an experimental study to assess the geotechnical ...€¦ · chapter 6: blake, a. p. and...

331
AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL PERFORMANCE OF DYNAMICALLY EMBEDDED PLATE ANCHORS by Anthony P. Blake B.Eng. (Hons) This thesis is presented for the degree of Doctor of Philosophy of The University of Western Australia Centre for Offshore Foundation Systems School of Civil, Environmental and Mining Engineering November 2015

Upload: others

Post on 30-Sep-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

AN EXPERIMENTAL STUDY TO ASSESS THE

GEOTECHNICAL PERFORMANCE OF DYNAMICALLY

EMBEDDED PLATE ANCHORS

by

Anthony P. Blake

B.Eng. (Hons)

This thesis is presented for the degree of

Doctor of Philosophy of

The University of Western Australia

Centre for Offshore Foundation Systems

School of Civil, Environmental and Mining Engineering

November 2015

Page 2: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed
Page 3: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

i

DECLARATION FOR THESES CONTAINING

PUBLISHED WORK AND/OR WORK PREPARED FOR

PUBLICATION

This thesis contains published work and work prepared for publication, which has

been co-authored. The bibliographical details of the work and where it appears in the

thesis are outlined below.

This thesis is presented as a series of academic papers in accordance with the

regulations of the University of Western Australia regarding higher degrees by research.

Chapter 1 and 9 comprise of the introduction and conclusions of the thesis respectively.

Chapters 2 and 3 contain the literature review. These chapters have been solely prepared

by the candidate. Chapters 4, 5, 6 and 7 are published journal papers. Chapter 8 has

been submitted to Géotechnique and was “under review” at the time of thesis

submission. Appendices I, II and III are papers published in peer reviewed conference

proceedings. The candidate is the first author on five of the eight publications arising

from this thesis and second author on the other three. Overall the candidate is

responsible for more than 80% of the content presented in this thesis. The publications

arising from this thesis are as follows:

Journal papers

Chapter 4: O’Loughlin, C. D., Blake, A. P., Richardson, M. D., Randolph, M. F.

and Gaudin, C. (2014). Installation and capacity of dynamically embedded plate

anchors as assessed through centrifuge tests. Ocean Engineering, Vol. 88, pp. 204–

213.

The centrifuge testing was carried out by the candidate along with preliminary data

processing and interpretation. The first author conducted further analyses of the data

Page 4: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

ii

and was responsible for an initial draft of the paper. All co-authors reviewed and

contributed to the final draft.

Chapter 5: Blake, A. P., O’Loughlin, C. D., Morton, J. M., O’Beirne, C., Gaudin,

C. and White D. J. (2016). In situ measurement of the dynamic penetration of free-

fall projectiles in soft soils using a low cost inertial measurement unit. Geotechnical

Testing Journal, forthcoming.

This paper presents field data gathered by the candidate (dynamically embedded

plate anchor data), and the third (instrumented free-fall sphere data) and fourth

(deep penetrating anchor data) authors. All data processing was conducted by the

candidate. The interpretation framework presented by in the paper was developed

and implemented by the candidate. The candidate prepared an initial draft of the

paper and all co-authors reviewed and contributed to the final draft.

Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically

embedded plate anchors as assessed through field tests. Canadian Geotechnical

Journal, Vol. 52, No. 9, pp. 1270–1282.

Field testing was conducted by the candidate along with data processing and

interpretation. The candidate prepared an initial draft of the paper and the co-author

all co-authors reviewed and contributed to the final draft.

Chapter 7: Blake, A. P., O’Loughlin, C. D. and Gaudin, C. (2014). Capacity of

dynamically embedded plate anchors as assessed through field tests. Canadian

Geotechnical Journal, Vol. 52, No. 1, pp. 87–95.

The field tests described in this paper and subsequent data processing and

interpretation were performed by the candidate. The candidate prepared an initial

draft of the paper and all co-authors reviewed and contributed to the final draft.

Page 5: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

iii

Chapter 8: O’Loughlin, C. D., Blake, A. P. and Gaudin, C. (2016). Towards a

simple design procedure for dynamically embedded plate anchors. Géotechnique,

under review.

The field tests described in this paper and subsequent data processing and

interpretation were performed by the candidate. The first author prepared an initial

draft of the paper. All co-authors reviewed and contributed to the final draft.

Conference papers

Appendix I: Blake, A.P., O’Loughlin, C. D. and Gaudin, C. (2010). Setup

following keying of plate anchors assessed through centrifuge tests in kaolin clay.

