenous.

On the other hand, pelagic sediments are generally fine grained and are instead classified
according to their composition. Organic, or biogenous, pelagic sediments derive from the
insoluble remains of marine organisms, e.g. shells, skeletons and teeth. Lithogenous pelagic
sediments are formed when particles are transported by wind into the ocean, before settling
through the water column. Figure 48 illustrates a schematic cross-sectional view of the
continental margin which shows various sediment types and their distribution across an idealized

48
Offshore Foundations: Technologies, Design and Application

passive margin. Passive margins are the margins which not extend down to subduction zones
(active continental margins).

Marine sediments can also form from biological and chemical reactions occurring in the
water column or within sediments, and are referred to as hydrogenous sediments. Figure 49
summarises some of the processes involved in the formation of marine sediments.

Figure 48 – Distribution of sediment across a passive continental margin, from:


http://classconnection.s3.amazonaws.com/927/flashcards/68927/jpg/4-91305062763712.jpg.

Figure 49 - Sedimentary process of marine deposits (after Silva 1974)

49
Offshore Foundations: Technologies, Design and Application

Lithogenous Sediments
The number one suppliers of oceanic sediments are rivers, as they contribute 20 billion
tonnes of sediment to the world’s oceans every year. The settlement of particles is controlled
primarily by grain size, such that generally finer particles are found with increasing distance from
shore, since these particles are lighter they remain suspended in the air and water longer and
travel further than the heaviest ones. However, lithogenous sediments can also come from
material transported during submarine slides such as debris flows and turbidity currents (heavy
fluid flows). The latter, in particular, can transport sediment from the continental margins to the
continental rise and onto the abyssal plain, Figure 50. Deposits formed by turbidity currents have
graded bedding, with larger particles overlain by progressively smaller particles, Figure 50(c)
shows this phenomena.

Wind also plays an important role, transporting around 100 million tonnes of sediment
into world’s oceans annually, and small particles can be carried for considerable distances before
being deposited in the sea as pelagic sediments (Keuen et al, 1950; Randolph et al., 2011).

Figure 50 – Turbidity current: (a) seabed topography where slope break may start a turbidity
current (b) turbidity current progress (c) settlement turbidity particles. From:
http://people.maths.ox.ac.uk/fay/research.html.

50
Offshore Foundations: Technologies, Design and Application

Biogenous Sediments
Biogenous sediments may either be siliceous or calcareous and are formed from
planktonic organisms, which are the most abundant organism in the oceans (now and through
geological history). The insoluble shell of these creatures can be formed of silica or calcite. While,
silica organisms are more common in the Polar Regions or in the very deep waters of the equator,
calcite organisms are more common in shallow water and temperate tropical climates. Calcium
carbonate, the main constituent of calcareous deposits, is soluble at high pressure, that’s why they
are not found in water depths greater than 4000 m (Randolph et al., 2011). Figure 51 shows the
worldwide distribution of neritic (nearshore) and pelagic (open ocean) sediments, it is conclusive
that neritic deposits are dominated by lithogeneous materials while pelagic deposits are
dominated by various types of biogenous oozes and lithogeneous abyssal clay.

Figure 51 – Distribution of the different sediments deposits across the world, from:
http://classconnection.s3.amazonaws.com/927/flashcards/68927/jpg/4-91305062763712.jpg.

All these pelagic sediments become from remains of ancient micro-organisms, such as
shells or skeletons (typically from an organism called foraminifera), and they are often perforated
and angular. These remains when coupled with the relative softness of calcium carbonate (of
which they are made), leads to the fragility and high compressibility which are characteristic of
calcareous sediments. Figure 52 compares micrographs of calcareous sand, from Goodwyn on the
North-West shelf of Australia, with quartz-grained silica sand, showing clearly the difference in
particle shape. Calcareous sand has fragile, angular and hollow particles as opposed to the hard,
rounded silica sand grains.

51
Offshore Foundations: Technologies, Design and Application

Figure 52 – Micrographs of (a) calcareous sand and (b) silica sand (Dean, 2009).

5.2.2. GEOTECHNICAL CHARACTERISTICS OF SOME OFFSHORE REGIONS


As a consequence of what was explained before, it is possible to say, in broad terms, that
finer-grained sediments predominate further from shore and in deeper water. Because coarser
terrigenous sediments cannot be transported this far, therefore pelagic sediments predominate.
Certain geohazards are also more prevalent in deep water, including steep scarps that are often
found at the margins of the continental shelf.

Even though, many factors may affect the characterization of a specific area, in broad
terms, some common seabed conditions in the major areas of oil and gas exploration can be
characterized as (on the following list, the numbers at the end of each description refer to the sites
on Figure 53), (Dean, 2009):

 Gulf of Mexico – soft, normally consolidated, medium high plasticity clays (30<PI<70),
often with interbedded sand layers (sites 11, 21 & 22).
 Campos and Santos Basins offshore Brazil – sands and clays with high carbonate content
(site 5).
 Western Africa – soft, normally consolidated, very high plasticity clays (70<PI<120), often
deposited rapidly by river deltas (sites 1 & 10).
 North Sea and other glaciated regions – stiff, overconsolidated clays and dense sands, with
a recent drape of softer material (site 14).
 South-East Asia – desiccated crusts of stiff soil, which are remnants of low sea levels during
the Pleistocene (2,588,000 to 11,700 years ago), with strengths one or two orders of
magnitude greater than the underlying soil (site 20).

52
Offshore Foundations: Technologies, Design and Application

 Australia’s North-West and Timor Sea- carbonate sands, silts and clay-sized soils, often
with variable cementation (site 13).

Figure 53 – Worldwide distribution of offshore oil and gas developments. Dean (2009) based on
McClelland (1974) and Poulos (1988).

5.3. GEOHAZARDS
By definition, a geohazard is a geological and environmental condition, and involve long-
term or short-term processes, that can lead to the movement of soil, rock, fluid or gas during
sudden episodic events or slow progressive deformations (Randolph et al., 2011). For offshore oil
and gas projects, geohazards have the potential to cause injury or loss of life, damage to the
environment or infrastructure, and can impose significant additional project costs.

Diverse geohazards are associated with engineering on the ocean floor and they become
significantly more dangerous in deeper water. Geohazards are grouped in two general categories:

 Hazardous events – events that are infrequent and episodic in nature, such as phenomena
associated with earthquakes, submarine slope movements, turbidity flows and gas
expulsions.
 Hazardous ground conditions – conditions that involve slow processes that are
progressive in nature, such as soil creep, non-tectonic fault creep and mud or salt
tectonics.

53
Offshore Foundations: Technologies, Design and Application

Geohazards may pose a threat to the integrity or serviceability of structures and their
foundations over their design lifetime. Therefore, with the purpose of their identification on the
site; studies of geology, geomorphology, and geography of a region, and thorough geophysical and
geotechnical surveys and investigations are executed (Dean, 2009).

Figure 54 illustrates the most typical geohazards to be considered, including many


geological features as well as landslides, carbonate sands, unconsolidated soils, gas hydrates, and
disturbed sediments. Submarine slope instabilities are discussed further in the following section,
since these are the common geohazard that needs to be dealt with.

Figure 54 – Schematic diagrams showing main offshore geohazards (a) widely used summary of
deepwater geohazards (Power et al., 2005) (b) more geohazards (Strout and Tjelta, 2007).

54
Offshore Foundations: Technologies, Design and Application

5.3.1. TRIGGERING MECHANISMS FOR SUBMARINE SLOPE FAILURES


There are a several prospective events capable of triggering submarine slope instabilities,
flow slides, debris flows, and mudflows and such events can be either sudden, or progressive.
Natural processes or anthropogenic actions cause an increment in soil stresses or decrease in soil
strength, leading to failure of a soil mass. The comprehension of pore pressure conditions and the
processes and mechanisms that lead to excess pore pressure generation are very important for
assessing potential geohazard triggering mechanisms (Randolph et al., 2011). Kvalsta et al. (2001)
identified a selection of triggering mechanisms (some events are illustrated in Figure 54b):

A) Natural Processes
 Rapid deposition leading to excess pore pressures, under-consolidation and
increased shear stresses in a slope
 Base erosion or top deposition leading to higher slope inclination and increased
gravity forces and shear stresses along potential failure surfaces
 Melting of gas hydrates caused by temperature increase or pressure reduction
leading to increased pore pressure and reduced soil strength
 Active fluid or gas flow and expulsion
 Mud volcano eruptions giving rise to mass wasting and soil displacements
 Tectonic fault displacements generating earthquakes, near-field displacement
pulses, and ground rupture
 Earthquake strong ground shaking causing short-term inertia forces and increase
in pore pressure
 Long wavelength wave loading
 Sea level lowering during glacial periods leading to lower hydrostatic pressure,
free gas expansion and gas hydrate ex-solution
 Increased seawater temperature at seabed level caused by changes in current
regime leading to temperature increase in the soil mass and ex-solution of
hydrates.
 Sensitive (contractive) and collapsible soils increasing the risk of retrogressive
sliding and increased areal extent of failure zones.
B) Human Activities
 Drilling wells, creating blowouts and cratering at the seabed
 Underground blowouts changing the pore pressure regime in shallow layers and
potentially creating instability in sloping areas

55
Offshore Foundations: Technologies, Design and Application

 Oil production increasing heat flow and temperature around wells leading to
hydrate ex-solution, increased pore pressures and strength loss of the adjacent
soil
 Depletion of reservoir pressure giving rise to reservoir collapse and changes in
overburden stresses
 Installation activities leading to increasing gravity forces
 Mooring installations and anchoring forces imposing short and long-term lateral
forces.

Marine geohazards generally involve mobilisation of the seabed and sub-bottom


sediments. This mobilisation may impact against or bury infrastructure or lead to loss of
foundation support. The volume of soil involved may range from a few cubic metres to thousands
of cubic kilometres with consequences ranging from local over-stressing of subsea infrastructure
to total loss of an installation with the associated human casualties and economic and
environmental impacts. If a very large seabed displacement occurs it has the potential to generate
a tsunami, which can cause massive human, economic and environmental losses. Therefore
platforms are designed to avoid generating a slope failure, and to resist the forces from turbidity
currents and debris flows generated elsewhere (Dean, 2009; Randolph et al., 2011).

5.3.2. GEOHAZARD IDENTIFICATION


Due to the large and complex nature of offshore projects, they often span a range of
different environments, extending from deep water on the continental rise, up the continental
slope, across the continental shelf through shallow water to the shoreline. Therefore, major
projects are exposed to a wide range of geological and geotechnical conditions, so these projects
evolve through time and may go through many design concepts and engineering stages.

Projects involve three phases, moving from the general to the specific; Figure 55
summarises these phases. First, pre-exploration interpretations of potential site conditions are
carried out, then post-discovery preliminary engineering evaluations are made, and finally, an
integrated site characterisation is performed in order to support detailed design.

The geological and geotechnical information collected during the development of projects
needs to be sufficient in order to support each stage in a project. As a result, a phased approach
often proves most effective in meeting engineering requirements for offshore developments
(Randolph et al., 2011).

56
Offshore Foundations: Technologies, Design and Application

First Phase. Pre-Drilling Second Phase. Post- Third Phase. Integrated Site
Activities Discovery Characterisation
Seismic inversion and
Screening of regional Preliminary engineering development of final
geological hazards evaluation geotechnical criteria

Assess Geohazards in Plan high-resolution Detailed geohazard assessment,


prospect area geophysical survey analyses for special engineering
programme issues

Assess hazards for Carry out high-resolution


specific well sites geophysical survey Risk assessment
programme
Prepare and process
Team meetings and Develop model with integrated
high-resolution
reporting site characteristics
geophysical data

Complete preliminary Prepare integrated


site characterisation report

Plan geotechnical site Team meetings and


investigation reporting

Carry out geotechnical


investigation

Sample preparation and


shipping to laboratories

Geological lab testing

Geotechnical lab testing

Team meetings and


reporting

Figure 55 – Summary of generalised investigation elements (Campbell et al. 2008)

The Phase 1 works essentially as a desktop study, to evaluate general constraints from
conditions such as extreme terrain, earthquake and fault activity, slope instability and broad
geotechnical soil properties. Generally this first stage involves: compilation and review of
published or unpublished data and reports, interpretation of exploration seismic testing results;
development of a project geographic information system (GIS), identification of critical
engineering issues and engineering support. The main product resulting from this initial stage of
study is a regional scale geohazard map, which identifies landslides, fault crossings, areas of
liquefaction potential, salt domes, mud volcano activity and areas of gas hydrates. However, these
results retain large uncertainties and are not suitable for detailed design (Randolph et al., 2011).

The next phase of work (Phase 2) occurs after a discovery has been made and builds upon
the first phase baseline geohazards assessment to define geohazard issues that may affect specific
components of a proposed system. The Phase 2 investigations will involve planning and execution

57
Offshore Foundations: Technologies, Design and Application

of detailed geophysical and geotechnical data acquisition programmes as well as preliminary site
characterisation activities. Phase 2 investigations may include the following:

 Acquiring high resolution geophysical and geotechnical data sets


 Developing a predictive soil model
 Developing the project GIS further and mapping geological conditions in the
foundation zone or along route alignments
 Conducting detailed terrain analysis, and interpretation of high-resolution
geophysical data to identify specific hazards within the project area
 Constructing preliminary hazard susceptibility maps showing, for example,
rugged terrain, faults, landslides, liquefiable terrain, and submarine canyon
crossings
 Identifying specific targets requiring investigation to develop final design
parameters
 Interacting with the engineering team to discuss geohazard constraints and
impacts on design
 Developing recommendations for special studies required to address specific
technical issues.

After these works are done, geohazard and design teams need to get together and review
the specific findings and determine whether additional investigations need to be completed
(Randolph et al., 2011).

Lastly, Phase 3 activities involve the final detailed integration of all of the various data sets.
This is an intensive stage of many projects and involves close interaction between geologists,
geotechnical engineers and owner’s representatives. During this stage, additional geophysical
processing may be carried out (e.g. seismic inversion), final sub-surface soil models are developed,
and soil parameters are defined for use in specialty studies such as site amplification, liquefaction
and slope stability analyses. Detailed geohazard assessments are also executed to address the
distribution, severity and frequency of geohazards such as submarine slope failures, mass gravity
flows, faulting, strong ground shaking, liquefaction, scour, gas hydrates and fluid expulsion.

There are two general types of data acquisition tools required to complete deep-water
developments. These include geophysical survey tools and geotechnical/geological sampling and
in situ testing tools. The latter will be discussed in the following section.

58
Offshore Foundations: Technologies, Design and Application

5.3.3. GEOTECHNICAL SITE INVESTIGATION


A geotechnical site investigation programme is most effectively carried out following
acquisition of site-specific geotechnical data, development of the preliminary soil model and
completion of the initial geohazard assessment. This enables the collection of targeted data and
samples to support the foundation engineering stage of a project, but also provides an opportunity
to acquire data for the assessment of specific seafloor features. The geotechnical investigation
provides data for specific soil properties and pore pressure conditions of seabed deposits,
involves borehole logging, field testing, sampling, and both geological and geotechnical laboratory
testing.

There are two main approaches for geotechnical site investigations: seabed mode and
down-hole mode (Campbell et al., 2008). Seabed mode methods characterise the shallow part of
the stratigraphic profile, which means depths less than 50 meters. These methods often include
large diameter piston core sampling and in situ testing, piezocone penetrometer testing (PCPT),
vane shear tests (VST) and T-bar and ball penetrometer testing (Peuchen and Rapp, 2007). Yet,
some facilities, such as TLPs, compliant platforms and mooring systems for SPARs, semi
submersibles and FPSOs vessels have driven pile foundations that require data on soils to a
greater depth than can be reached using seabed investigations approaches. In these cases,
sampling is completed using down-hole approaches. This involves using rotary drilling techniques
from a ship-based drilling platform. Sample types may include: piston samplers, push tubes,
rotary corers, percussion corers and pressurised corers to sample gas hydrates.

It is also important to execute field tests to identify pore pressure conditions, temperature
distributions and remoulded and residual shear strength of soft sediments. High-quality sampling
techniques should be used to retrieve cores for laboratory testing. Figure 56 enumerates the tests
that a testing programme should include (Randolph et al., 2011).

59
Offshore Foundations: Technologies, Design and Application

Soil texture and mineralogy

Clay mineralogy

Classification tests to
Age
identify
Pore water
salinity

Laboratorial Thermal
testing properties
programme
Peak strength

Critical state
Strength tests to Stress
identify Remoulded and residual dependency
properties
Strain rate
Stress tests to investigate effects
Strength
anisotropy
Cyclic
loading

Figure 56 – Tests that should be performed during a laboratorial testing program (Randolph et al.,
2011).

5.3.4. SUBMARINE SLOPE FAILURES AND SLIDES


Seafloor instability is the main geohazard threat encountered offshore and can have
catastrophic effects on offshore developments. Seabed instability is an issue even on the
continental shelf with seafloor gradients as low as 0.5ᵒ. Submarine slides can range in size from
relatively small coastal slides of less than a cubic kilometre to vast slides involving thousands of
cubic kilometres of material.