Proceedings of the 2nd International Symposium on Frontiers in Offshore

Geotechnics, Perth, Australia, 8–10 November 2010, pp. 705–710.

The centrifuge tests and subsequent data processing and interpretation were

performed by the candidate. The candidate prepared an initial draft of the paper and

all co-authors reviewed and contributed to the final draft.

Appendix II: Blake, A. P. and O’Loughlin, C. D. (2012). Field testing of a reduced

scale dynamically embedded plate anchor. Proceedings of the 7th International

Conference in Offshore Site Investigation and Geotechnics, London, United

Kingdom, 12–14 September 2012, pp. 621–628.

The field tests described in this paper and subsequent data processing and

interpretation were performed by the candidate. The candidate prepared an initial

draft of the paper and the co-author reviewed and contributed to the final draft.

Appendix III: O’Loughlin, C. D., Blake, A. P., Wang, D., Gaudin, C. and

Randolph, M. F. (2013). The dynamically embedded plate anchor: results from an

experimental and numerical study. Proceedings of the 32nd International

Conference in Ocean, Offshore and Artic Engineering, Nantes, France, 9–14 June

2013, OMAE2013-11571.

Page 6: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

iv

The centrifuge and field testing described in this paper were carried out by the

candidate. Numerical analysis was conducted by the third author. All co-authors

reviewed and contributed to the final draft.

Page 7: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

v

ABSTRACT

The dynamically embedded plate anchor (DEPLA) has been proposed as a cost effective

and technically efficient anchor for deepwater mooring applications. The DEPLA is

rocket or torpedo shaped anchor that comprises a removable central shaft and a set of

four flukes. The DEPLA penetrates to a target depth in the seabed by the kinetic energy

obtained through free-fall in water. After embedment the central shaft is retrieved

leaving the anchor flukes vertically embedded in the seabed. The flukes constitute the

load bearing element as a plate anchor. The DEPLA combines the installation

advantages of dynamically installed anchors (no external energy source or mechanical

operation required during installation) and the capacity advantages of vertical loaded

anchors (sustain significant horizontal and vertical load components).

Despite these advantages there are no current geotechnical performance data for the

DEPLA, as development of the anchor is in its infancy. This thesis has focused on

assessing the geotechnical performance of DEPLAs through an experimental study

involving a centrifuge testing program and an extensive field testing campaign. The

centrifuge tests provided early stage proof of concept for the DEPLA and motivation for

subsequent field testing. The field tests were carried out on reduced scale DEPLAs at

two sites: (i) Lough Erne, which is an inland lake located in County Fermanagh,

Northern Ireland and (ii) the Firth of Clyde which is located off the West coast of

Scotland. The reduced scale anchors were instrumented with a custom-design, low cost

inertial measured unit (IMU) for measurement of anchor motion during installation. The

field tests were designed to assess the behaviour of the DEPLA during free-fall in water

and dynamic embedment in soil, and quantify the loss in embedment during keying and

the subsequent plate anchor capacity. Analytical design tools for the prediction of

Page 8: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

vi

DEPLA embedment and capacity were verified, refined and calibrated using the

centrifuge and field data.

Back analysis of the IMU data showed that the fluid drag resistance acting on the

DEPLA during free-fall in water may be approximated using a constant mean drag

coefficient of 0.7. DEPLA impact velocities derived from the IMU data were in the

range 2.7 to 12.9 m/s corresponding to anchor release heights of 0.4 to 21.3 times the

anchor length. The DEPLA required a release height in the range 6.5 to 7 anchor lengths

to reach terminal velocity but then slowed slightly due to the increasing fluid drag

resistance afforded by the increasing length of mooring and follower recovery lines that

mobilised in the water. DEPLA tip embedment derived from the IMU data were in the

range 2.1 to 3.5 times the anchor length in Lough Erne and 1.4 to 2 times the anchor

length in the Firth of Clyde. These tip embedments were in good agreement with the

centrifuge data (1.6 to 2.8 times the anchor length). The shallower embedments

measured in the Firth of Clyde tests reflect the much higher strength gradient at this site.

The results indicated that anchor embedment increases with increasing impact velocity.

The appropriateness of an embedment prediction model for DEPLA based on strain rate

enhanced shear resistance and fluid mechanics drag resistance was demonstrated

through comparison with the IMU measurements. Qualitatively, the embedment depth

prediction model captures the dynamic response at both test sites adequately and is

capable of predicting the final embedments to within 10% of the measurements.

Back analysis of the IMU data indicated that a power function is capable of describing

the strain rate dependence of undrained shear strength at the very high strain rates

associated with the dynamic embedment process, although complexities associated with

initial embedment (up to one anchor length) were not captured by the model. However,

these shortcomings did not appear to unduly lessen the capability of the embedment

Page 9: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

vii

model to predict the measured velocity profiles and importantly the final anchor

embedment depth.

Video captured by a remotely operated vehicle in the Firth of Clyde showed an open

crater at the anchor drop sites which suggest that the hole caused by the passage of the

anchor remains open.

The keying response in the field was qualitatively consistent with the DEPLA

centrifuge data. However, despite the geometrical similarity between the field and

centrifuge DEPLAs, the loss in embedment of 1 to 1.8 times the plate diameter in

Lough Erne (could not be determined in the Firth of Clyde), was much higher than that

in the centrifuge tests (0.5 to 0.7 times the plate diameter). This was attributed to the

soil strength ratio which was much higher in Lough Erne than in the centrifuge tests.

Field test data indicated that the end of keying coincides with the peak anchor capacity

and the plate anchor bearing capacity factor decreases with reducing load inclination for

plate embedment ratios less than four. Back analysed peak bearing capacity factors from

the centrifuge and field tests were in the range 14.2 to 15.8 (centrifuge), 14.3 to 14.6

(Lough Erne) and 4.2 to 12.9 (Firth of Clyde). These bearing capacity factors were

compared with corresponding numerically derived values. This comparison indicated

the potential for soil tension to be lost at the underside of the anchor plate at shallow

embedment, where the plate embedment is less than 2.5 times the diameter, resulting in

a lower bearing capacity factor. At deep embedment, where the plate embedment is at

least 2.5 times the diameter, the experimental bearing capacity factors reached limiting

values in good agreement with that determined numerically.

A simple design procedure, based on a design chart presented in terms of the total

energy of the anchor at impact with the mudline for a range of strength gradients and

soil sensitivities, was shown to be useful approach for preliminary sizing of a DEPLA.

Page 10: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

viii

ACKNOWLEDGEMENTS

I am grateful for the financial support I received during my candidature, which included

a Scholarship for International Research Fees and a University International Stipend

from the University of Western Australia (UWA), a top-up scholarship from the Centre

for Offshore Foundation Systems (COFS), maintenance and fee grants from the Mayo

Vocational Education Committee, and a scholarship provided by the Society for

Underwater Technology through their Educational Support Fund. I am also grateful to

Enterprise Ireland for funding the centrifuge and field testing reported in this thesis,

through their Commercialisation Fund Technology Development Programme.

I would like to take this opportunity express my sincere gratitude to my supervisors

Associate Professor Conleth O’Loughlin and Professor Christophe Gaudin for their

encouragement, guidance, time and patience. It has been a great privilege to have

worked with both of you.

I appreciate the assistance provided by the COFS staff, particularly Bart Thompson,

Dave Jones, John Breen, Lisa Melvin, Monica Mackman, Keith Russell, Ivan Kenny

and Monika Mathyssek-Kilburn.

I acknowledge the professionalism, experience and skill of the individuals I have

worked with during field testing, especially John Taylor, Michael and John Patrick

McCaldin of Aghinver Boat Company, Toms Wadsworth of Hydrabotix, and Tom

Stevenson and Duncan Fraser of the University Marine Biological Station Millport.

I thank Cathal Colreavy for conducting site investigation and characterisation at both

test locations reported in this thesis.

Page 11: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

ix

I thank Assistant Professor Scott Draper for helping to enhance my understanding of

fluid mechanics and being a sounding board for my ideas relating to the topic.

I thank the people whose friendship and camaraderie I enjoyed over the course of my

studies such as Tom Burke, Cathal Colreavy, John Cunningham, John Heneghan,

Richard Hughes, Paul McGarry, Simon Leckie, Neil Marshall, Gary Jennings, Keith

Jennings, Colm O’Beirne, Shane O’Doherty, Joe Phillips, Lucile Quѐau, Raffaele

Ragni, Amin Rismanchian, Sam Stanier, Yusuke Suzuki, Pauline Truong and Padraig

Varley. Special thanks to Siobhan Durkan who provided me with encouragement and

support throughout the course of my studies.

I thank my secondary school physics and chemistry Austin Egan for the excellent

foundation for learning that he provided me with, which has helped me enormously

throughout my subsequent studies. I am fortunate to have been educated by Austin.