The mobility of submarine landslides can be characterised geometrically by the run-out


ratio L/H where L is the horizontal distance from source to deposit and H is the vertical elevation
of the debris flow source above the deposit, Figure 57. First introduced by Heim (1932) and used
later by Scheidegger (1973), the ratio can be predicted by considering the energy balance for a
dry mass sliding down a slope. Figure 58 shows the volume of material involved and run-out ratio
of various submarine and sub-aerial slides. From the compiled data, it appears that submarine
slides can be much larger than sub-aerial landslides (due to the concentration of each group of
dots), and they also tend to exhibit larger run-out distances for the same volume of sliding
material. This indicates that water plays a particular role in the mobility of the sliding mass.

60
Offshore Foundations: Technologies, Design and Application

A submarine slide involves an initiation event, usually a shallow or deep-seated slope


failure, often with retrogressive slumping, followed by mass gravity flow involving laminar visco-
plastic debris flow, and loose suspension turbidity currents. The slope failure and mass gravity
flow must be analysed as part of a geohazard assessment (Randolph et al., 2011).

Figure 57 – Definition of slide mobility

Figure 58 – Comparison of volume and run-out distance of submarine and sub-aerial slides
(Randolph et al. 2011 after Scheidegger 1973, Edgers and Karlsrud 1982, Hampton et al. 1976, Dade
and Huppert 1998, De Blasion et al. 2006)

61
Offshore Foundations: Technologies, Design and Application

62
Offshore Foundations: Technologies, Design and Application

6. CASE STUDY: GEOTECHNICAL CONSIDERATIONS FOR ULTRA-


DEEP OIL FIELDS OFF S ÃO TOME E PRINCIPE

6.1. PROPOSED DEVELOPMENT


In the past decade São Tomé e Princípe (STP) has been reported as future potential oil supplier
to the world, yet little progress has been made, largely because potential exploitation sites have water
columns ranging from 1800 to 3000 meters. With advances in deep sea production technology and
given that the most readily exploited production fields worldwide are now being exploited, STP has
once again come under greater attention.

As a member of the Community of Portuguese Speaking Countries (CPLP), STP has been used
to provide a focus for the development of this thesis, and this work provides a preliminary review on
an ultra-deep water design proposal for its offshore sites. The foundation solution that is presented
takes into account the economic issues, technical novelty and site conditions specific to STP. The
foundation proposed an anchor type system to which a floating platform is moored. Possible anchor
solutions might include suction caissons, torpedo anchors, and/or plate anchors; specifically, solutions
using a Torpedo Anchor and a SEPLA will be evaluated.

6.2. GEOTECHNICAL SITE CONDITIONS


São Tomé & Principe is a group of islands situated on the Gulf of Guinea, the island of
Principe is the nearest to the site where possible oil exploration is more likely. Figure 59 shows
the location of STP and the surrounding geology, it is also possible to see two red lines which refer
to the schematic cross sections presented in Figure 60 and Figure 61; the cross sections extend
from Principe Island to Nigeria, and Equatorial Guinea respectively.

Based on the information obtained from Figure 60 and Figure 61, the seabed soil is
composed of shales with a lithogenous nature, since they come from turbidity currents along the
continental slope of Equatorial Guinea and Nigeria. In the direction of Nigeria, operating water
depths range from 2600 to 3000 meters while towards Equatorial Guinea, they are 1800 to 2200
metres.

Unfortunately there is not a precise characterization of the ultra-deep seabed soil offshore
of Principe, though some investigation has been made in other parts of the Gulf of Guinea (GoG).
Starting from these studies and extrapolating some results, a pre-design characterization can be
carried out. Figure 62 illustrates the limits of GoG and the nations that border it.

63
Offshore Foundations: Technologies, Design and Application

Figure 59 – São Tomé & Principe location and surrounding geology, Courtesy of Agencia Nacional do
Petroleo of STP from: http://www.stp-eez.com/DownLoads/Posters/3_STP_RegionalGeol.pdf .

Figure 60 – Cross section from Nigeria cost to Principe Island, Courtesy of Agencia Nacional do
Petroleo of STP from: http://www.stp-eez.com/DownLoads/Posters/3_STP_RegionalGeol.pdf .

64
Offshore Foundations: Technologies, Design and Application

Figure 61 – Cross section from Principe Island to Equatorial Guinea Cost, Courtesy of Agencia
Nacional do Petroleo of STP from: http://www.stp-
eez.com/DownLoads/Posters/3_STP_RegionalGeol.pdf .

Figure 62 – The red circle limits the extent of Gulf of Guinea.

65
Offshore Foundations: Technologies, Design and Application

6.2.1. INDEX PROPERTIES OF GOG SEDIMENTS


The sediments present in GoG deep-waters are characterised by very high water contents,
values which can be between 150% and 250% at the seabed and gradually decrease with depth.
Figure 63 shows three different soil properties (water content, wet unit weight and submerged
unit weight) from 10 sites in the GoG. Figure 63(b) suggests soil unit weights starting from 12-13
kN/m3 at seabed and increasing to 13-15 kN/m3 below 6-8m, Puech (2004). Water content and
wet unit weight results are also supported by the studies carried out by the International Ocean
Discovery Program (IODP) in the GoG approximately 200 km off of the Ivory Coast, Figure 64 and
Figure 65. However, the submerged unit weight graph suggests higher values (approx. 1.5 to 2
times) than Puech (2004) proposes for the first 20 meters. The locations of the IODP boreholes
are identified in Figure 66.

The soils present in the Gulf of Guinea have plasticity indexes (PI) ranging from 70% to
120%, and as high as 150% near the seabed, Figure 67. Liquid limits (wL) range between 125%
and 175%, and when plotted in the Casagrande diagram (with its associated PI), Figure 68; reveals
that the soils are classified as highly plastic clays (CH) and highly plastic silts (MH).

Using the above liquid limit values and Figure 69, it is possible to determine that the
coefficient of consolidation (cv) of these soils is likely to be somewhere between 10-8 and 0.5 10-8.

When thoroughly analysed using X-Ray diffraction, the soil is found to have a high
percentage of clay (60% to 80%) (Puech, 2004). The majority of the clay is Kaolinite (more than
50%) but an important proportion of smectites is also found (15 % to 20%), which explains the
high plasticity of the soils (Puech, 2004).

Figure 63 – Deep-water sediments physical properties on Gulf of Guinea (Puech, 2004).

66
Offshore Foundations: Technologies, Design and Application

Figure 64 – Deep-water sediments physical properties, data from International Ocean Discovery
Program testing results, after Mascle et al. (1998). Sites are located offshore of Ivory Coast and with
depths ranging from 2090 to 4637 meters, Figure 66.

Site 959A

Site 959B

Site 960C

Site 962B

Site 959A

Site 959B

Site 960C

Site 962B

Figure 65 – Comparison of the IODP testing results (offshore of Ivory Coast) and Puech (2004)
sediment physical properties in GoG.

67
Offshore Foundations: Technologies, Design and Application

Figure 66 – Map showing the drilling locations of IODP in Gulf of Guinea, from: http://www-
odp.tamu.edu/publications/159_IR/images/map.jpg .

Figure 67 - Plasticity index of GoG sediments in depth (Puech, 2004).

68
Offshore Foundations: Technologies, Design and Application

Figure 68 – Plasticity chart of GoG deep-water sediments, Casagrande diagram (Puech, 2004).

Figure 69 – cv versus wL chart (Kulhawy & Mayne, 1990)

6.2.2. IN SITU STRESSES AND STRESS HISTORY


On the continental slope of the GoG, the upper part of the sediment column is a
consequence of sedimentation processes. When sedimentation progresses slowly, normally
consolidated soils, where the submerged weight of the particles is entirely supported by the soil
skeleton (no excess pore pressure), are more likely to be formed.

An over-consolidated soil is one that the effective stress currently applied (p’0) is lower
than the maximum effective stress applied previously (p’c). The ratio between these two (p’c/p’0)

69
Offshore Foundations: Technologies, Design and Application

is called the over-consolidation ratio (OCR). The OCR is commonly obtained from standard
oedometer test procedures.

In situ vertical effective stress profiles from the GoG, as well as from Gulf of Mexico are
presented in Figure 70. Whereas in the first few meters it is common to find OCR values around 2
in the GoG, the OCR decreases to values around 1.5 at depths of 15 to 20 meters. Below that depth,
the OCR decreases further, reaching values around 1.2 and 1.3. This profile differs a lot from Gulf
of Mexico, since in the GoM it is common to have an OCR of about 1 until 15 m depth, and a small
increment with depth afterwards (Puech, 2004). Whereas, the difference between p’c and p’0 is
almost constant all the way down in the GoG, in the GoM this difference gets greater with
penetration.

Figure 70 – Typical in situ stress profiles for Gulf of Guinea deep-water soils and the comparison
with Gulf of Mexico (Puech, 2004).

The latest work that has been done in this field suggests that using the term yield stress
ratio (YSR) rather than OCR to define the stress history of the Gulf of Guinea soils is more
appropriate, since no implicit assumption regarding the past over-burden pressure is introduced

70
Offshore Foundations: Technologies, Design and Application

(J.L. Colliat & H. Dendani et al., 2011). The YSR is determined from the ratio between vertical yield
stress σ’vy to the actual overburden pressure σ’v0 (YSR= σ’vy/ σ’v0).

J.L. Colliat & H. Dendani et al. (2011) derived the vertical yield pressures from a number
of standard oedometer tests performed on samples from several typical deep-water sites. They
plotted the results versus depth, Figure 71, and used them to compute the YSR as well. Once again,
the difference between σ’vy and σ’v0 is quasi constant with depth. Furthermore, they report this
difference to be typically between 15 kPa and 40 kPa for most sediments. The corresponding YSR
values start at about 3 in the first couple of meters, and tend to reduce to between 1 and 2, below
after about 10 m depth.

Even though the YSR exceeds 1, it does not imply over-consolidation in the geological
sense (i.e. no past overloading of the material). De Gennaro et al. (2005) have shown that these
soils exhibit a significant structuration effect, and that the difference between σ’vy and σ’v0 is a
quantitative measurement of the “extra-strength” due to soil structure.

Figure 71 – Vertical yield pressure and YSR versus penetration depth for typical GoG sites (J.L. Colliat
& H. Dendani et al., 2011).

Within the first 2 m below the seabed, higher values of YSR can be found. This occurrence
is called “crust”. The origin of this “crust” is still undefined and open to debate. It is known from
geological evidence that these sediments have never seen overburden stresses in excess of the
present state, meaning that the relatively high shear strengths of the “crust” are not the result of

71
Offshore Foundations: Technologies, Design and Application

mechanical over-consolidation. Based on a study by Brausse (2001), there are three families of
mechanisms that might be a possible cause for this phenomena:

 A strong inter-particle physiochemical bonding and ion exchange which may be combined
with effects of cementation and cause higher attraction between particles (through ionic
and van der Waals forces). This proposition is supported by Sultan et al. (2001) who
propose an interpretation of the apparent over-consolidation using a microscopic model
based on the theory of the diffuse double layer, which relates the mechanical behavior of
the soil to the variations in ionic concentration.
 An inorganic cementation involving cementing agents such as silicates, carbonates, iron
oxides, or alumina.
 Bioturbation (biological activity leading to reworking of soils and sediments) and organic
cementation. Ehlers et al. (2005) have presented X-ray radiographs of surficial cores from
offshore Nigeria showing a correlation between the most intensively burrowed sections
and the highest soil shear strengths, thus favouring the role of biological activity.

Currently, there is research work being done on this issue, however further effort is needed
to better understand the origin of the “crust”, the conditions for its development and the reasons
for its absence in some areas.

6.2.3. SHEAR STRENGTH PROFILES


Shear strength profiles are a determinant part of the final design solution, as well as when
choosing which structure it is more appropriate to adopt. Usually these profiles are established
from in situ CPTU and VST tests (uncorrected) in the precise location where the foundation will
be positioned, but in the STP case they have not yet been done. Thus, information was gathered
from multiple sites across the GoG (data collected from the sites in Figure 66by the International
Ocean Discovery Program), along with data from other similar places, such as: Brazil, Gulf of
Mexico and Mississippi Delta.

IODP performed several soil tests in four different sites, from the location map Figure 66
& Figure 73 below, and where it is possible to see that the sites were located in different
geotechnical environments. This implies different undrained shear strength profiles; for instance,
Figure 72 suggests that site 960 has very low undrained shear strength in the first 30 meters,
while sites 959 and 962 have a constant growth throughout their depth. This difference is justified
by the sedimentary activity in the continental shelf; while sites 959 and 962 are in a relatively
stable location, site 960 is at the edge of a slope failure where the soil has been disturbed recently
and has not reconsolidated yet. This reveals how important is to choose the location before

72
Offshore Foundations: Technologies, Design and Application

installing any foundation, since, despite the proximity between the sites (note 10’ is approx. 20
km), the resistance achieved would be very different.

Undrained Shear Strength


Su (kPa)
- 50 100 150 200 250
0

20

40

60 Site 962B
Site 961A
80
Depth (m)

Site 960A
100 Site 960C
Site 959A
120 Site 959B

140

160

180

200

Figure 72 - Undrained shear strength variation with depth in Gulf of Guinea, determined by
International Ocean Discovery Program using Vane shear and penetrometer tests, and test
locations.

In his analysis, Puech (2004) recognised that GoG soils have:

 Typical gradient in undrained shear strength that is often close to 1.5 kPa/m.
 Su/p’o ratios are high in comparison with onshore clays, where the values are about 0.25,
here they exceed 1 in the first meters of penetration and decrease to about 0.4-0.5 below
10 m.
 GoG clays are typically medium sensitive with sensitivity values ranging from 3 to 4, but
can reach higher values. Sensitivity is the ratio between the undisturbed and remoulded
undrained shear strength of the soils, and varies from about 1 for heavily over-
consolidated clays to values of over 100 for the so-called extra-sensitive or “quick” clays.
 Steel-soil interface adhesion factor (α) in these soft plastic clays is considered to be
between 0.8 and 1 (Tomlinson, 1957).

Puech (2004) also observed that the first 2 m of penetration can have two distinct types
of profiles (Figure 73):

73
Offshore Foundations: Technologies, Design and Application

 A “no peak” profile where shear strength starts from 1-2 kPa near the seabed and
increases linearly with depth (magenta line)
 A “peak” profile which is characterised by a sharp increase of s u, reaching 8.5-14.3 kPa at
0.5m penetration, then decreases to progressively get back to the deep gradient at about
2 m penetration (red line). This initial “peak” is what is called as “crust”. This profile is
more common in greater water depths, i.e. greater than 1000 m.

Figure 74 illustrates four different profiles of (undrained) shear strenght, τ: two normally
consolidated clays from the Gulf of Mexico (GoM) (Quiros et al., 2003), a Mississippi Delta clay
(Quiros et al., 2003), and the GoG 1.5 kPa/m profile proposed by Puech (2004). The various
symbols represent the results obtained by the IODP in field tests executed in the Brazil Basin, it is
reasonable to say that these test results have an acceptable fit to the 1 kPa/m profile of GoM.

From Figure 74 is comprehensible that the shear strength gradient in the Gulf of Mexico
has a range of about +/- 50-60% from the characteristic 1.25 kPa/m (red line, Figure 74). This
gradient obviously depends on the location; the closer it is to the Mississippi Delta, the lower the
gradient will be because of the soft under-consolidated clay layers. When the location is far from
the influence of the river the layers are much stiffer and the shear strength gradient is about
2kPa/m. In the Mississippi Delta profile, below 75 m depth the gradient is similar to the stronger
GoM strength profiles (dashed line).

Figure 73 – Typical cone resistance profile for GoG deep water sediments (WD – water depth),
Puech (2004).

74
Offshore Foundations: Technologies, Design and Application

GoG, Puech 1.5 kPa/m

Miss Delta, Quiros (2003)

GoM NC, Queiros 1.25


kPa/m (2003)
GoM NC, Quieros (2003)

Brazil Basin, DSDP (1977)

Figure 74 – Undrained shear strength profiles of Gulf of Mexico, Brazil and GoG proposed by Puech
(2004).

Figure 75 reveals that the majority of the shear strengths measured near the Ivory Coast
(sites 959, 960, 961 and 962) lie below the 1.5 kPa/m gradient line. However, this does not mean
that this gradient is not typical for the GoG. This soil, like those in the GoM may have a range of
about 50-60% in its gradient and these measured values may be in the lower range. Assuming this
similarity to the GoM, the gradient of undrained shear strength in the GoG may range between
0.75 kPa/m and 2.25 kPa/m.