I thank my parents Patrick and Mary Blake for their encouragement, support and

patience throughout my education. During my time in Australia, I have missed my

family (Mom, Dad, Helen, Noel, Matthew and Jamie) dearly.

Page 12: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

x

TABLE OF CONTENTS

Chapter 1 Introduction .................................................................................................... 1

1.1 Anchoring systems ............................................................................................. 6

1.1.1 Anchor piles ................................................................................................ 6

1.1.2 Suction caissons .......................................................................................... 7

1.1.3 Drag embedment anchors............................................................................ 8

1.1.4 Vertical loaded anchors ............................................................................... 9

1.1.5 Dynamically installed anchors .................................................................. 11

1.1.6 Dynamically embedded plate anchor ........................................................ 13

1.2 Research objectives .......................................................................................... 16

1.3 Research methodology ..................................................................................... 17

1.4 Thesis structure................................................................................................. 18

Chapter 2 Dynamic embedment of projectiles into the seabed .................................... 23

2.1 Introduction ...................................................................................................... 23

2.2 Free-falling penetrometers ............................................................................... 23

2.2.1 Field studies .............................................................................................. 25

2.2.2 Laboratory studies ..................................................................................... 28

2.3 Dynamically installed anchors ......................................................................... 29

2.3.1 Torpedo anchor ......................................................................................... 29

2.3.2 Deep penetrating anchor ........................................................................... 30

2.3.3 OMNI-Max anchor ................................................................................... 32

2.3.4 Laboratory studies ..................................................................................... 33

2.4 Resistance forces acting on projectiles during dynamic embedment in soil .... 41

2.4.1 Fluid drag resistance ................................................................................. 42

2.4.2 Added mass ............................................................................................... 45

2.4.3 Shear strain rate effects in clay ................................................................. 46

2.5 Embedment depth prediction............................................................................ 54

2.5.1 Analytical modelling ................................................................................. 54

2.5.2 Numerical modelling................................................................................. 56

2.6 Summary .......................................................................................................... 59

Chapter 3 Capacity of plate anchors ............................................................................. 60

3.1 Introduction ...................................................................................................... 60

3.2 Keying .............................................................................................................. 60

3.2.1 Installation method .................................................................................... 61

3.2.2 Padeye eccentricity ratio ........................................................................... 61

Page 13: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

xi

3.2.3 Pullout angle ............................................................................................. 62

3.2.4 Soil properties ........................................................................................... 63

3.2.5 Anchor thickness ratio and submerged unit weight .................................. 66

3.2.6 Anchor aspect ratio ................................................................................... 68

3.2.7 Anchor roughness...................................................................................... 68

3.2.8 Anchor geometry coefficient..................................................................... 69

3.3 Capacity ............................................................................................................ 71

3.3.1 Failure mechanism .................................................................................... 71

3.3.2 Suction ...................................................................................................... 72

3.3.3 Previous studies ......................................................................................... 73

3.4 DEPLA capacity ............................................................................................... 90

3.5 Summary .......................................................................................................... 92

Chapter 4 Installation and capacity of dynamically embedded plate anchors as assessed

through centrifuge tests ................................................................................ 93

4.1 Abstract ............................................................................................................ 93

4.2 Introduction ...................................................................................................... 94

4.3 Centrifuge testing program ............................................................................... 96

4.3.1 Model anchors ........................................................................................... 97

4.3.2 Soil properties ........................................................................................... 99

4.3.3 Experimental arrangement and testing procedures ................................. 101

4.4 Results and discussion .................................................................................... 104

4.4.1 Impact velocity ........................................................................................ 105

4.4.2 Embedment depth.................................................................................... 106

4.4.3 Follower extraction ................................................................................. 107

4.4.4 Plate anchor keying and capacity ............................................................ 108

4.5 Conclusions .................................................................................................... 115

Chapter 5 In situ measurement of the dynamic penetration of free-fall projectiles in

soft soils using a low cost inertial measurement unit ................................. 118

5.1 Abstract .......................................................................................................... 118

5.2 Introduction .................................................................................................... 118

5.3 Free-falling projectiles ................................................................................... 122

5.3.1 Deep penetrating anchor ......................................................................... 122

5.3.2 Dynamically embedded plate Anchor ..................................................... 123

5.3.3 Instrumented free-falling sphere ............................................................. 124

5.3.4 Inertial measurement unit ........................................................................ 125

5.4 Interpretation of IMU measurements ............................................................. 127

Page 14: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

xii

5.4.1 Rotations ................................................................................................. 128

5.4.2 Accelerations ........................................................................................... 129

5.4.3 Velocities and distances .......................................................................... 130

5.4.4 Tilt angles ................................................................................................ 131

5.5 Test sites and soil properties .......................................................................... 132

5.6 Test procedure ................................................................................................ 134

5.7 Results and discussion .................................................................................... 137

5.7.1 Rotations ................................................................................................. 137

5.7.2 Accelerations ........................................................................................... 141

5.7.3 Velocity profiles ...................................................................................... 144

5.7.4 Verification of the IMU derived measurements ..................................... 145

5.7.5 Example application of projectile IMU data ........................................... 147

5.8 Conclusions .................................................................................................... 151

Chapter 6 Installation of dynamically embedded plate anchors as assessed through

field tests .................................................................................................... 153

6.1 Abstract .......................................................................................................... 153

6.2 Introduction .................................................................................................... 153

6.3 Experimental program .................................................................................... 155

6.3.1 Instrumented model anchors ................................................................... 155

6.3.2 Inertial Measurement Unit ...................................................................... 157

6.3.3 Test site and soil properties ..................................................................... 159

6.3.4 Testing setup and procedure ................................................................... 160

6.4 Results and discussion .................................................................................... 164

6.4.1 Accelerations ........................................................................................... 164

6.4.2 Free-fall in water ..................................................................................... 165

6.4.3 Soil penetration ....................................................................................... 170

6.5 Conclusions .................................................................................................... 181

Chapter 7 Capacity of dynamically embedded plate anchors as assessed through field

tests ............................................................................................................. 183

7.1 Abstract .......................................................................................................... 183

7.2 Introduction .................................................................................................... 183

7.3 Experimental program .................................................................................... 186

7.3.1 Model anchors and instrumentation ........................................................ 186

7.3.2 Test site and soil properties ..................................................................... 187

7.3.3 Testing setup and procedure ................................................................... 188

7.4 Results and discussion .................................................................................... 191

Page 15: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

xiii

7.4.1 Follower extraction ................................................................................. 191

7.4.2 Plate keying and capacity ........................................................................ 193

7.5 Conclusions .................................................................................................... 200

Chapter 8 Towards a simple design approach for dynamically embedded plate anchors

.................................................................................................................... 202

8.1 Abstract .......................................................................................................... 202

8.2 Introduction .................................................................................................... 203

8.3 Field Trials ..................................................................................................... 205

8.4 Instrumented reduced scale DEPLA .............................................................. 208

8.5 Test procedure ................................................................................................ 209

8.6 Results and discussion .................................................................................... 210

8.6.1 Impact velocity and embedment ............................................................. 211

8.6.2 Plate anchor capacity............................................................................... 214

8.6.3 Embedment depth prediction model ....................................................... 218

8.7 Design approach ............................................................................................. 228

8.8 Conclusions .................................................................................................... 231

Chapter 9 Conclusions ................................................................................................ 234

9.1 Summary ........................................................................................................ 234

9.2 Main findings ................................................................................................. 235

9.2.1 Inertial measurement unit ........................................................................ 235

9.2.2 Free-fall and dynamic embedment in soil ............................................... 236

9.2.3 Follower extraction ................................................................................. 238

9.2.4 Plate anchor keying response and capacity ............................................. 239

Appendix I – Setup following keying of plate anchors assessed through centrifuge tests

in kaolin clay .............................................................................................. 272

Appendix II – Field testing of a reduced scale dynamically embedded plate anchor ... 279

Appendix III – The dynamically embedded plate anchor: results from an experimental

and numerical study ................................................................................... 288

Page 16: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

xiv

LIST OF TABLES

Table 4.1. Centrifuge scaling laws (after Scholfield 1980, Taylor 1995) ....................... 96

Table 4.2. Model anchors ................................................................................................ 98

Table 4.3. Kaolin clay properties (after Stewart 1992) ................................................... 99

Table 4.4. Summary of test results ................................................................................ 117

Table 6.1. DEPLA anchors ........................................................................................... 157

Table 7.1. Summary of DEPLA field test results ......................................................... 192

Table 7.2 Comparison between the field and centrifuge tests of the non-dimensional

groups that govern keying behaviour ............................................................................ 197

Table 8.1. Summary of field test results at the Firth of Clyde ...................................... 211