75
Offshore Foundations: Technologies, Design and Application

Site 962B

Site 961A

Site 960A

Site 960C

Site 959A

Figure 75 – Shear strength profiles of diverse marine sites

Sultan et al. (2007) reported some studies from the continental slope of Nigeria. Even
though the water depth where this study was executed ranged between only 1100 m and 1250 m,
the study area was actually very close to Principe Island as shown in Figure 76. The shear strength
profile based on laboratory geotechnical tests is represented in Figure 77. This profile has a
similar development to the Puech (2004) proposal of a gradient of 1.5 kPa/m, therefore in future
calculations this is the gradient that will be used. However, the shear strength profile does not
start from the zero, therefore it is assumed to be 5 kPa in the first 3 m, and after that assumes the
proposed 1.5 kPa/m gradient (blue line Figure 77).

Príncipe
Island

Figure 76 –Nigerian continental slope study area, Sultan et al. (2007).

76
Offshore Foundations: Technologies, Design and Application

0
GoG, Puech 1.5
kPa/m
GoM NC, Queiros

2 GoM NC, Quieros

4
Depth (m)

10

12

Figure 77 – Shear strength profile of Nigerian continental slope obtained in laboratory geotechnical
tests and its comparison with shear strength profiles proposed by others.

6.3. DESIGN OF ANCHOR SOLUTIONS


Since the water depths in the zone of proposed offshore development near São Tomé &
Príncipe range from 1800 m to 3000 m, future installed facilities must be floating platforms and
hence the foundation systems in the seabed will be resisting tensile forces instead of compression.
Therefore, the only types of foundation solution suitable for this region are anchoring systems.
Anchoring systems that are selected for evaluation in this case study are the ones for which fewer
studies and investigation has been made, but on the other hand are likely to be the most economic
systems.

6.3.1. TORPEDO ANCHORS


The design process for Torpedo anchors, Figure 78, is rather complex due to the difficulties
of predicting the anchor embedment and set-up after installation. These two factors have a direct
effect on the anchor capacity and they depend on the geometry and characteristics of the anchor,
as well as the soil properties such as undrained shear strength and coefficient of consolidation
(horizontal and vertical).

77
Offshore Foundations: Technologies, Design and Application

Anchor Soil
geometry: Characteristics:
-Diameter Impact -Undrained
shear strength Determination Soil-structure
-Length Velocity: of Tip interaction and Pull-out
(su) Embedment soil capacity
- nº of flukes -Free-fall
- Coefficients Depth consolidation
-Fluke height
of
configuration consolidation
-Weight (Cv and Ch)

Figure 78 – Design process for torpedo anchors.

6.3.1.1. Anchor Geometry


The design of torpedo anchor structures is based on four parameters:

 Diameter, which ranges between 0.76 m and 1.07 m


 Length, which may vary between 12 m and 17 m
 Mass which depends on the filling and may be between 24.6 and 98 tons
 The number of flukes; usually 4 but it can also be 3 or none, and the flukes may have a
width from 0.45 m to 0.9 m, and be 9 m to 10 m long.

The main problem with torpedo anchors is the lack of field experience, especially outside
Brazil. Therefore, the geometry suggested in this text will be based on that used in the Albacora
Leste Field (FPSO P-50), a FPSO unit in water depth of 1400 m with required capacity of 7500 kN
(Araujo et al., 2004).

The torpedoes used for the FPSO P-50 were type T-98, which are illustrated in Figure 79
and Figure 80, and had the following characteristics (Brandão et al., 2006):

 Total mass of 98 tons


 Diameter of 1.07 meters
 Length of 17 meters
 Four stiffener wings (flukes): 0.9 m wide x 10m long.

Figure 79- Schematic longitudinal section drawing of the T-98, Brandão et al. (2006).

78
Offshore Foundations: Technologies, Design and Application

Figure 80 – Photos of the T-98 body sections welding and its final adjustments, Brandão et al. (2006).

The ballast does not full-fill the torpedo interior, Figure 79; the reason is to lower the
gravity centre of the structure closer to the nose. This greater proximity to the nose creates more
stability while free-falling before reaching the seabed. Apart from oceanic currents and the
misalignment of fins, stability is the principal factor in the torpedo trajectory. It has been
recommended that, the centre of gravity should be 10% of the total anchor length below the
hydrodynamic (geometric) centre (Hickerson et al., 1988; Raie & Tassoulas, 2009).

6.3.1.2. Impact Velocity


Before defining the velocity reached by the torpedo when impacting the seabed, the
equation for terminal velocity will be demonstrated. Normally, the impact velocity will less than
the terminal velocity, since to achieve this it must have a higher fall than those recommended for
torpedo anchor installation. These recommendations ensure torpedo stability during free fall.
Higher drops may result in the torpedo trajectory crossing ocean currents, and greater velocities
may induce stronger turbulence putting in risk the torpedo verticality and impact location.

To determine the terminal velocity, Newton’s second law is considered. The forces acting
on the penetrator during the fall are: drag force, anchor weight and buoyancy force (Raie, 2009).
Therefore:

𝑑𝑉
𝑊𝑆𝑢𝑏 − 𝐹𝑑 = (𝑚 + 𝑚𝑎 ). (1)
𝑑𝑡
1
𝐹𝐷 = . 𝐶𝐷 . 𝜌𝑤 . 𝐴𝑝 . 𝑣 2 (2)
2
𝑊𝑆𝑢𝑏 = 𝑚. 𝑔 (3)

Where

 Wsub is the submerged weight of the anchor in the water

79
Offshore Foundations: Technologies, Design and Application

 Fd is the drag force


 m is the torpedo mass
 ma is the added mass of ballast
 V is the instant anchor velocity
 t is the time
 𝜌𝑤 is the water density
 Ap is the anchor frontal area
 CD is the drag coefficient
 g is the acceleration of gravity

After release, the velocity increases until the drag force and anchor weight have the same
intensity. At this point, the incremental velocity change and thus the acceleration is zero, and the
torpedo has attained the so-called terminal velocity. Considering the previous equations (Eqs. 1
to 3), the terminal velocity (VT) can be calculated:

𝑊𝑆𝑢𝑏 = 𝐹𝑑 (4)
1
𝑚. 𝑔 = . 𝐶𝐷 . 𝜌𝑤 . 𝐴𝑝 . 𝑣 2 (5)
2
𝑚. 𝑔
𝑉𝑇 = √ (6)
1
.𝐶 .𝜌 .𝐴
2 𝐷 𝑤 𝑝
The difficulty on this expression is in the determination of CD. Although, the drag
coefficient value should change throughout the torpedo’s fall, a study from Hasanloo (2011)
showed a good agreement between the measured velocities and the calculated results when using
a constant drag coefficient. Table 5 summarises the CD values proposed for different penetrometer
types that have been investigated.

Drag coeff., CD Penetrometer type Reference


Cylindrical penetrometers with a
0.70 True (1976)
pointed nose

0.15 to 0.18 European Standard Penetrators Freeman et al. (1984)

0.03+0.0085L/D European Standard Penetrators Freeman and Hollister (1988)

0.63 Four fluke DPAs Øye (2000)

0.33 Torpedo anchors Fernandes et al. (2005)

Table 5 – Drag Coefficients according to different penetrometer shape and reference, in the Freeman
and Hollister (1988) formula, L is the penetrometer length and D is the diameter.

80
Offshore Foundations: Technologies, Design and Application

Given that the company with most experience of this kind of anchor system is the Brazilian
petroleum company Petrobras, the drag coefficient purposed by Fernandes et al. (2005) will be
considered for the torpedo anchor assessment in this study. That value is CD = 0.33, so the terminal
velocity for the T-98 is:

98000 × 9.8
𝑉𝑇 = √ = 80 𝑚/𝑠
0.5 × 0.33 × 999.972 × 𝜋(1.07)2 /4 (7)

Nevertheless, it is advised, as previously noted in Section 4.6, to use drop heights above
the seabed between 30 m and 150 m, which usually result in impact velocities from 0.5 to 0.33
times the terminal velocity (Medeiros, 2002). Hence, for this study, an impact velocity of 40m/s
will be considered. Using equations 1 to 3 it is possible to define the acceleration when the defined
velocity is reached (in this case 7.4 m/s-2), and consequently, using the equations for conservation
of mechanical energy, determine the height needed to achieve the chosen impact velocity as the
anchor reaches the seabed:

1
𝐸𝑐 = . 𝑚. 𝑣 2 = 0.5 × 98000 × 402 (8)
2
𝐸𝑝 = 𝑚. 𝑎. ℎ = 98000 × 7.4 × ℎ (9)
𝐸𝑐 = 𝐸𝑝 ⟹ 𝒉𝒇𝒓𝒆𝒆−𝒇𝒂𝒍𝒍
(10)
= 𝟏𝟎𝟖. 𝟒 𝒎
6.3.1.3. Soil Characteristics
As previously mentioned, accurate information about the constitution of the seabed of the
São Tomé & Príncipe (STP) Economic Exclusive Zone (EEZ) is currently very limited. So, in order
to continue the design of the torpedo solution, the following table summarises the characteristics
of the soils adopted in Principe Island surroundings that are needed for further analysis.

Soil Characteristic Adopted Value Reference


Undrained shear strength
1.5 kPa/m Puech (2004)
gradient
Kulhawy & Mayne
Coefficient of consolidation 10-8 m2/s
(1990)

Wet unit weight 14 kN/m3 Puech (2004)

Sensitivity 4 Puech (2004)

Adhesion factor (α) 0.8 Puech (2004)

Table 6 – Adopted values for different soil characteristics and references from which they were
based on.

81
Offshore Foundations: Technologies, Design and Application

6.3.1.4. Tip Embedment Depth


Tip embedment depth is the penetration into the seabed achieved by the pointed tip of a
torpedo anchor. The deeper a torpedo is installed, the higher the pull-out capacity will be for a
given shear strength profile. Therefore, it is absolutely crucial to accurately predict this
penetration so that the resistance is well defined. The penetration of a torpedo strongly depends
in three aspects: geometry of the penetrator (L/D), impact velocity of the penetrator and strength
of the soil.

The prediction of anchor embedment has been evolving over the years, not only because
of the use of dynamic penetration models based on empirical results, e.g. True Model, and
Computational Fluid Dynamics (CFD) (Raie & Tassoulas, 2009; O’Loughlin et al., 2013); but from
the implementation of numerical procedures for coupled finite element analysis of dynamic
problems in geomechanics as well (Aguiar et al., 2011; Carter & Nazem, 2013; Sabetamal et al.,
2014). Each of these approaches have proved to be reasonably accurate.

Petrobras, which developed the torpedo technology, perform a conductor drive analysis
to evaluate the penetration of the torpedo base system using a computer program based on the
model proposed by True (1976). This model relies on soil parameters whose values are assumed
as known, fixed, and deterministic. Yet, it is common sense that these soil parameters have a
significant degree of variability, which may affect negatively the accuracy of the response given
by the simulation method (Kunitaki et al., 2008).

On the other hand, a more accurate approach is to predict the penetration based on
empirical model trials in the field or in lab tests. Though, first carrying out a CFD or FEM prediction
and confirming it with field trials or lab tests is the recommended procedure. Applying numerical
procedures is out of the scope on this work, but would be recommended for future studies.

So, making a prediction for the STP offshore area requires a review of studies made in this
field. Richardson et al. (2009) carried out several laboratory tests and compared the results to
empirically based predictions. These laboratory tests involved a series of centrifuge model tests
of 1:200 reduced scale model anchors in normally consolidated clay (Kaolin). Table 7 summarises
some relevant characteristics of the tests that were performed. These tests studied the influence
of different soil properties in the embedment of model projectiles, which were designated as
either quasi-static or dynamic installation, with impact velocities up to 30 m/s. The influence of
soil properties such as shaft adhesion and undrained shear strength gradient in the resulting
penetration are illustrated in the graphics of Figure 81 and Figure 82 respectively.

82
Offshore Foundations: Technologies, Design and Application

Dimension Model Prototype

Anchor Length 75 mm 15 m

Anchor Diameter 6 mm 1.2 m

Mass 6.2-12.7 g 49.6-101.6 tons

Average Undrained Shear


0.21 kPa/mm 1.05 kPa/m
Strength Gradient
Table 7 – Model and Prototype anchor dimensions and soil property.

Using the studies by Richardson et al. (2008) of the sensitivity of embedment depth
predictions to shaft adhesion factor (Figure 81), it was possible to interpolate for an adhesion
factor of 0.8 and extrapolate an impact velocity of 40 m/s to predict a penetration of 40.4 m for
the T-98 torpedo being considered. Nevertheless, this value is not accurate since the undrained
shear strength gradient in this case is lower than the 1.5 kPa/m, presumed value for STP.
However, on examination Figure 82, contains a curve for an undrained shear strength gradient of
1.5 kPa/m, and it indicates that the penetration would be greater than 40.4 m. This happens
because the adhesion factor used for this trial was 0.4, much lower than the value of 0.8 assumed
for the T-98 torpedo. As a result, this prediction of 40.4 m is an upper bound limit for the final
value.

0 40

α…

Figure 81 – Sensitivity of embedment depth predictions to shaft adhesion factor according to


Richardson et al. (2008) and interpolation for adhesion factor of 0.8 and impact velocity of 40m/s.

83
Offshore Foundations: Technologies, Design and Application

Figure 82 – Sensitivity of embedment depth predictions to undrained shear strength gradient,


Richardson et al. (2008).

O’Loughlin et al. (2013) after gathering penetration data from worldwide field tests and
comparing them to centrifuge tests of equivalent prototype scale models, were able to propose a
relationship that predicted penetration depth with reasonable accuracy for this very large dataset
that encompassed a wide range of anchor masses, geometries and impact velocities, Figure 83.

Figure 83 – Comparison of centrifuge and field test embedment data, O’Loughlin et al. (2013).

84
Offshore Foundations: Technologies, Design and Application

From Figure 83 it is possible to see the good agreement between the formulated curve and
the dataset. So, the prediction for STP using the O’Loughlin et al. (2013) proposed relationship,
would be:

1/3
𝑧𝑒 𝐸𝑡𝑜𝑡𝑎𝑙
≈( 4 )
(11)
𝑑𝑒𝑓𝑓 𝑘. 𝑑𝑒𝑓𝑓
𝐸𝑡𝑜𝑡𝑎𝑙 = 𝐸𝑐 + 𝐸𝑝
2
(12)
= 0,5. 𝑚. 𝑣𝑖𝑚𝑝𝑎𝑐𝑡 + 𝑚′ . 𝑔. 𝑧𝑒
1
2
0,5. 𝑚. 𝑣𝑖𝑚𝑝𝑎𝑐𝑡 + 𝑚′ . 𝑔. 𝑧𝑒 3
𝑧𝑒+1 ≈ ( ) × 1.17
1.5 × 1.174 (13)

= 39.24 𝑚
Where m’ is the effective mass of the anchor (anchor mass minus mass of soil displaced)
and g is the Earth’s gravitational acceleration. Penetration had to be calculated using an iterative
process in Equation 13. It is important to notice that the ratio Ze/deff=33.54, according to Figure
83, is not as accurate as to lower Z e/deff ratios (0 to 20), and when examining Figure 83, it is
apparent that Equation 13 provides a conservative prediction of Z e/deff ratio for a given situation
or installation energy value. The two grey lines traced in the graph bound almost every result,
additionally the grey line below the original equation line intersects the field test result from T-
98 (Brandão et al., 2006). Therefore, this ratio may vary 13% from the real depth, so the
penetration achieved is expected to be between 36 m and 43 m.

Freeman et al. (1984) reported penetrations of 30 meters for penetrators with an aspect
ratio of 10, weight of 1.8 tons, and an impact velocity of 50 m/s when installed off the West coast
of Africa, and Freeman and Burdett (1986) report penetrators reaching 62 m/s with an associated
burial depth of up to 35 m at the NAP site (offshore of the Bahamas).

6.3.1.5. Effects of installation in the resistance of the soil


The effect of installation cannot be ignored, as it generates an initial perturbation of the
soil during installation. A reduction in the undrained shear strength in the vicinity of the anchor
occurs not only because of the change in the initial stress state, but also because of the shearing of
the soil caused by the penetration and the soil deformation with a constant volume (Aguiar, 2011).

As the pile penetrates into the soil, the soil moves around the pile in the vertical direction
(Komurka, 2003). Randolph et al. (1979) comment that the penetration of a cylindrical pile into a
clay soil affects a region of up to 20 times the diameter of the penetrating pile.

The disturbance in the soil around the pile produces an excess of pore-water pressure,
which reduces the effective stress in the vicinity of the pile. As the volume of mobilised soil tends

85
Offshore Foundations: Technologies, Design and Application

to be equal to the volume of the pile, this issue is more significant when installing piles with larger
diameters. The process of reconsolidation of the soil is described as the dissipation of the high
pore-pressure generated during the pile installation, while this occurs the effective pressure
increases, thus increasing the lateral resistance of the pile (Richardson et al., 2009).

Therefore, this reconsolidation process is a very important aspect in the torpedo design,
and must be carefully predicted. Several studies have been made in order to better understand
this phenomenon. Richardson et al. (2009) carried out a series of experimental tests. Comparing
the results with theoretical models, they indicated that 50% of the resistance acquired by anchors
with dynamic installation was achieved within 35 to 350 days after the pile installation. In the
same studies, the authors comment that 90% of the capacity is achieved within 2.4 to 40 years
after the pile installation. It is important to notice that these values are highly dependent on the
soil characteristics where the pile is installed, and the authors recommend the use of this kind of
anchor where reconsolidation occurs rapidly, i.e. locations where the pile achieves 50% of the
resistance capacity in a period within 30 to 90 days.

It is also important to mention that the applied loads are much lower than the total
capacity of the pile, for the reason that safety factors used in this type of anchor is between 1.5
and 2.0 (Eltaher et al., 2003). Therefore, because the design load conditions are usually associated
with a long period, this means that the reconsolidation period is sufficient to achieve the
resistance capacity taking into account the safety factor used.

6.3.1.6. Pull-Out Capacity


The pull-out capacity depends on the characteristics of the soil (i.e. undrained shear
strength), the geometry of penetrator (e.g. weight, number and geometry of the flukes, shaft area
of the pile) and the angle between the applied load and the plane of the flukes.

Therefore, it is important to predict the time the soil takes to reconsolidate, thus,
estimating the short and long-term pull-out capacities. Despite the importance of soil
reconsolidation, this will not be approached in this work, and only the means for prediction of the
long-term pull-out capacity will be examined and evaluated for the case being considered.

Normally, pull-out capacity is predicted using FEM programs and laboratory model tests.
The estimation using the finite-element method has some limitations, since the pile is usually
considered to be vertically oriented in the soil with an initial state of stress in the soil equal to the
state prior to the anchor installation and the initial excess pore-water pressure equal zero,
ignoring the installation effect (Medeiros, 2002; Lieng et al., 2000).

86
Offshore Foundations: Technologies, Design and Application

The holding capacity behaviour of the FPSO P-50 in Campos Basin (Brazil) was studied
using a 3-D finite element program, called AEEPEC3D, Brandão et al. (2006). During the torpedo
pile design the following aspects were considered:

 The wings of the torpedo pile were considered to be located in the alignment that
produces the minimum load capacity (i.e. 45O off the alignment of the load – Plane
1 in Figure 84).
 The submerged self-weight of the torpedo pile was considered in the load capacity
evaluation.

Figure 84 – Cross section of a torpedo anchor with four flukes (Aguiar el al., 2009).

In Figure 85, the graph shows the results for loads applied to a vertically installed pile at
three different angles (0, 43.6 and 90 degrees to the horizontal) and for three piles installed 10,
20 and 30 degrees from the vertical (tilt) with the load applied 43.6 degrees from the horizontal
(Brandão et al., 2006).

Figure 85 – T-98 Holding Capacity for 4 tilts and 3 different load directions in Campos Basin, Brandão
et al. (2006).

87
Offshore Foundations: Technologies, Design and Application

The minimum ultimate capacity is associated with the pull-out load being applied along
the axis of the pile (pink line in Figure 85). Even if, the pile suffers a deviation in its penetration,
the pull-out capacity will be the lowest when the load acts in the same direction as the pile is
installed. This tendency is proven by the decline of capacity when the tilt on the pile tends to be
in the direction of the load (lines green, blue and red).

Therefore, determining the vertical pile resistance is a good approach in terms of the
evaluation of a safe design load. This resistance can easily be determined using some simple
calculations; the vertical pile resistance is given by the sum of three components:

𝑅 = 𝑅𝑠 + 𝑊𝑝 + 𝑅𝑏 (14)
Where,

𝑅𝑠 = 𝐴𝑠 × 𝑠̅𝑢 × 𝛼 (15)
𝑅𝑏 = 𝐴𝑡 × 𝑠𝑢 (𝑧𝑖) × 𝑁𝑐 (16)
𝑅𝑠 friction resistance obtained from the torpedo shaft plus the flukes
𝑊𝑝 weight of the torpedo pile (980 kN)
𝑅𝑏 resistance obtained from the soil that has to move so the pile can be pulled out (reverse
end bearing resistance)
𝐴𝑠 shaft area plus the fluke area that contributes to friction resistance and it is 129 m 2
𝑠̅𝑢 average shaft undrained shear strength along the pile
𝛼 adhesion factor
𝐴𝑡 transversal area of the torpedo pile
𝛾𝑠 unit weight of the soil
NC end bearing capacity factor (12)
Offshore Site Campos Basin, Brazil São Tomé & Príncipe

Tip embedment depth (m) 33 40

Average undrained shear strength


5+2z 1.5z
gradient (kPa)
Average undrained shear strength on
54 47.25
shaft (kPa)

Adhesion factor (α) 0.6 0.7

𝜸𝒔 (kN/m3) 15 14

Rs (kN) 4180 4880

Rb (kN) 2590 3470

88
Offshore Foundations: Technologies, Design and Application

Vertical pile resistance (kN) 7750 8720


Table 8 – Soil properties, penetration and vertical resistance of Campos Basin in Brazil and STP
offshore. Average undrained shear strength gradient was proposed by Medeiros (2001).

Table 8 shows the properties of the soil and its interaction with the pile, as well as the
installation specifications of the pile, and finally, the vertical pile resistance determined using
Equations 14 to 16. The friction resistance and reverse end bearing resistance are the two
principal sources of resistance; however, both of them are very sensitive to an empirical factor.

On one hand, the reverse end bearing factor has a huge effect on the resistance.
Unfortunately, at present, there is no agreement on which N c value should be used. Based on
experimental results, Watson et al. (2000) suggested that the tension bearing resistance is similar
to the compression bearing resistance in terms of magnitude, therefore, the reverse end bearing
value may be calculated in a similar way to that for compression end bearing.

On the other hand, shaft friction resistance is highly dependent on the adhesion factor, this
factor changes a lot within the consolidation of the soil around the pile. Although, adopted value
of 0.8 was adopted for the adhesion factor, as proposed by Puech (2004), there are some divergent
opinions from others authors. The α factor is often estimated as 1/St, where St is the sensitivity,
though, this would suggest quite adhesion low factors in some clays which are not borne out by
testing, i.e. in this case, a sensitivity of 4 would imply α = 0.25, which is quite low (Houlsby &
Byrne, 2005).

Dendani & Jaeck (2007) reported that the axial resistances recorded during model tests of
pipelines in clay recovered from the GoG were best predicted by taking α=0.7 and α=0.35 for the
peak resistance and residual resistance, respectively. Thus, and assuming that the general
structure and undrained shear strength of the clay in the tests was the same as that in situ, the
shaft resistance for the torpedo anchor considered here could be estimated as 2130 kN in the
short-term (with the soil remoulded, i.e. residual strength applies) and 4270 kN for the peak
resistance, i.e. after the soil has reconsolidated following installation and recovers the natural
structure present prior to installation. Note that α = 0.7 is lower than the 0.8 value proposed by
Puech (2004), which may indicate that it is not expected that the soil recovers all its initial
structure.

89
Offshore Foundations: Technologies, Design and Application

6.3.2. SEPLA
The development of suction embedded plate anchors (SEPLA) emerged from the need for
a reliable alternative to driven piles, suction embedded piles and drag-embedded plate anchors
(VLAs). With this in mind, SEPLA was conceived to combine the advantages of suction piles and
VLAs without their drawbacks:

Anchors
Advantages Drawbacks
type
 Large
 Known penetration
Suction pile  Costly
 Geographical location
 Difficulty to handle
 High loads required to achieve initial
 Geotechnical penetration and final capacity.
VLA
efficiency  May not be suitable where precise
positioning specified.
Table 9 - Advantages and drawbacks of suction piles and drag embedded plate anchors (VLAs),
Wilde et al. (2001)

The application of this type of anchor is not entirely new in the GoG. SEPLA anchors were
selected by ExxonMobil in 2003 for mobile offshore drilling unit (MODU) moorings between the
Xicomba and the Kizomba A fields offshore from Angola (Dahlberg et al., 2004).

The functional requirement of SEPLA is to resist the specified maximum factored mooring
line load, while avoiding significant displacements, both in the direction of the applied load or
vertically. SEPLA holding capacity is related primarily to three basic aspects, which must be
defined for the design of the solution, those aspects are:

 Anchor plate area


 Undrained shear strength
 Penetration depth

6.3.2.1. Anchor Geometry and Installation

The SEPLA consists of a solid rectangular steel fluke (approx. 20 cm thick), a shank
connecting the fluke to a pad-eye (i.e. the loading point) and a full-length keying flap hinged to the
top edge of the fluke, Wang et al. (2012), Figure 86.

90
Offshore Foundations: Technologies, Design and Application

Figure 86 – Typical SEPLA with keying flap, Wang et al. (2012)

Typically, for mobile offshore drilling units (MODU), the fluke is a solid steel plate 2.5 m to
3.0 m in width and 6 m to 7.3 m in length. For permanent installations, the fluke is usually 4.5 m
wide and 10 m long (Wilde et al., 2001). According to information published by Technip, for two
example applications of these solutions for MODU and a permanent installation called Red Hawk,
the suction embedded followers had a diameter of 3 m and 5.5 m, and the installations depths
were 21.5 m and 24 m respectively. Courtesy of Technip, http://events.energetics.com
/deepwater/pdfs/presentations/session5/gengshenliu.pdf .
Based on the above, the geometry and the characteristics proposed for SEPLA applied
offshore of STP are:

 Plate area: 10 m length by 4.5 m width


 Shank: 2.5 m high (padeye eccentricity) and at an angle of 60ᵒ with the fluke.
 Anchor thickness: 0.20 m
 Anchor weight: 50 tons
 Installation penetration: 24 m

During installation, the SEPLA uses a suction pile (referred to as a "follower") for
embedment to the design penetration. The SEPLA is inserted in slots at the bottom of the follower
and retained by the mooring line and recovery bridle.

The suction follower, with the SEPLA slotted into its base, is lowered to the seafloor,
allowed to self-penetrate and then suction embedded in a manner similar to a suction pile. Once
the SEPLA has reached its design penetration depth, the mooring line and retrieval bridle that
hold the SEPLA secure in the bottom of the follower are released by the installation ROV

91
Offshore Foundations: Technologies, Design and Application

(remote operated vehicle). The pump flow direction is then reversed and water is pumped back
into the follower; the follower moves upward, leaving the SEPLA in place.

The SEPLA fluke will be in a vertical orientation at this time. For a recoverable
application, a small submersible buoy will support the recovery bridle. At this time, the mooring
line is tensioned by the installation vessel in the direction the SEPLA is to be loaded. This
"keying” tension will:

• Pull the initially vertical mooring line through the soil so that it forms an inverse
catenary shape from the mudline to the anchor shackle.

• Start rotation of the SEPLA fluke to an orientation perpendicular to the direction of the
mooring line at the anchor end.

• Set the keying flap to prevent further loss of penetration beyond that necessary to set
the keying flap.

At this time, the SEPLA is ready to develop its full pull-out capacity. Regardless of the initial
orientation of the fluke to the mooring line, the SEPLA with its long, rigid shank will rotate to
present the maximum projected area to the direction of pull. This ensures that the ultimate pull-
out capacity, based on the anchor's penetration depth and soil properties, is achieved (Wilde
et al., 2001).

However, keying also represents the major issue of the SEPLA, since the plate moves
vertically and horizontally in addition to rotating, this leads to several uncertainties associated
with the final embedment depth during operation.

6.3.2.2. Keying

The anchor keying process promotes two negative effects in the plate holding capacity.
First, it induces an upward movement on the anchor during the plate rotation, hence reducing the
embedment depth; and second, the soil in the immediate vicinity of the plate anchor is remoulded,
therefore reducing the soil strength (Randolph et al., 2005). Even though, this latter effect may be
recovered as the soil reconsolidates, the loss of embedment is permanent. As clay deposits in Gulf
of Guinea are typically characterized by an increasing strength profile with depth, any loss in
embedment will correspond to a non-recoverable loss in potential anchor capacity.

The U.S. Naval Civil Engineering Laboratory (NCEL) has produced some guidelines (NCEL,
1985) about the loss of embedment during anchor keying; recommending the design takes into
account a loss of twice the anchor breadth (2B) in cohesive soils. However, NCEL (1985)

92
Offshore Foundations: Technologies, Design and Application

recognizes that the loss of embedment is also a function of anchor geometry, soil type, soil
sensitivity, and duration of time between penetration and keying, thus these aspects may reduce
or enlarge the embedment loss. More recently, Wilde et al. (2001) reported in situ full scale and
reduced scale onshore and offshore test results for SEPLAs in clay. Soil sensitivity was in the range
of 1.8 to 4.0 for the different test sites and the loss of embedment during keying was 0.5 to 1.7
times the anchor breadth, with lower embedment losses corresponding to higher soil sensitivities.

None of the studies previous to O’Loughlin et al. (2006) studied the effect of loading
eccentricity on the keying process. The loading eccentricity is the distance between the padeye,
where the keying forces are applied, and the centre of rotation of the plate anchor; Figure 87
illustrates the plate anchor setup, where the loading eccentricity represented by the letter “e”.
O’Loughlin et al. (2006) showed a strong dependence between the loss in embedment and loading
eccentricity (e). At a large eccentricity ratio (e/B≥1), the loss in anchor embedment was no greater
than 0.1B. When the eccentricity ratio was less than 1, the loss in anchor embedment increased
linearly with the reduction of the eccentricity ratio, as is shown in Figure 88. As the plate displaces
and rotates, F may be expressed in terms of forces perpendicular (Fn) and parallel (Fs) to the plate
and a moment (M) at the mid-point of the plate, Figure 87 . The larger the eccentricity is, the larger
will be the moment on the plate anchor caused by the keying forces, the more rapidly the plate
anchor will rotate and the less it will translate. Figure 89 illustrates the effect of high eccentricity
and low eccentricity loading paths, low eccentricity requires a much higher Fs to achieve the same
Fn. This greater parallel force makes the plate translate upwards before rotating.

Figure 87 – Plate anchor setup before keying process.

93
Offshore Foundations: Technologies, Design and Application

Figure 88 – Graph showing the results achieved by O’Loughlin et al. (2006) for eccentricity ratios
between 0.17and 1.0.

Figure 89 – Combined loading paths for high and low eccentricity plate anchors, O’Loughlin et al.
(2006).

94
Offshore Foundations: Technologies, Design and Application

Gaudin et al. (2006) investigated the influence of suction installation on anchor keying and
anchor capacity. They found that the loss in embedment after a suction installation is lower than
after a jacked-in installation, which is when the plate is pushed into the soil. For a square anchor
with e/B=0.66 and a 45ᵒ pull-out angle, the loss of embedment was reduced from the range of 1.3
to 1.5 times the anchor breadth for a jacked-in plate to a range of 0.9 to 1.3 times the anchor
breadth for a suction installed plate anchor. Gaudin et al. (2006) concluded that greater soil
disturbance during suction installation was responsible for the lower loss in embedment during
anchor keying. As the disturbance drops the shear strength of the soil, the resistance that the soil
gives to the rotation of the plate is lower, thus the rotation is faster and the loss in embedment is
lower. They also observed a strong correlation between the loading angle and the loss in anchor
embedment.

Taking into account the issues discussed in these previous studies, Song et al. (2009) ran
several centrifuge tests and developed large deformation finite-element (FE) analyses. From the
results of those studies, they found that the loss in anchor embedment during anchor keying may
be expressed in terms of a non-dimensional anchor geometry factor, which is a function of the
eccentricity of the padeye, angle of loading, and the net moment applied to the anchor at the stage
where the applied load balances the anchor weight. The procedure to estimate the loss in anchor
embedment to any given pull-out angle during keying is summarized as follows:

1. Calculate the loss in anchor embedment (Δze/B) during vertical pull-out (θ=90ᵒ), using
eq. (17) taking into account all the anchor geometry measurements and anchor
submerged unit weight. The initial moment M0 corresponding to zero net vertical load on
the anchor is given by eq. (18).

∆𝑧𝑒 0.2
=
𝐵 𝑒 𝑡 0.3 𝑀 (17)
( ) ( ) ( 0 )0.1
𝐵 𝐵 𝐴𝐵𝑠𝑢
𝑀0 = (𝑓 + 𝑊 ′ 𝑎 )𝑒 − 𝑓𝑒𝑓 − 𝑊′𝑎 𝑒𝑤 (18)

2. Determine the constant Cθ, using the gradient kθ=0.005 (per degree of pull-out angle) and
the loss in anchor embedment (Δze/B) calculated in the first step for a pull-out angle
θ=90ᵒ.

∆𝑧𝑒
𝐶𝜃 = − 𝑘𝜃 ∙ 𝜃 (19)
𝐵
3. Finally the loss in anchor embedment under any given pull-out angle is obtained from eq.
(20), substituting the constant Cθ found in the second step.