Page 17: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

xv

LIST OF FIGURES

Figure 1.1. Worldwide progression of water depth capabilities for offshore drilling and

production (after FloaTEC 2012) ...................................................................................... 1

Figure 1.2. BW Pioneer FPSO (www.shipspotting.com) ................................................. 2

Figure 1.3. Global deepwater oil production capacity 2010 to 2016 (after International

Energy Agency 2011) ....................................................................................................... 2

Figure 1.4. Comparison of shallow and deepwater oil production in the Gulf of Mexico

(after Nixon et al. 2009) .................................................................................................... 3

Figure 1.5. Total projected oil and gas production in the Gulf of Mexico 2011 to 2020

(after PFC Energy 2011) ................................................................................................... 3

Figure 1.6. Offshore development systems (after FloaTEC 2012) ................................... 5

Figure 1.7. Catenary system and taut leg system mooring systems (after Vryhof

Anchors 2010) ................................................................................................................... 5

Figure 1.8. Prelude FPSO and shuttle tanker (after Offshore Magazine 2012) ................ 6

Figure 1.9. Anchor piles (after Eltaher et al. 2003) .......................................................... 7

Figure 1.10. Suction caissons (www.delmarus.com) ........................................................ 8

Figure 1.11. Drag embedment anchor (www.vryhof.com) ............................................... 9

Figure 1.12. Drag-in plate anchor (www.bruceanchor.co.uk) .......................................... 9

Figure 1.13. SEPLA (after Gaudin et al. 2006) .............................................................. 11

Figure 1.14. SEPLA concept: 1. Suction installation, 2. Caisson retrieval, 3. Anchor

keying, 4. Mobilised anchor (after Gaudin et al. 2006) .................................................. 11

Figure 1.15 Deep penetrating anchor installation procedure (after Lieng et al. 2000) ... 12

Figure 1.16. Torpedo anchor (after Araujo et al. 2004) .................................................. 13

Figure 1.17. Dynamically embedded plate anchor.......................................................... 14

Figure 1.18. DEPLA free-fall installation ....................................................................... 15

Figure 1.19. DEPLA keying, from left to right: dynamic free-fall installation, follower

removed and recovered on the vessel deck for reuse leaving DEPLA flukes embedded

in soil, DEPLA flukes keyed into position...................................................................... 16

Figure 2.1. FFPs developed for seabed characterisation (a) MIP (after Dayal et al.

1975), (b) MSP (after Colp et al. 1975), (c) Free-fall penetrometer (after Denesse et al.

1981), (d) XBP (after Stoll and Akal 1999), (e) FF-CPT (www.brooke-ocean.com), (f)

Page 18: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

xvi

CPT-Lance (after Strak et al. 2009b), (g) Nimrod (after Strak et al. 2009b) and (h)

LIRmeter (after Stephan et al. 2012) .............................................................................. 24

Figure 2.2. FFPs developed for naval mine countermeasure studies (a) XDP

(www.sonatech.com), (b) STING (after Abelev et al. 2009b), (c) ESP (after Mulhearn et

al. 1998), (d) BMMB (after Chow 2013), (e) AUSSI (after Mulhearn et al. 1999), (f)

FEP (after Chow 2013) and (g) PROBOS (after Stoll et al. 2007) ................................. 26

Figure 2.3. European standard penetrometer (after Freeman et al. 1984) ...................... 27

Figure 2.4. 98 t torpedo anchor for the Albacora Leste field in the Campos Basin (after

Brandão et al. 2006) ........................................................................................................ 30

Figure 2.5. Deep penetrating anchor (www.ngi.no) ....................................................... 31

Figure 2.6. Measured and predicted impact velocities and tip embedment depths for 1:4

scale model DPA test at Troll Field – grey band denotes range of measured values (after

Strum et al. 2011) ............................................................................................................ 32

Figure 2.7. OMNI-Max anchor (after Zimmerman et al. 2007) ..................................... 33

Figure 2.8. Dependence of anchor tip embedment depth on impact velocity and mass

(after O’Loughlin et al. 2013b) ....................................................................................... 34

Figure 2.9. Dependence of anchor tip embedment depth on aspect ratio (after

O’Loughlin et al. 2013b) ................................................................................................. 35

Figure 2.10. Dependence of anchor tip embedment depth on mass (after Richardson et

al. 2006) .......................................................................................................................... 35

Figure 2.11. Comparison of measured and predicted tip embedment depths for 1:30

reduced scale model torpedo anchors in kaolin clay (after Gilbert et al. 2008) ............. 36