95
Offshore Foundations: Technologies, Design and Application

∆𝑧𝑒
= 𝐶𝜃 + 𝑘𝜃 ∙ 𝜃 (20)
𝐵

Where,
∆𝑧𝑒 Loss in anchor embedment
𝑒 Loading eccentricity for pull-out force
𝑡 Anchor thickness
𝑀0 Initial moment corresponding to 0 net vertical load
𝐴 Anchor area
𝑘𝜃 Gradient for estimating anchor loss in anchor embedment
𝑓 Shank resistance
𝑊′𝑎 Overall submerge anchor weight
𝑒𝑓 Loading eccentricity for friction resistance
𝑒𝑤 Loading eccentricity for anchor weight

In Appendix II, the calculations made to determine the loss in anchor embedment
according to Song et al. (2009) for this study are detailed. Below the results of the calculations
using equations (17) to (20) are presented:

∆𝑧𝑒 /𝐵= 0.94


𝑀0 = 5080.71 kN.m
𝐶𝜃 = 0.49
The equation that gives the loss of embedment under a given pull-out angle in the studied
area is:

∆𝑧𝑒
(𝜃) = 0.49 + 0.005 ∙ 𝜃 ⟹ ∆𝑧𝑒 = 0.94 × 4.5 = 4.19 𝑚
𝐵 (21)
≈ 4𝑚
This prediction (0.94B), according to Song et al. (2009), has a good agreement with that
expected according to Wilde et al. (2001), which gives a range between 0.5B and 1.7B, and also
with Gaudin et al. (2006) that predicts displacement in the range of 0.9B to 1.3B for suction
embedded plate anchors.

From O’Loughlin et al. (2006), the expected loss of embedment is 0.85B for a ratio e/B=0.58,
however the Song et al. (2009) formulation is more conservative because the equation that fitted
the results obtained in their studies with correlation coefficient of R2=0.95 was

96
Offshore Foundations: Technologies, Design and Application

∆𝑧𝑒 0.15
=
𝐵 𝑒 𝑡 0.3 𝑀 0.1 (22)
( )( ) ( 0 )
𝐵 𝐵 𝐴𝐵𝑠𝑢

Equation 17 is an upper bound of the fitted line which provides a conservative design curve,
Figure 90. Therefore, for this case study, it is plausible to say that the SEPLA embedment depth
after installation is about 20 m.

Eq.17

Eq.22

Figure 90 – Loss in anchor embedment during keying versus anchor geometry factor, Song et al.
(2009).

6.3.2.3. Plate Holding Capacity

As previously stated, the functional requirement of SEPLA is to resist the applied loads
without undergoing significant displacements. SEPLA holding capacity can be divided into short
and long term capacities. The short term capacity represents the capacity the soils have when they
mostly have an undrained behaviour. This situation suits best in typical storm conditions, and
where soils have low permeability like highly plastic clays.

Long term capacity is the capacity the SEPLA must have to resist long term quasi-static
loads in permanent installations. This term is not considered critical for deeply embedded plate
anchors in normally consolidated plastic clays, since the effective stress conditions around a plate
anchor will gradually change as the excess pore pressure dissipates. Hence, the soil will have
greater long term capacity value than the short term for normally consolidated soils.

97
Offshore Foundations: Technologies, Design and Application

There are several suggested procedures for evaluating the plate holding capacity in clay,
and in this work this estimation will be made according to three different procedures, which are:

 Wilde et al. (2001)


 Merifield et al. (2001)
 DNV-RD-E302 (2002)

Wilde et al. (2001) proposal


Wilde et al. (2001) based their procedure on the general bearing capacity equation for
square or rectangular foundations, which has the following formulation:

𝐵
𝐹𝑛 = 𝐴𝑝𝑙𝑎𝑡𝑒 × 𝑁𝑐 × 𝑠𝑢,𝑎𝑣𝑔 × (1 + 0.2 × ) (23)
𝐿
Where,

𝐹𝑛 bearing capacity of the anchor


𝐴𝑝𝑙𝑎𝑡𝑒 projected bearing area of the fluke and keying flap
Nc bearing capacity factor
su,avg average undrained shear strength across the anchor plate after keying
B width of the plate
L length of the plate

This expression assumes full suction can develop behind the fluke. Therefore, the bearing
capacity factor used is 12.5 (Brinch-Hansen, 1970). However, if somehow water is able to flow to
the back of the plate anchor, and the suction formed on the back is relieved, the value of Nc should
be reduced to 7.5 for deep failures modes (i.e. more than 5 fluke widths deep), for shallow failure
modes the author does not give any guidance (Wilde et al., 2001).

Yet this expression can be refined, because during deployment and proof loading a certain
amount of remoulding and soil disturbance near the SEPLA occurs due to the installation process
(penetration plus keying). This results in a temporary loss in available bearing capacity (NCEL,
1977). To take this effect into account, an additional term αn has been incorporated. Thus, the
revised equation is:

𝐵
𝐹𝑛 = 𝐴𝑝𝑙𝑎𝑡𝑒 × 𝑁𝑐 × 𝑆𝑢,𝑎𝑣𝑔 × (1 + 0.2 × ) × 𝛼𝑛 (24)
𝐿
Based on US Naval Civil Engineering Laboratory (NCEL) experience and tests, the value
for αn for clays is expected to range from 0.7 to 1.0, depending on clay type and sensitivity. For

98
Offshore Foundations: Technologies, Design and Application

clay sensitivities of 2 to 5, which are characteristic in the GoG, Wilde et al. (2001) recommend the
αn take a value of 0.7.

Merifield et al. (2001) proposal


Merifield et al. (2001) applied numerical limit analysis to rigorously evaluate the stability
of vertical and horizontal strip anchors in undrained clay. By using two numerical procedures that
are based on finite element formulations of the upper and lower bound theorems of limit analysis,
the ultimate pull-out capacity was obtained. Those formulations followed a standard procedure
by assuming a rigid perfectly plastic clay model with a Tresca yield criterion. The results were
compared with existing numerical and empirical solutions, and for most of the cases, the exact
anchor capacity was proved to be predicted to within 5%.

The equation used to quantify the ultimate anchor pull-out capacity in undrained clay is:

𝑄𝑢 = 𝐴𝑝𝑙𝑎𝑡𝑒 × 𝑁𝑐 × 𝑠𝑢,0 + 𝑊𝑎 (25)


Where Wa is the weight of the anchor.

It is apparent that this equation is quite similar to that proposed by Wilde et al. (2001), the
main difference is found in the estimation of the bearing capacity factor (N c). Once again, Nc relies
in the soil-anchor behaviour. The anchor behaviour can be divided into two distinct categories:
immediate breakaway and no breakaway. The immediate breakaway case is when the soil-anchor
interface cannot sustain tension, while the no breakaway case is assumed when there is adhesion
or suction between the anchor and the soil. In reality, it is likely that the true breakaway state will
fall somewhere between the extremities of the immediate breakaway and no breakaway case
(Rowe & Davis, 1982).

The suction force developed between the anchor and soil is a function of several variables,
such as: embedment depth, soil permeability, undrained shear strength and loading rate.
Therefore, with so many variables, Merifield et al. (2001) propose taking into account the
immediate breakaway case only, resulting in a conservative estimation of the actual pull-out
resistance with suction. Figure 91 shows the general layout out of the problem considered in the
analyses.

99
Offshore Foundations: Technologies, Design and Application

Figure 91 – Problem notation

Merifield et al. (2001) studied the influence of the failure mode in the ultimate resistance
of the anchor. The anchors are classified as shallow, if at collapse, the observed failure mechanism
reaches the surface, Figure 92(a) and (b). In contrast, a deep anchor is one whose failure mode is
characterized by localized shear around the anchor and is not affected by the location of the soil
surface, Figure 92(c).

For a given anchor geometry and soil type, there is a critical embedment depth, H cr, at
which the failure mechanism no longer extends to the soil surface and becomes localized around
the anchor. When this type of failure occurs, the ultimate capacity reaches a maximum limiting
value. Physically, this transition arises because the undrained shear strength is assumed to be
independent of the mean normal stress. From a practical point of view, this result is important as
embedding the anchor beyond Hcr will not lead to an appreciable increase in anchor capacity,
Figure 92(d).

Even though, the studies made by Merifield et al. (2001) cover most problems of practical
interest, the situation defined in this work is out of this range of theoretical solutions. These are
derived for problems where H/B ranges from 1 to 10 and ρB/su,o varies from 0.1 to 1. In this work
H/B=20/4.5=4.44, but ρB/su,o = 1.5x4.5/5 = 1.35.

100
Offshore Foundations: Technologies, Design and Application

Figure 92 – Shallow and deep anchor behaviour, Merifield et al. (2001)

The suggested procedure for estimation of uplift capacity is as follows:

1. Determine representative values of the material parameters s u,o (undrained shear


strength at the ground surface), γ (soil wet unit weight) and ρ (undrained shear
strength gradient).
2. Knowing the anchor size B (width) and embedment depth H, calculate the
embedment ratio H/B and overburden ratio γH/s u,o.
3. Calculate the break-out factor Nco using Figure 93 depending on the anchor shape.
4. Adopt Nc*=11.16 (for horizontal anchors)
5. In soil with 𝜌 > 0, calculate the break-out factor Ncoρ using:

𝜌𝐵 2𝐻
𝑁𝑐𝑜𝜌 = 𝑁𝑐𝑜 [1 + 0.383 ( − 1)] (26)
𝑠𝑢,0 𝐵

6. Calculate the break-out factor Nc using:

𝛾𝐻
𝑁𝑐 = 𝑁𝑐𝑜𝜌 + (27)
𝑠𝑢,0
7. Calculate the limiting value of the break-out factor 𝑁𝑐𝜌∗ using:

101
Offshore Foundations: Technologies, Design and Application

𝜌𝐻
𝑁𝑐𝜌∗ = 𝑁𝑐∗ [1 + ] (28)
𝑠𝑢,0
8. If 𝑁𝑐 ≥ 𝑁𝑐𝜌∗ then the anchor is a deep anchor. The ultimate pull-out capacity is
given by (25), where 𝑁𝑐 = 𝑁𝑐𝜌∗ .
9. If 𝑁𝑐 ≤ 𝑁𝑐𝜌∗ then the anchor is a deep anchor. The ultimate pull-out capacity is
given by (25), where 𝑁𝑐 is the value obtained in Equation (27).

Figure 93 – Design chart for rectangular anchors in clay, allows determining anchor break-out factor
Nco at various embedment ratios (Merifield et al., 2003).

DNV-RP-E302 (2002)
This design code is based on the results from the Joint Industry Projects “Reliability
analysis of deep water plate anchors” and “Design procedures for deep water anchors” (Dahlberg
et al., 2004). It applies to the geotechnical design and installation of plate anchors in NC clay,
including the suction embedded anchors, for the several types of moorings systems.

According to this code, the design of the anchor must account for installation effects on the
plate anchor resistance. In particular, the design must specify a minimum anchor penetration,
which should be verified during the anchor installation.

The safety requirements are based on the limit state method of design, and the design
criterion shall be satisfied by:

𝑅𝑑 (𝑧𝑖 ) − 𝑇𝑑 ≥ 0 (29)

102
Offshore Foundations: Technologies, Design and Application

Where 𝑅𝑑 (𝑧𝑖 ) is the design value for the anchor resistance and 𝑇𝑑 is the design value of the
mooring line tension. The design anchor resistance is given by:

𝑅𝑑 (𝑧𝑖 ) = 𝑁𝑐 ∙ 𝑠𝑐 ∙ 𝜂 ∙ 𝑠𝑢,𝑚𝑒𝑎𝑛 (𝑧𝑖 ) ∙ 𝐴𝑝𝑙𝑎𝑡𝑒 ∙ 𝑈𝑐𝑦 ∙ (1⁄𝛾𝑚 ) (30)

The parameters in eq. (30) are:

𝑈𝑐𝑦 Cyclic loading factor


= 𝜏𝑓,𝑐𝑦 /𝑠𝑢,𝑚𝑒𝑎𝑛
𝜏𝑓,𝑐𝑦 Cyclic shear strength
𝑠𝑢,𝑚𝑒𝑎𝑛 Mean static undrained shear strength within soil volume influencing the
anchor resistance
𝑁𝑐 Bearing capacity factor
𝑠𝑐 = 1 + 0.2 ∙ 𝑊𝐹 Shape factor
/𝐿𝐹
𝑊𝐹 Equivalent plate width
𝐿𝐹 Equivalent plate length
𝜂 Empirical reduction factor
𝐴𝑝𝑙𝑎𝑡𝑒 Plate area
𝑧𝑖 Installation penetration depth of plate
𝛾𝑚 Partial safety factor

Most anchors can utilise the beneficial effect of cyclic loading, which may increase the
anchor resistance by 10% to 20% in normally consolidated clay when subjected to storm loading,
because of the suction forces developed in the back of the plate anchor. For special environmental
loading conditions, such as loop currents or wind-generated squall conditions, the beneficial effect
of cyclic loading will be marginal. Therefore, in a conservative approach U cy is set to be equal to
1.0, since it normally is larger than 1.0.

The bearing capacity factor, Nc, is valid for plane strain conditions (strip footing) and
isotropic, incompressible material, and is adjusted for the plate geometry through the shape factor
sc. If the anchor is installed at a depth equal or superior to 4.5W F, then it is considered a deep
5.14

anchor and Nc is set to 12.0. However, if the installation depth is less than 4.5WF, the penetration
is considered shallow and the equation or graph in Figure 94 must be used to obtain the
appropriate bearing capacity factor.

103
Offshore Foundations: Technologies, Design and Application

12

Figure 94 – Equation and curve describing the variation of Nc value in shallow failure zone, Dahlberg
et al. (2004).

The factor η is an empirical reduction factor, set equal to 0.75. This factor accounts for:

 Progressive failure (strain-softening) of the clay within the soil volume involved as the
plate anchor is loaded to failure
 Strength anisotropy, to the extent the actual average undrained shear within the soil
volume affected by the failure differs from the assumed shear strength
 Load eccentricity if the anchor is not loaded normally to the plate

This factor may be higher or lower depending on laboratory soil strength test data. For
example, clay with a high sensitivity, which exhibits a significant strain-softening, may require a
lower η value, whereas one with insignificant strain-softening may use a higher η value. In
addition, a greater load eccentricity may also need a lower η factor.

The design criteria can be formulated in terms of two limit state equations:

a) An ultimate limit state (ULS) to ensure that the individual mooring lines have adequate
strength to withstand the load effects imposed by extreme environmental actions
b) An accidental limit state (ALS) to ensure that the mooring system has adequate
resistance to withstand the failure of one mooring line, failure of one thruster, or a
failure in the thruster system for unknown reasons.

There are also two different classes of consequences, which are considered for both ULS
and ALS:

104
Offshore Foundations: Technologies, Design and Application

1. Class 1 – Failure is unlikely to lead to unacceptable consequences such as loss of life,


collision with an adjacent platform, uncontrolled outflow of oil or gas, capsizing or sinking.
2. Class 2 – Failure may well lead to unacceptable consequences of the above types.

Limit State: ULS ALS


Consequence Class:
1 2 1 2
Partial safety factor
𝜸𝒎 1.40 1.40 1.00 1.30
Table 10 – Partial safety factor for anchor resistance, Dahlberg et al. (2004).

Predicted Plate Anchor Capacity


The predicted plate anchor capacity according to the three different methods is expressed
in Table 11, the calculation steps made to achieve those results are presented in Appendix III. The
Merifield et al. (2001) procedure returned an ultimate holding capacity 50% higher than the
others, and this may be explained by the two facts. Firstly, one parameter of the evaluated
𝜌𝐵
situation is outside of the range of the theoretical solutions, the value of must be between 0.1
𝑠𝑢,0

and 1.0, but in this case is equal to 1.35. Secondly, Merifield et al. (2001) do not use any empirical
reduction factor whereas the other two methods do; 0.70 is used by Wilde et al. (2001) and 0.75
in DNV-RP-E302 (2002).

Wilde et al. (2001) and DNV-RP-E302 (2002) have similar results, and the major source of
their difference is the usage of a partial safety factor by DNV-RP-E302 (2002), for that reason a
partial factor of 1.4 is also applied in Wilde et al. (2001) method to obtain a design holding
capacity.

For the design holding capacity in Merifield et al. (2001) method if safety factors (of 1.4)
are applied to both shear strength gradient (ρ) and initial shear strength (s u,0), the design holding
capacity will be about 3.5 MN higher than DNV-RP-E302 (2002). The difference is explained by
the fact that Merifield et al. (2001) does not use and empirical reduction factor and they include
de self-weight of the anchor in the plate resistance. If these were similarly taken into account for
Merifield et al. (2001) the resistance would be about 9.4 MN, which is quite similar to the Wilde
et al. (2001) and DNV-RP-E302 (2002) results.

Contrary to Merifield et al. (2001), Wilde et al. (2001) and DNV-RP-E302 (2002)
procedures do not take into account the self-weight of the plate anchor in their resistance. This
conservative decision means not taking into consideration 500 kN of resistance, which represents
4 to 5% of the anchors holding capacity.