Figure 2.12. Dependence of anchor tip embedment on impact velocity (after Cenac Π et

al. 2011) .......................................................................................................................... 37

Figure 2.13. Comparison of anchor embedment prediction model results and OMNI-

Max anchor field data (after Shelton et al. 2011) ........................................................... 37

Figure 2.14. 1:40 reduced scale model torpedo anchor (after Hasanloo et al. 2012) ..... 38

Figure 2.15. Comparison between the drag coefficients of torpedo anchors with

different (a) densities, (b) aspect ratio, (c) scale ratios, (d) fin sizes (after Hasanloo et al.

2012) ............................................................................................................................... 39

Figure 2.16. 1:200 reduced scale model OMNI-Max anchor (after Gaudin et al. 2013) 39

Figure 2.17. Embedment depths achieved by 1:200 reduced scale OMNI-Max anchors

from centrifuge tests compared with field data (after Gaudin et al. 2013) ..................... 40

Figure 2.18. 1:200 reduced scale torpedo anchors (after Hossain et al. 2014) ............... 41

Page 19: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

xvii

Figure 2.19. Embedment depths achieved by 1:200 reduced scale torpedo anchors from

centrifuge tests compared with field data (after Hossain et al. 2014) ............................. 41

Figure 2.20. Forces acting on a DEPLA during installation ........................................... 42

Figure 2.21. Dependence of drag coefficient on Reynolds number for (a) torpedo anchor

(after Hasanloo et al .2012) and (b) OMNI-Max anchor (after Cenac Π 2011) ............. 45

Figure 2.22. Illustration of viscous effects for normally consolidated and lightly

overconsolidated soils – representative of typical deepwater seabed deposits (after

Lehane et al. 2009) .......................................................................................................... 48

Figure 2.23. Back-analysed strain rate parameters for reduced scale mode DPAs from

centrifuge tests in kaolin clay (after O’Loughlin et al. 2013b) ....................................... 50

Figure 2.24. Back-analysed strain rate parameters for a reduced scale mode OMNI-Max

anchor from centrifuge tests in kaolin clay and calcareous silt (after Gaudin et al. 2013)

......................................................................................................................................... 50

Figure 3.1. Effect of padeye eccentricity ratio on keying behaviour .............................. 62

Figure 3.2. Influence of pullout angle on keying behaviour ........................................... 63

Figure 3.3. Effect of soil rigidity ratio, E/su, on keying behaviour (after Wang et al.

2011) ............................................................................................................................... 64

Figure 3.4. Influence of soil shear strength profile on keying behaviour (after Wang et

al. 2011)........................................................................................................................... 65

Figure 3.5. Effect of soil strength ratio on keying behaviour (a) after Song et al. (2009)

and (b) after Wang et al. (2011) ...................................................................................... 66

Figure 3.6. Effect of anchor thickness ratio on keying behaviour: (a) after Song et al.

(2009) and (b) after Wang et al. (2011) .......................................................................... 67

Figure 3.7. Influence of anchor submerged unit weight on keying behaviour (after Song

et al. 2009) ....................................................................................................................... 67

Figure 3.8. Combined effects of anchor roughness, thickness ratio and eccentricity ratio

on keying behaviour (after Wang et al. 2013b)............................................................... 68

Figure 3.9. Influence of anchor roughness on the keying behaviour (after Song et al.

2009) ............................................................................................................................... 69

Figure 3.10. Embedment depth loss expressed in terms of a non-dimensional anchor

coefficient: (a) after Song et al. (2009) and (b) after Gaudin et al. (2010) ..................... 70

Figure 3.11. Shallow and deep anchor behaviour ........................................................... 72

Figure 3.12. Effect of overburden stress ratio on bearing capacity factor ...................... 75

Figure 3.13. Effect of pullout angle on bearing capacity factor (after Song et al. 2005) 76

Page 20: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

xviii

Figure 3.14. Effect of pullout angle on bearing capacity factor (after Song et al. 2008a)

......................................................................................................................................... 77

Figure 3.15. Effect of soil strength gradient on bearing capacity factor (after Song et al.

2008a).............................................................................................................................. 77

Figure 3.16. Influence of normalised strength ratio on bearing capacity factor (after

Song et al. 2008a)............................................................................................................ 78

Figure 3.17. Bearing capacity factors and separation depth for continuous pullout of

rough circular plate anchors (after Song et al. 2008b) .................................................... 79

Figure 3.18. Influence of normalised strength ratio on separation depth (Song et al.