105
Offshore Foundations: Technologies, Design and Application

As a result, is believed to be more reliable to use DNV-RP-E302 (2002) method, since it


seems to have a better calibration of its empirical reduction and safety factors. Thus, plate anchor
capacity adopted is 9.5 MN, determined after DNV-RP-E302 (2002) plus the anchor self-weight,
0.5 MN, therefore the predicted plate anchor capacity in this case study is predicted to be 10 MN.

Procedure suggested Ultimate Holding Design Holding


by: Capacity (MN) Capacity (MN)
Wilde et al. (2001) 13 9.2
Merifield et al. (2001) 18 13
DNV-RP-E302 (2002) 13.3 9.5
Table 11 – Plate anchor holding capacities in the idealized STP offshore conditions according to
three different methods.

6.4. DISCUSSION OF RESULTS


Both of the considered anchor solutions are likely to be economically competitive
alternatives to conventional offshore anchors, for application in STP. In Table 12 and Table 13 are
enumerated, respectively, the advantages and disadvantages of both SEPLA and torpedo anchors.

Torpedo type anchors are simple and economic to fabricate, and are able to be installed
easily and quickly.

On the other hand, the anchor element of the SEPLAs represents the lowest cost of all the
deep-water anchors (including torpedoes). Moreover, it only needs one follower (suction caisson)
to install any required number of anchors and due to their small size; multiple SEPLAs might be
carried on a single installation vessel in a single voyage (Ehlers et al., 2004). For SEPLA, all these
aspects make the system significantly less costly than e.g. compared to suction anchor systems;
the cost is about half to one-third (http://events.energetics.com/deepwater/pdfs/presentations/
session5/gengshenliu.pdf).

Both of the torpedo and the SEPLA solutions have a disadvantage in common which is their
patented installation method.

The installation methods for torpedo anchors and SEPLAs are very distinct. The latter uses
a proven installation methodology based on suction caisson technology that has been in use for
the last 30 years, this technique provides an accurate measure of embedment and position of the
anchor, before keying. The torpedo anchor installation is simply achieved by dropping the torpedo
from a certain height above the seafloor and letting it drive freely into the soil. Despite the
simplicity of the torpedo method, it has the disadvantage of the final orientation (inclination) of

106
Offshore Foundations: Technologies, Design and Application

the torpedo pile being unknown, which has been shown, can lead to the load resistance being
lower than expected.

In the economic aspect of the installation, torpedo anchors have a small advantage over
SEPLA, in that while they both need an Anchor Handling Vessel, the torpedo has no need of a
remote operating vehicle, while SEPLA does.

The geotechnical design of these anchor systems is very different as well. The design of
Torpedo anchors is patent protected by Petrobras, which means there are very few documented
installation and the design methods are unavailable to the public. However, it is known that a
computer program based on the True model (True, 1974) is used to evaluate the penetration of
the torpedo, and finite element analysis is used to carry out a detailed non-linear analysis to
predict the pull-out capacity of the pile.

The estimation of the penetration depends on several factors such as the layering of the soil,
the precise soil strength profile and the presence of stiff elements. These factors may be difficult
or impossible to identify and they have the potential to stop the penetration of the anchor or to
deviate the penetration path from the vertical. Furthermore, the pull-out capacity relies on the
adhesion between the soil and the shaft of the pile, the adhesion increases with the
reconsolidation of the soil in the vicinity of pile after installation. It is very difficult however to
quantify this increment without field trials and long term experience. Since there is no experience
outside Brazil, and the soils are not exactly similar to the soils present in Gulf of Guinea, it would
be very important to carry out those full-scale tests before adopting this solution.

Apart from the loss of embedment during keying, the position of the SEPLA is precisely
known, because it is positioned wherever it is needed in a controlled manner. SEPLA design is
based on classic limit equilibrium methods, for which the most difficult parameter to determine
is the bearing capacity factor. However, intense studies have been made to understand the
behaviour of plate anchor failure and to quantify the bearing capacity factor. Besides, the design
of the SEPLA being more reliable than the torpedo, the SEPLA have already been put into use in
the Gulf of Guinea, which can provide important information of the SEPLA behaviour in a similar
situation.

Finally, when comparing the resistance obtained by each foundation element, the SEPLA has
a greater advantage. After applying safety factors in the proposed SEPLA, solution the predicted
resistance is about 10 MN. For torpedo anchors the ultimate resistance is predicted to be 8.7 MN,
however the use of a safety factor of 2 is advised (Eltaher et al., 2003), which reduces the design

107
Offshore Foundations: Technologies, Design and Application

resistance to about 4.4 MN. Thus, double the number of torpedoes will be required to provide the
same load resistance as a single plate anchor.

In other words, the resistance obtained per unit weight of the anchor element is higher for
10000 𝑘𝑁 4350 𝑘𝑁
SEPLA, = 200 𝑘𝑁⁄𝑡𝑜𝑛𝑛𝑒; than for the torpedo anchor, = 44 𝑘𝑁⁄𝑡𝑜𝑛𝑛𝑒.
50 𝑡𝑜𝑛𝑛𝑒𝑠 98 𝑡𝑜𝑛𝑛𝑒𝑠

Even though, both torpedo and suction embedded plate anchors have smaller unit costs in
comparison with the other types of anchoring systems, one of them is more suitable than the other
to be recommended for use in this case-study. The best solution is the one which has more unit
resistance per foundation element, the design and installation method is more reliable, research
has provided a more fundamental basis for understanding the method, and if possible, should
already been used in the Gulf of Guinea or a similar geotechnical region. Taking into account the
above criteria, the more appropriate solution for this case study is considered to be the Suction
Embedded Plate Anchor.

SEPLA advantages Torpedo anchor (DPA) advantages


 Cost of anchor element is the
lowest of all the deep-water  Simple and economic to fabricate
anchors.  Simple to design.
 Uses proven suction caisson  Accurate to position with no
installation methods. requirements for proof loading.
 Provides an accurate measure of  Rapid installation
embedment and position of the  Robust and compact design makes
anchor. handling and installation simple and
 Design based on well-developed economic with only one Anchor
procedures for plate anchors. Handling Vessel (AHV) and no ROV.
 Experience in the Gulf of Guinea
Table 12 – SEPLA and Torpedo anchor advantages.

SEPLA disadvantages Torpedo anchor (DPA) disadvantages


 Patented installation method.  Patented installation method.
 Installation time greater than for a  No experience outside Brazil.
caisson.  Lack of documented installation and
 Requires keying and proof loading. design methods with verification
 Requires a ROV. agencies.
 Limited field load tests.  Unknown orientation once embedded.
Table 13 – SEPLA and Torpedo anchor disadvantages.

108
Offshore Foundations: Technologies, Design and Application

7. CONCLUSION AND FURTHER RESEARCH

7.1 CONCLUSION
This thesis has presented an outlined of existing platform and associated foundation
technologies used in offshore developments, and the most important aspects involved in their
design. Also, geotechnical characterization issues relating to the offshore environment such as
topography, seabed composition and geohazards were discussed.

In order to relate the foundation technologies discussed with a real scenario, a specific
region was chosen to assess the viability of two different foundation solutions. Proposed offshore
developments near São Tomé & Príncipe were chosen for the case study, because it is a region
that has recently gathered attention from the oil & gas industry, and has not been subjected to any
platform construction yet.

Water depths in the zone of proposed offshore development near São Tomé & Príncipe
range from 1800 m to 3000 m; thus, future installed facilities must be floating platforms and hence
the foundation systems in the seabed will be resisting tensile forces instead of compression.
Therefore, the only types of foundation solution suitable for this region are anchoring systems.
Anchoring systems that were selected for evaluation in this case are those for which fewer studies
and investigation have been made, but on the other hand are likely to be the most economic
systems – in this case, torpedo anchors and suction embedded plate anchors were evaluated.

Geotechnical characterization of the zones offshore from São Tomé & Príncipe was based on
investigations performed in the Gulf of Guinea over more than 10 years. This large database on
the behaviour of deep-water sediments made possible the assumption of several soil parameters.
The soil in the Gulf of Guinea typically is a highly sensitive clay (St = 2 to 4), and one of the most
important parameters is the undrained shear strength profile, it was apparent that it would have
a positive gradient with depth of about 1.5 kPa/m. It also became clear that many sites in this
region exhibit a greater resistance (up to about 15 kPa) in the first 2 m, this phenomenon is called
a “crust”, and no unique or convincing explanation has been proposed for its existence so far.
Other important properties, which are not well defined yet, are the interface soil-steel friction
resistance and the set-up effects.

The selected torpedo anchor to be employed in this case study would be the same that was
used in Albacora Lest Field in Brazil Basin by Petrobras, and it is the T-98 torpedo. The design of
the torpedo anchor was based on the design of simple cylindrical pile. Its pull-out capacity comes
from three different sources: shaft friction resistance, self-weight of the pile and reverse end

109
Offshore Foundations: Technologies, Design and Application

bearing capacity. The shaft friction resistance is the fraction that gives the greatest contribution,
thus the correct assessment of the adhesion factor (α) plays an important role.

The assessment of the torpedo anchor free penetration into the soil is also important to
recognise the torpedo embedment, and therefore the soil shear strength along the torpedo shaft.
Following O’Loughlin et al. (2013), after releasing the torpedo from 108 m above the seafloor, it
should reach a velocity of about 40 m/s before impacting with the soil and the torpedo should
penetrate about 40 m into the soil. An adhesion factor, α=0.7 was considered, which corresponds
to full reconsolidation of the soil in the vicinity of the pile after being remoulded by the torpedo
penetration, consequently the pull-out resistance is expected to be 8.7 MN.

The SEPLA solution considered is a 4.5 m x 10 m plate anchor, which is proposed by Wilde
et al. (2001) for permanent installations. Plate anchor keying induced loss of embedment was
calculated according to Song et al. (2009) and is expected to be about 0.81B (B is the anchor
breadth), i.e. approximately 4 m.

The holding capacity of the SEPLA is provided by the end bearing resistance plus the self-
weight of the anchor. This capacity was calculated according to three different design procedures:

 Wilde et al. (2001)


 Merifield et al. (2001)
 DNV-RP-E302 (2002)

The most conservative procedure proved to be from DNV-RP-E302 (2002), with a holding
capacity of about 9.5 MN. However, this do not take into account the self-weight of the anchor and
includes a partial resistance factor, therefore an additional 0.5 MN may be added to provide a total
resistance of 10 MN.

Based on a criteria that involved the resistance, installation process, experience, knowledge
and reliability of both anchoring systems, is was considered that the use of suction embedment
plate anchor systems (SEPLA 4 x 10 m2) would be more appropriate offshore from São Tome &
Principe.

7.2. FURTHER RESEARCH


In terms of geotechnical issues associated with this case study, there is still need for further
studies, in particular:

 Site specific characterisation of the seabed in the São Tomé & Principe region,
instead of the broad characterisation of the Gulf of Guinea made in this study and
based on the amalgamation of various published data from the Gulf generally.

110
Offshore Foundations: Technologies, Design and Application

 Understanding of the origin and characterisation of the near seabed “crust”


particular to this region and the effect this may have on foundation installation
processes.
 The interface friction resistance between the soil and the steel elements in the short
and long-term.

Since there is no experience of torpedo anchors in the Gulf of Guinea, it would be of great
interest to develop in situ model scale tests to study the behaviour of the torpedo during
penetration of the soil, and its resistance in the short and long-term as well. As this in situ tests
are very expensive, it would be very interesting to evaluate the influence of the near seabed “crust”
on the torpedo penetration, using for that computational programs or laboratorial tests.

A modulation of both SEPLA and torpedo solutions would be interesting to assess a more
efficient evaluation of their potential resistance in STP.

A study to evaluate the forces that the platforms will be subjected in the offshore of São
Tomé & Principe would be of great interest, because depending on those forces the foundation
solutions may be more or less economic, i.e. number and size of the foundation elements.

111
Offshore Foundations: Technologies, Design and Application

112
Offshore Foundations: Technologies, Design and Application

REFERENCES
AGUIAR, C. S., DE SOUSA, J.R., AND ELLWANGER, G. B. (2009), “ANÁLISE DA INTERAÇÃO SOLO-ESTRUTURA DE
ÂNCORAS DO TIPO TORPEDO PARA PLATAFORMAS OFFSHORE”, PHD THESIS, UFRJ.

AGUIAR, C. S., DE SOUSA, J.R., AND ELLWANGER, G. B. (2011), “UNDRAINED LOAD CAPACITY OF TORPEDO
ANCHORS IN COHESIVE SOILS”, PROCEEDINGS OF THE ASME 28TH OMAE CONFERENCE, 79465,
HONOLULU.

ALDRIDGE, T.R., CARRINGTON, T.M. AND KEE, N.R. (2005). “PROPAGATION OF PILE TIP DAMAGE DURING
INSTALLATION”. PROC. INT. CONF. FRONTIERS IN OFFSHORE GEOTECHNICS, EDS S. GOURVENEC AND M.

CASSIDY, TAYLOR AND FRANCIS, 823—827.

AMERICAN PETROLEUM INSTITUTE (2010). “API RP 2A: RECOMMENDED PRACTICE FOR PLANNING,
DESIGNING AND CONSTRUCTING FIXED OFFSHORE PLATFORMS – WORKING STRESS DESIGN”.

ARAÚJO, J. B., MACHADO, R. D., AND MEDEIROS JR., C. P. (2004), “HIGH HOLDING POWER TORPEDO PILE –
RESULTS FOR THE FIRST LONG TERM APPLICATION”, PROCEEDINGS OF THE ASME 23RD OMAE
CONFERENCE, 51201, VANCOUVER.

BARBOUR, R.J. AND ERBRICH C.T. (1994). “ANALYSIS OF IN-SITU REFORMATION OF FLATTENED LARGE
DIAMETER FOUNDATION PILES USING ABAQUS”. PROC. UK ABAQUS USERS CONF., OXFORD.

BRANDÃO, F. E. N., HENRIQUES, C. C. D., ARAÚJO, J. B., FERREIRA, O. C. G. & AMARAL, C. D. S. (2006).
“ALBACORA LESTE FIELD DEVELOPMENT: FPSO P-50 MOORING SYSTEM CONCEPT AND INSTALLATION”.
PROC. OFFSHORE TECHNOL. CONF., HOUSTON, TX, PAPER OTC 18243.

BRAUSSE, M. (2001). “UNDERCONSOLIDATION AND APPARENT OVERCONSOLIDATION IN MARINE SEDIMENTS: A


STUDY OF STRESS STATES IN THE NORTHWESTERN GULF OF MEXICO”. CIVIL ENGINEERING MSC THESIS,

UNIVERSITY OF RHODE ISLAND.

BRAUSSE, M. (2001). “UNDERCONSOLIDATION AND APPARENT OVERCONSOLIDATION IN MARINE SEDIMENTS: A


STUDY OF STRESS STATES IN THE NORTHWESTERN GULF OF MEXICO”. CIVIL ENGINEERING MSC THESIS,

UNIVERSITY OF RHODE ISLAND.

BS ISO 19903:2006, PETROLEUM AND NATURAL GAS INDUSTRIES (2006). “FIXED CONCRETE OFFSHORE
STRUCTURES”.

CAMPBELL, K.J., HUMPHREY, G.D. AND LITTE, R.L. (2008). “MODERN DEEPWATER SITE INVESTIGATION:
GETTING IT RIGHT THE FIRST TIME”. PAPER OTC 19535, OFFSHORE TECHNOLOGY CONFERENCE.

113
Offshore Foundations: Technologies, Design and Application

CARTER, J. P., NAZEM, M. (2013), “ANALYSIS OF DYNAMIC PENETRATION OF SOILS”. CENTRE FOR
GEOTECHNICAL AND MATERIAL MODELLING, THE UNIVERSITY OF NEWCASTLE, CALLAGHAN NSW,
AUSTRALIA.

CHRISTOPHERSEN, H.P. (1993). “THE NON-PILED FOUNDATION SYSTEMS OF THE SNORRE FIELD. OFFSHORE
SITE INVESTIGATION AND FOUNDATION BEHAVIOUR”, SOC. FOR UNDERWATER TECHNOLOGY, 28, 433-
447.

CLAUSEN, C.J.F., DIBIAGIO, E., DUNCAN, J.M. AND ANDERSEN, K.H. (1975). “OBSERVED BEHAVIOUR OF THE
EKOFISK OIL STORAGE TANK FOUNDATION”. PROC. ANNUAL OFFSHORE TECHNOLOGY CONF., HOUSTON,
OTC 2373.

CLUKEY, E.C., MORRISON, M.J., GARNIER, J. AND CORTE, J.-F. (1995). “THE RESPONSE OF SUCTION CAISSONS
IN NORMALLY CONSOLIDATED CLAYS TO CYCLIC TLP LOADING CONDITIONS”. PAPER OTC 7796,
OFFSHORE TECHNOLOGY CONFERENCE.

COLLIAT, J.-L. & DENDANI, H., PUECH, A. AND NAUROY, J.-F. (2011). “GULF OF GUINEA DEEPWATER
SEDIMENTS: GEOTECHNICAL PROPERTIES,DESIGN ISSUES AND INSTALLATION EXPERIENCES”. RONTIERS

IN OFFSHORE GEOTECHNICS II – GOURVENEC & WHITE: 59-86.

DADE, B.W AND HUPPERT, H.E. (1998). “LONG-RUNOUT ROCKFALLS”. GEOLOGY, 26(9): 803-806.

DAHLBERG, R., RONOLD, K. O., STRØM, P. J., & MATHISEN, J. (2004). “NEW CALIBRATED DESIGN CODE FOR
PLATE ANCHORS IN CLAY”. OFFSHORE TECHNOLOGY CONFERENCE. DOI:10.4043/16109-MS.

DE BLASIO, F.V., ELVERHOI, A., ENGVIK, L.E., ISSLER, D., GAUER, P. AND HARBITZ, C.B. (2006).
“UNDERSTANDING THE HIGH MOBILITY OF SUBAQUEOUS DEBRIS FLOWS”. NORW. J. GEOL., 86: 275-
284.

DE GENNARO, V., PUECH, A. & DELAGE, P. (2005). “ON THE COMPRESSIBILITY OF DEEPWATER SEDIMENTS OF
THE GULF OF GUINEA” .PROC. 1ST INT. SYMPOSIUM ON FRONTIERS IN OFFSHORE GEOTECHNICS,
ISFOG, PERTH.

DEAN, E. T. (2009). “OFFSHORE GEOTECHNICAL ENGINEERING”. THOMAS TELFORD.

DENDANI, H. & JAECK, C. (2007). “PIPE-SOIL INTERACTION IN HIGHLY PLASTIC CLAYS”. PROC. INT.
CONFERENCE ON OFFSHORE SITE INVESTIGATION AND GEOTECHNICS, SUT OSIG, LONDON.

DNV-OS-C502 (2012). “OFFSHORE CONCRETE STRUCTURES”. DET NORSK VERITAS.

DNV-RP-E302 (2002). “DESIGN AND INSTALLATION OF PLATE ANCHORS IN CLAY- RECOMMENDED


PRACTICE”, DET NORSK VERITAS, 43 PAGES.

114
Offshore Foundations: Technologies, Design and Application

DOVE, P., TREU, H., AND WILDE, B. (1998). “SUCTION EMBEDDED PLATE ANCHOR (SEPLA): A NEW
ANCHORING SOLUTION FOR ULTRA-DEEP WATER MOORING.” PROC., D.O.T. 10TH INT. CONF. AND
EXHIBITION, DEEP OFFSHORE TECHNOLOGY, LONDON.

EDGERS, L. AND KARLSRUD, K. (1982). “SOIL FLOWS GENERATED BY SUBMARINE SLIDES - CASE STUDIES AND
CONSEQUENCES”. PROC. INT. CONF. BEHAV. OFFSHORE STRNCT. (BOSS), CAMBRIDGE,
MASSACHUSETTS, 2: 425-437.

EHLERS, C. J., YOUNG, A. G., CHEN, J. (2004). “TECHNOLOGY ASSESSMENT OF DEEPWATER ANCHORS.”
PROCEEDINGS OF THE 36TH ANNUAL OFFSHORE TECHNOLOGY CONFERENCE, HOUSTON, TEXAS, 3-6
MAY 2004, PAPER NO. OTC 16840.

EHLERS, C.J., CHEN, J., ROBERTS, H.H. & LEE, Y.C. (2005). “THE ORIGIN OF NEAR-SEAFLOOR ‘CRUST ZONES’ IN
DEEPWATER”. PROC. 1ST INT. SYMPOSIUM ON FRONTIERS IN OFFSHORE GEOTECHNICS, ISFOG, PERTH.

ELTAHER, A., RAJAPASKA, Y., AND CHANG, K. T. (2003). “INDUSTRY TRENDS FOR DESIGN OF ANCHORING
SYSTEMS FOR DEEPWATER OFFSHORE STRUCTURES”. PROC., OFFSHORE TECHNOLOGY CONF., OTC NO.

15265.

ERBRICH, C.T. AND TJELTA, T.I. (1999). “INSTALLATION OF BUCKET FOUNDATIONS AND SUCTION CAISSONS IN
SAND – GEOTECHNICAL PERFORMANCE”. PROC. ANNUAL OFFSHORE TECHNOLOGY CONF., HOUSTON,
PAPER OTC 10990.

FERNANDES, A. C., DOS SANTOS, M. F., ARAÚJO, J. B., ALMEIDA, J. C. L., DINIZ, R. AND MATOS, V. (2005).
“HYDRODYNAMIC ASPECTS OF THE TORPEDO ANCHOR INSTALLATION”. PROC. INTERNATIONAL
CONFERENCE ON OFFSHORE MECHANICS AND ARTIC ENGINEERING, HALKIDIKI, GREECE, OMAE 2005-
67201.

FREEMAN, T. J. AND HOLLISTER, C. D. (1988). “MODELLING WASTE EMPLACEMENT AND THE PHYSICAL
CHANGES TO THE SEDIMENT BARRIER. PROC. INTERNATIONAL CONFERENCE ON THE DISPOSAL OF
RADIOACTIVE WASTE IN SEABED SEDIMENTS, OXFORD, 199-219.

FREEMAN, T. J., BURDETT, J. R. F., (1986). “DEEP OCEAN MODEL PENETRATOR EXPERIMENTS.” FINAL REPORT
TO COMMISSION OF THE EUROPEAN COMMUNITIES, CONTRACT NO. 392-83-7-WAS-UK.

FREEMAN, T. J., MURRAY, C. N., FRANCIS, T. J. G., MCPHAIL, S. D., AND SCHULTHEISS, P. J. (1984). “MODELING
RADIOACTIVE WASTE DISPOSAL BY PENETRATOR EXPERIMENTS IN THE ABYSSAL ATLANTIC
OCEAN.”NATURE, 310, 130–133.

115
Offshore Foundations: Technologies, Design and Application

FRENCH, L. S., RICHARDSON, G. E., KAZANIS, E. G., MONTGOMERY, T.M., BOHANNON, C.M. AND GRAVOIS, M. P.
(2006). "DEPWATER GULF OF MEXICO 2006: AMERICA’S EXPANDING FRONTIER. U.S. DEPARTMENT
OF INTERIOR, MINERALS MANAGEMENT SERVICE, REPORT NO. MMS 2006-022.

GAUDIN, C., O’LOUGHLIN, C. D., RANDOLPH, M. F. AND LOWMASS, A. (2006), “INFLUENCE OF THE
INSTALLATION PROCESS ON THE PERFORMANCE OF SUCTION EMBEDDED PLATE ANCHORS”.

GÉOTECHNIQUE, VOL. 56, NO. 6, 381-391.

GERWICK, B. C. (2007). “CONSTRUCTION OF MARINE AND OFFSHORE STRUCTURES”. BOCA RATON: CRC PRESS.

HAMPTON, M.A., LEE, H.J., AND LOCAT, J. (1996). “SUBMARINE LANDSLIDES”. REV. GEOPHYS., AMERICAN
GEOPHYSICAL UNION, 34(1), 33-59.

HASANLOO, D., & YU, G. (2011). “A STUDY ON THE FALLING VELOCITY OF TORPEDO ANCHORS DURING
ACCELERATION”. IN THE TWENTY-FIRST INTERNATIONAL OFFSHORE AND POLAR ENGINEERING
CONFERENCE. INTERNATIONAL SOCIETY OF OFFSHORE AND POLAR ENGINEERS.

HEIM, A. (1932) “BERGSTURZ UND MENSCHENLEBEN”. FRETZ UND WASMUTH, ZURICH, SWITZERLAND.

HICKERSON, J., FREEMAN, T. J., BOISSON, J., GERA, F., MURRAY, N., NAKAMURA H., NIEUWENHUIS, J. D.,
SCHULLER, K. H. (1988). “FEASIBILITY OF DISPOSAL OF HIGH-LEVEL RADIOACTIVE WASTE INTO THE
SEABED.” VOLUME 4: ENGINEERING, NUCLEAR ENERGY AGENCY, PARIS.

HOULSBY, G.T. AND BYRNE, B.W. (2005) “DESIGN PROCEDURES FOR INSTALLATION OF SUCTION CAISSONS IN
CLAY AND OTHER MATERIALS”, PROC. ICE, GEOTECHNICAL ENG., VOL. 158 NO. GE2, PP 75–82.

KEUNEN, PH. H., (1950). “MARINE GEOLOGY”, JOHN WILEYS & SONS, INC., NEW YORK

KOLK, H.J. (2000). “DEEP FOUNDATIONS IN CALCAREOUS SEDIMENTS” .PROC. INT. CONF. ENGINEERING FOR
CALCAREOUS SEDIMENTS, ED. K.A. AL-SHAFEI, BALKEMA, 2, 313—344.

KOMURKA, V. E., WAGNER, A. B., (2003). “ESTIMATING SOIL/PILE SET-UP”. WISCONSIN HIGHWAY RESEARCH
PROGRAM – FINAL REPORT.

KULHAWY, F.H. AND MAYNE, P.W., (1990). “EL-6800: MANUAL FOR ESTIMATING SOIL PROPERTIES FOR
FOUNDATION DESIGN”, ELECTRIC POWER RESEARCH INSTITUTE, PALO ALTO, CALIFORNIA, USA.

KUNITAKI, D. M. K. N., DE LIMA, B. S. L. P., EVSUKOFF, A. G., & JACOB, B. P. (2008). “PROBABILISTIC AND FUZZY
ARITHMETIC APPROACHES FOR THE TREATMENT OF UNCERTAINTIES IN THE INSTALLATION OF TORPEDO

PILES”. MATHEMATICAL PROBLEMS IN ENGINEERING, VOL. 2008, ARTICLE ID 512343, 26 PAGES.

116
Offshore Foundations: Technologies, Design and Application

KVALSTAD, T.]., NADIM, E AND HARBITZ, E.B. (2001) “DEEPWATER GEOHAZARDS: GEOTECHNICAL CONCERNS
AND SOLUTIONS”. PROC. ANNU. OFFSHORE TECH. CONF., HOUSTON, TEXAS, PAPER OTC 12958.

LIENG, J. T., HOVE, F., TJELTA, T. I. (1999). “DEEP PENETRATING ANCHOR: SUBSEABED DEEPWATER ANCHOR
CONCEPT FOR FLOATERS AND OTHER INSTALLATIONS.” PROCEEDINGS OF THE 9TH INTERNATIONAL
OFFSHORE AND POLAR ENGINEERING CONFERENCE, BREST, FRANCE, 30 MAY- 4 JUNE 1999, VOL. I,
PP. 613-619.

LIENG, J. T., KAVLI, A., HOVE, F., TJELTA, T.I. (2000). “DEEP PENETRATING ANCHOR: FURTHER DEVELOPMENT,
OPTIMIZATION AND CAPACITY VERIFICATION.” PROCEEDINGS OF THE 10TH INTERNATIONAL
OFFSHORE AND POLAR ENGINEERING CONFERENCE, SEATTLE, WASHINGTON, 28 MAY- 2 JUNE 2000,
PP 410-416.

MASCLE, J., LOHMANN, G.P., AND MOULLADE, M. (EDS.) (1998). PROC. ODP, SCI. RESULTS, 159: COLLEGE
STATION, TX (OCEAN DRILLING PROGRAM). DOI:10.2973/ODP.PROC.SR.159.1998.

MCCLELLAND, B., (1974). “DESIGN OF DEEP PENETRATION PILES FOR OCEAN STRUCTURES”. ASCE JOURNAL OF
THE GEOTECHNICAL ENGINEERING DIVISION, 100(GT7), 705—747.

MEDEIROS JR., C. J., HASSUI, L. H. MACHADO, R. D., (1997). “PILE FOR ANCHORING FLOATING STRUCTURES
AND PROCESS FOR INSTALLING THE SAME.” UNITED STATES PATENT NUMBER 6,106,199.

MEDEIROS, C.J. (2001). “TORPEDO ANCHOR FOR DEEP WATER”. PROC. DEEPWATER OFFSHORE TECHNOLOGY
CONF., RIO DE JANEIRO.

MEDEIROS, C.J. (2002). “LOW COST ANCHOR SYSTEM FOR FLEXIBLE RISERS IN DEEP WATERS”. PROC. ANNUAL
OFFSHORE TECHNOLOGY CONF., HOUSTON, PAPER OTC 14151.

MERIFIELD, R. S., SLOAN, S. W. & YU, H. S. (2001). “STABILITY OF PLATE ANCHORS IN UNDRAINED CLAY”
GÉOTECHNIQUE 51, NO. 2, 141-153.

MSL, 2001. “A STUDY OF PILE FATIGUE DURING DRIVING AND IN-SERVICE AND OF PILE TIP INTEGRITY”,
OFFSHORE TECHNOLOGY REPORT 2001/018. PREPARED BY MSL ENGINEERING LIMITED FOR THE
HEALTH AND SAFETY EXECUTIVE, HSE BOOKS.

O’LOUGHLIN, C. D., RANDOLPH, M. F. & RICHARDSON, M. D. (2009). “CENTRIFUGE TESTS ON DYNAMICALLY


INSTALLED ANCHORS”. PROC. 28TH INT. CONF. ON OCEAN, OFFSHORE AND ARCTIC ENGINEERING,
HONOLULU, PAPER OMAE2009-80238.

117
Offshore Foundations: Technologies, Design and Application

O’LOUGHLIN, C. D., RANDOLPH, M. F., RICHARDSON, M. (2004). “EXPERIMENTAL AND THEORETICAL STUDIES
OF DEEP PENETRATING ANCHORS.” PROCEEDINGS OF THE 36TH ANNUAL OFFSHORE TECHNOLOGY
CONFERENCE, HOUSTON, TEXAS, MAY 3-6, 2004, PAPER NO. OTC 16841.

O’LOUGHLIN, C. D., RICHARDSON, M. D., RANDOLPH, M. F., & GAUDIN, C. (2013). “PENETRATION OF
DYNAMICALLY INSTALLED ANCHORS IN CLAY”. GÉOTECHNIQUE, 63(11), 909-919.

O’LOUGHLIN, C.D., LOWMASS, A.C., GAUDIN, C., RANDOLPH, M.F. (2006). “PHYSICAL MODELLING TO ASSESS
KEYING CHARACTERISTICS OF PLATE ANCHORS”. PROCEEDINGS OF THE 6TH INTERNATIONAL
CONFERENCE PHYSICAL MODELLING IN GEOTECHNICS, PP. 659–665.

ØYE, I. (2000). “SIMULATION OF TRAJECTORIES FOR A DEEP PENETRATING ANCHOR”. CFD NORWAY, REPORT
NO. 250:2000.

PEUCHEN, J. AND RAPP, J. (2007). “LOGGING SAMPLING AND TESTING FOR OFFSHORE GEOHAZARDS”, PROC.
ANNUAL OFFSHORE TECH. CONF., HOUSTON, PAPER OTC 18664.

POULOS, H.G., (1988). “MARINE GEOTECHNICS”, UNWIN.

POWER, P., GALAVAZI, M. AND WOOD, G., (2005). “GEOHAZARDS NEED NOT BE — REDEFINING PROJECT RISK”.
PAPER OTC 17634, OFFSHORE TECHNOLOGY CONFERENCE.

PUECH, A., DENDANI, H., NAUROY, J-F. AND MEUNIER, J. (2004). “CHARACTERISATION OF GULF OF GUINEA
DEEPWATER SOILS FOR GEOTECHNICAL ENGINEERING: SUCCESSES AND CHALLENGES”. 21-22 OCTOBRE

2004, SEATECH WEEK COLLOQUE “CARACTÉRISATION IN SITU DES SOLS MARINS” BREST, FRANCE.

QUIROS, G.W. & LITTLE, R 2003. “DEEPWATER SOIL PROPERTIES AND THEIR IMPACT ON THE GEOTECHNICAL
PROGRAM”. PROC. OFFSHORE TECHNOLOGY CONFERENCE, OTC PAPER 15262, HOUSTON.

RAIE, M. S. (2009). “A COMPUTATIONAL PROCEDURE FOR SIMULATION OF TORPEDO ANCHOR INSTALLATION,


SET-UP AND PULL-OUT”. CIVIL ENGINEERING PHD THESIS, UNIVERSITY OF TEXAS.

RAIE, M. S., & TASSOULAS, J. L. (2009). “INSTALLATION OF TORPEDO ANCHORS: NUMERICAL MODELING”.
JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING, 135(12), 1805-1813.

RANDOLPH M.F. AND HOUSE A.R. (2002). “ANALYSIS OF SUCTION CAISSON CAPACITY IN CLAY”. PROC. ANNUAL
OFFSHORE TECHNOLOGY CONF., HOUSTON, PAPER OTC 14236.

RANDOLPH, M. F., CARTER, J. P., AND WROTH, C. P. (1979). “DRIVEN PILES IN CLAY—THE EFFECTS OF
INSTALLATION AND SUBSEQUENT CONSOLIDATION.” GEOTECHNIQUE, 29(4), 361–393.

118
Offshore Foundations: Technologies, Design and Application

RANDOLPH, M. F., GAUDIN, C., GOURVENEC, S. M., WHITE, D. J., BOYLAN, N., & CASSIDY, M. J. (2011). RECENT
ADVANCES IN OFFSHORE GEOTECHNICS FOR DEEP WATER OIL AND GAS DEVELOPMENTS. OCEAN
ENGINEERING, 38(7), 818-834.

RANDOLPH, M., CASSIDY, M., GOURVENEC, S., & ERBRICH, C. (2005, SEPTEMBER). “CHALLENGES OF OFFSHORE
GEOTECHNICAL ENGINEERING”. IN PROCEEDINGS OF THE INTERNATIONAL CONFERENCE ON SOIL
MECHANICS AND GEOTECHNICAL ENGINEERING (VOL. 16, NO. 1, P. 123). AA BALKEMA
PUBLISHERS.

RICHARDSON, M. D. (2008). “DYNAMICALLY INSTALLED ANCHORS FOR FLOATING OFFSHORE STRUCTURES”.


PHD THESIS, THE UNIVERSITY OF WESTERN AUSTRALIA, CRAWLEY, AUSTRALIA.

RICHARDSON, M. D., O’LOUGHLIN, C. D., RANDOLPH, M. F. AND GAUDIN, C. (2009). “SETUP FOLLOWING
INSTALLATION OF DYNAMIC ANCHORS IN NORMALLY CONSOLIDATED CLAY”. JOURNAL OF
GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING. ASCE, SUBMITTED.

ROWE, R. K. & DAVIS, E. H. (1982). “THE BEHAVIOUR OF ANCHOR PLATES IN CLAY”. GÉOTECHNQIUE 32, NO. 1,
9-23.

RUINEN, R.M. (2005). “INFLUENCE OF ANCHOR GEOMETRY AND SOIL PROPERTIES ON NUMERICAL MODELING
OF DRAG ANCHOR BEHAVIOR ON SOFT CLAY! .PROC. INT. CONF. FRONTIERS IN OFFSHORE GEOTECHNICS,

EDS S. GOURVENEC AND M.J. CASSIDY, TAYLOR AND FRANCIS, 165—169.

SABETAMAL, H., NAZEM, M., CARTER, J. P., & SLOAN, S. W. (2014). “LARGE DEFORMATION DYNAMIC ANALYSIS
OF SATURATED POROUS MEDIA WITH APPLICATIONS TO PENETRATION PROBLEMS”. COMPUTERS AND

GEOTECHNICS, 55, 117-131.

SCHEIDEGGER, A.E. (1973) “ON THE PREDICTION OF THE REACH AND VELOCITY OF CATASTROPHIC
LANDSLIDES”. ROCK MECH., 5: 231-236.

SILVA, A.J. (1974). “MARINE GEOMECHANICS: OVERVIEW AND PROJECTIONS”. IN: DEEP SEA SEDIMENTS. A.L.
INDERBITZEN (ED.), PLENUM PRESS, NEW YORK.

SONG, Z., HU, Y., O’LOUGHLIN, C., & RANDOLPH, M. F. (2009). “LOSS IN ANCHOR EMBEDMENT DURING PLATE
ANCHOR KEYING IN CLAY”. JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING,

135(10), 1475-1485.

STOVE, O.J., BYSVEEN, S. AND CHRISTOPHERSEN, H.P. (1992). “NEW FOUNDATION SYSTEMS FOR THE SNORRE
DEVELOPMENT”. PROC. ANNUAL OFFSHORE TECHNOLOGY CONF., HOUSTON, PAPER OTC 6882.

119
Offshore Foundations: Technologies, Design and Application

STROUT, J.M. AND TJELTA, T.I., (2007). “EXCESS PORE PRESSURE MEASUREMENT AND MONITORING FOR
OFFSHORE INSTABILITY PROBLEMS”. PAPER OTC 18706, OFFSHORE TECHNOLOGY CONFERENCE.

SULTAN, N., COCHONAT, P., CAUQUIL, E. & COLLIAT, J.L. (2001). “APPARENT OVERCONSOLIDATION AND
FAILURE MECHANISMS IN MARINE SEDIMENTS” .PROC. OTRC INT. CONFERENCE HONORING PROF. W.

DUNLAP, HOUSTON.

SULTAN, N., VOISSET, M., MARSSET, T., VERNANT, A.M., CAUQUIL, E., COLLIAT, J.L. & CURINIER, V. (2007).
“DETECTION OF FREE GAS AND GAS HYDRATE BASED ON 3D SEISMIC DATA AND CONE PENETRATION
TESTING: AN EXAMPLE FROM THE NIGERIAN CONTINENTAL SLOPE” .MARINE GEOLOGY, 240, 235–255.

TRUE, D. G. (1974). “RAPID PENETRATION INTO SEAFLOOR SOILS.” PROCEEDINGS OF THE 6TH ANNUAL
OFFSHORE TECHNOLOGY CONFERENCE, HOUSTON, TEXAS, MAY 1974, PAPER NO. OTC 2095.

TRUE, D. G. (1976). “UNDRAINED VERTICAL PENETRATION INTO OCEAN BOTTOM SOILS.” PHD DISSERTATION,
UNIVERSITY OF CALIFORNIA, BERKELEY, CALIFORNIA.

WANG, D., GAUDIN, C., & RANDOLPH, M. F. (2013). “LARGE DEFORMATION FINITE ELEMENT ANALYSIS
INVESTIGATING THE PERFORMANCE OF ANCHOR KEYING FLAP”. OCEAN ENGINEERING, 59, 107-116.
ISO 690.

WATSON, P. G., RANDOLPH, M. F. AND BRANSBY, M. F. (2000), “COMBINED LATERAL AND VERTICAL LOADING
OF CAISSON FOUNDATIONS”. PROCEEDINGS, OFFSHORE TECHNOLOGY CONFERENCE, HOUSTON, USA,
OTC 12195.

WILDE, B., TREU, H., AND FULTON, T. (2001). “FIELD TESTING OF SUCTION EMBEDDED PLATE ANCHORS”.
PROC., 11TH INT. OFFSHORE AND POLAR ENGINEERING CONF., INT. SOCIETY OF OFFSHORE AND POLAR
ENGINEERS, MOUNTAIN VIEW, CA, 544–551.

YANG, M., AUBENY, C. P., & MURFF, J. D. (2011). “BEHAVIOUR OF SUCTION EMBEDDED PLATE ANCHORS DURING
KEYING PROCESS”. JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING, 138(2),
174-183.

120
Offshore Foundations: Technologies, Design and Application

Appendixes
INDEX:
Appendix I – Details of Shallow Foundation Studies………………………………………………. 136
Appendix II – Calculation to determine Loss in Anchor Embedment……………………….. 138
Appendix III – Calculations to determine the resistance of the SEPLA…………………….. 140

121
Offshore Foundations: Technologies, Design and Application

122
Offshore Foundations: Technologies, Design and Application

APPENDIX I – DETAILS OF SHALLOW FOUNDATION STUDIES.

123
Offshore Foundations: Technologies, Design and Application

124
Offshore Foundations: Technologies, Design and Application

APPENDIX II – CALCULATION TO DETERMINE LOSS IN ANCHOR


EMBEDMENT.
INITIAL MOMENT M0 CALCULATION:
𝑀0 = (𝑓 + 𝑊 ′ 𝑎 )𝑒 − 𝑓𝑒𝑓 − 𝑊′𝑎 𝑒𝑤

𝑊 ′ 𝑎 = 𝑊𝑎 − 𝑉𝑎 × 𝛾𝑤 = 500 − 9.58 × 10 = 404.2 𝑘𝑁

4 × 2.9
𝑉 = 4.5 × 10 × 0.2 + × 0.05 × 2 = 9.58 𝑚3
2

𝑓 = 𝐴𝑝𝑙𝑎𝑡𝑒 × 𝑠𝑢 = 45 × 1.5 × 24 = 1620 𝑘𝑁

𝑒𝑠ℎ𝑎𝑛𝑘 × 𝑉𝑠ℎ𝑎𝑛𝑘 2.5/3 × 0.58


𝑒𝑤 = = 0.05 𝑚
𝑉𝑠ℎ𝑎𝑛𝑘 + 𝑉𝑝𝑙𝑎𝑡𝑒 9 + 0.58

𝑒𝑓 = 0.1 𝑚

𝑒 = 2.5 + 0.1 = 2.6 𝑚

𝑀0 = (1620 + 404.2)2.6 − 1620 × 0.1 − 404.2 × 0.05 = 5080.71 𝑘𝑁. 𝑚

DETERMINATION OF THE LOSS IN ANCHOR EMBEDMENT, ΔZE/B:


∆𝑧𝑒 0.2
=
𝐵 𝑒 𝑡 0.3
𝑀 0.1
( )( ) ( 0 )
𝐵 𝐵 𝐴𝐵𝑠𝑢

𝑒 2.6
= = 0.58
𝐵 4.5

𝑡 0.2
= = 0.04
𝐵 4.5

𝑀0 5080.71
= = 0.70
𝐴𝑝𝑙𝑎𝑡𝑒 𝐵𝑠𝑢 45 × 4.5 × 24 × 1.5

∆𝑧𝑒 0.2
= = 0.94
𝐵 (0.58)(0.04)0.3 (0.70)0.1

DETERMINE THE CONSTANT CΘ:


∆𝑧𝑒
𝐶𝜃 = − 𝑘𝜃 ∙ 𝜃 = 0.94 − 0.005 × 90 = 0.49
𝐵

125
Offshore Foundations: Technologies, Design and Application

126
Offshore Foundations: Technologies, Design and Application

APPENDIX III – CALCULATIONS TO DETERMINE THE RESISTANCE OF


THE SEPLA.

WILDE ET AL. (2001)


𝐵
𝐹𝑛 = 𝐴𝑝𝑙𝑎𝑡𝑒 × 𝑁𝑐 × 𝑆𝑢,𝑎𝑣𝑔 × (1 + 0.2 × ) × 𝛼𝑛
𝐿

𝐴𝑝𝑙𝑎𝑡𝑒 = 𝐵 × 𝐿 = 10 × 4.5 = 45 𝑚2

𝑁𝑐 = 12.5 , assuming full suction is developed

𝑆𝑢,𝑎𝑣𝑔 = 𝜌 × 𝑧𝑎𝑣𝑔 = 1.5 × 20 = 30 𝑘𝑃𝑎

𝛼𝑛 = 0.7

𝛾𝑑 = 1.4, design safety factor

4.5
𝐹𝑛 = 45 × 12.5 × 30 × (1 + 0.2 × ) × 0.7 = 12875.6 𝑘𝑁
10

𝐹𝑛 13
𝐹𝑛 = 13 𝑀𝑁 ⟹ 𝐹𝑛,𝑑 = = = 9.2 𝑀𝑁
𝛾𝑑 1.4

MERIFIELD ET AL. (2001)


𝑄𝑢 = 𝐴𝑝𝑙𝑎𝑡𝑒 × 𝑁𝑐 × 𝑠𝑢,0 + 𝑊𝑎

𝐴𝑝𝑙𝑎𝑡𝑒 = 𝐵 × 𝐿 = 10 × 4.5 = 45 𝑚2

𝑊𝑎 = 500 𝑘𝑁

The ranges of theoretical solutions are:

𝐻 20
1< < 10 ⟺ 1 < < 10 ⟺ 1 < 4.44 < 10 , verifies
𝐵 4.5

𝜌𝐵 1.5×4.5
0.1 < < 1 ⟺ 0.1 < < 1 ⟺ 0.1 < 1.35 < 1 , does not verify
𝑠𝑢,0 5

However, the calculations will proceed.

→ 𝐷𝑒𝑡𝑒𝑟𝑚𝑖𝑛𝑒 𝑁𝑐 𝑎𝑐𝑐𝑜𝑟𝑑𝑖𝑛𝑔 𝑡𝑜 𝑠𝑢𝑔𝑔𝑒𝑠𝑡𝑒𝑑 𝑝𝑟𝑜𝑐𝑒𝑑𝑢𝑟𝑒:

127
Offshore Foundations: Technologies, Design and Application

1. 𝑠𝑢,0 = 5 𝑘𝑃𝑎
𝑘𝑁
𝛾 = 14
𝑚3
𝜌 = 1.5 𝑘𝑃𝑎/𝑚
2. 𝐻 = 20
𝐵 = 4.5
H/B=4.44
𝛾𝐻
=56
𝑠𝑢,0

3. From Figure 93
L/B=2.22
𝑁𝑐0 = 8
H/B=4.44
4. 𝑁𝑐∗ = 11.16 , for horizontal anchors
𝜌𝐵 2𝐻 1.5×4.5 2×20
5. 𝑁𝑐0𝜌 = 𝑁𝑐𝑜 [1 + 0.383 ( − 1)] = 8 [1 + 0.383 ( − 1)] = 40.63
𝑠𝑢,0 𝐵 5 4.5
𝛾𝐻 14×20
6. 𝑁𝑐 = 𝑁𝑐𝑜𝜌 + = 40.63 + = 96.63
𝑠𝑢,0 5

𝜌𝐻 1.5×20
7. 𝑁𝑐𝜌∗ = 𝑁𝑐∗ [1 + ] = 11.16 [1 + ] = 78.12
𝑠𝑢,0 5

8. 𝑁𝑐 > 𝑁𝑐𝜌∗ ⟹ 𝑁𝑐 = 78.12

𝑄𝑢 = 45 × 78.12 × 5 + 500 = 18077 𝑘𝑁 = 18 𝑀𝑁

𝑄𝑢 18
𝑄𝑢.𝑑 = = = 9 𝑀𝑁
𝛾𝑑 2.0

If partial safety factors are applied, instead of global safety factor:

𝑠𝑢,0 5
1. 𝑠𝑢,0𝑑 = = = 3.6 𝑘𝑃𝑎
𝛾𝑑 1.4

𝜌 1.5
𝜌𝑑 = = = 1.07 𝑘𝑃𝑎/𝑚
𝛾𝑑 1.4
1.07×4.5 2×20
2. 𝑁𝑐0𝜌 = 8 [1 + 0.383 ( − 1)] = 40.6
3.6 4.5
14×20
3. 𝑁𝑐 = 40.6 + = 118
3.6
1.07×20
4. 𝑁𝑐𝜌∗ = 11.16 [1 + ] = 78.1
3.6

𝑄𝑢,𝑑 = 45 × 78.1 × 3.6 + 500 = 13150 𝑘𝑁 = 13.15 𝑀𝑁

128
Offshore Foundations: Technologies, Design and Application

DNV-RP-E302 (2002)
𝑅𝑑 (𝑧𝑖 ) = 𝑁𝑐 ∙ 𝑠𝑐 ∙ 𝜂 ∙ 𝑠𝑢,𝑚𝑒𝑎𝑛 (𝑧𝑖 ) ∙ 𝐴𝑝𝑙𝑎𝑡𝑒 ∙ 𝑈𝑐𝑦 ∙ (1⁄𝛾𝑚 )

𝐴𝑝𝑙𝑎𝑡𝑒 = 45 𝑚2

𝑈𝑐𝑦 = 1.0 , 𝑐𝑜𝑛𝑠𝑒𝑟𝑣𝑎𝑡𝑖𝑣𝑒

𝐵 4.5
𝑠𝑐 = 1 + 0.2 × = 1 + 0.2 × = 1.09
𝐿 10

𝜂=0.75

𝑠𝑢,𝑚𝑒𝑎𝑛 (𝑧𝑖 ) = 20 × 1.5 = 30 𝑘𝑃𝑎

𝛾𝑚 = 1.40

→ 𝐷𝑒𝑡𝑒𝑟𝑚𝑖𝑛𝑒 𝑁𝑐 :

𝐻 20
= = 4.44 < 4.5, 𝑖𝑠 𝑐𝑜𝑛𝑠𝑖𝑑𝑒𝑟𝑒𝑑 𝑠ℎ𝑎𝑙𝑙𝑜𝑤 𝑝𝑙𝑎𝑡𝑒 𝑓𝑜𝑢𝑛𝑑𝑎𝑡𝑖𝑜𝑛
𝑊𝐹 4.5

𝑧𝑖
(𝑁𝑐 )𝑠ℎ𝑎𝑙𝑙𝑜𝑤 = 5.14(1 + 0.987 tan−1 ( )) = 11.98
𝑊𝐹

𝑅𝑑 (𝑧𝑖 ) = 11.98 ∙ 1.09 ∙ 0.75 ∙ 30 ∙ 45 ∙ 1.0 ∙ (1⁄1.40) = 9444 𝑘𝑁 = 9.5 𝑀𝑁

129

You might also like