2008b) ............................................................................................................................. 79

Figure 3.19. Influence of loading rate on bearing capacity factor (after Wang et al.

2008) ............................................................................................................................... 80

Figure 3.20. Influence of padeye eccentricity ratio on bearing capacity factor (after Yu

et al. 2009)....................................................................................................................... 81

Figure 3.21. Bearing capacity factor for various pullout angles (after Yu et al. 2009) .. 81

Figure 3.22. Bearing capacity factors of plate anchors in uniform clay (after Wang et al.

2010) ............................................................................................................................... 82

Figure 3.23. Effect of soil rigidity ratio on the bearing capacity factor (after Wang et al.

2010) ............................................................................................................................... 82

Figure 3.24. Dependence of separation depth on soil strength ratio (after Wang et al.

2010) ............................................................................................................................... 83

Figure 3.25. Dependence of bearing capacity factor on anchor aspect ratio (after Wang

et al. 2010)....................................................................................................................... 83

Figure 3.26. Bearing capacity factors of strip anchors in (a) uniform clay and (b)

normally consolidated clay (after Yu et al. 2011) ........................................................... 84

Figure 3.27. Effect of soil non-homogeneity on the ultimate bearing capacity of strip

anchors (after Yu et al .2011) .......................................................................................... 85

Figure 3.28. Influence of soil roughness on bearing capacity (after Yu et al. 2011) ...... 85

Figure 3.29. Influence of anchor thickness ratio on the bearing capacity factor of deeply

embedded circular footings (Zhang et al. 2012) ............................................................. 87

Figure 3.30. Bearing capacity factors as a function overburden pressure ratio (after Chen

et al. 2013)....................................................................................................................... 87

Figure 3.31. Effect of soil non-homogeneity factor on bearing capacity factor (after Tho

et al. 2013)....................................................................................................................... 88

Figure 3.32. Effect of soil rigidity ratio on bearing capacity (after Chen et al. 2013) .... 89

Page 21: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

xix

Figure 3.33. Effect of loading rate on the bearing capacity factor of square anchors

(after Chen et al. 2013).................................................................................................... 89

Figure 3.34. Nc of DEPLAs as a function of embedment ratio ....................................... 91

Figure 3.35. Influence of anchor roughness on NC of DEPLAs ..................................... 91

Figure 3.36. Effect of fluke thickness on NC of DEPLAs ............................................... 91

Figure 3.37. Effect of plate inclination on Nc of DEPLAs ............................................. 92

Figure 3.38. Effect of soil seperation on Nc of DEPLAs ................................................ 92

Figure 4.1. 1:200 reduced scale DEPLAs ....................................................................... 98

Figure 4.2. Undrained shear strength profiles ............................................................... 100

Figure 4.3. Anchor installation guide ............................................................................ 102

Figure 4.4. Experimental arrangement and testing procedure: (a) prior to installation, (b)

DEPLA embedment, (c) follower retrieval and (d) anchor keying and pullout............ 104

Figure 4.5. Velocity profiles during centrifuge free-fall ............................................... 105

Figure 4.6. Determining anchor tip embedment during centrifuge tests....................... 106

Figure 4.7. Follower load-displacement response......................................................... 108

Figure 4.8. Load-displacement response during anchor keying and pullout ................ 109

Figure 4.9. Dimensionless load-displacement response during keying and pullout ..... 110

Figure 4.10. Loss in anchor embedment during keying ................................................ 113

Figure 4.11. Comparison of experimental and numerical bearing capacity factors...... 114

Figure 5.1. Schematic representation of the operational principle of: (a) MEMS

accelerometers and (b) MEMS gyroscopes................................................................... 120

Figure 5.2. Deep penetrating anchor ............................................................................. 123

Figure 5.3. Dynamically embedded plate anchor.......................................................... 124

Figure 5.4. Instrumented free-falling sphere ................................................................. 124

Figure 5.5. Inertial measurement unit ........................................................................... 126

Figure 5.6. Body frame of reference ............................................................................. 126

Figure 5.7. Body and inertial frames of reference ........................................................ 127

Figure 5.8. Test site locations ....................................................................................... 133

Figure 5.9. Undrained shear strength profiles: (a) Firth of Clyde and (b) Lough Erne 134

Figure 5.10. RV Aora .................................................................................................... 135

Figure 5.11. Self-propelled barge.................................................................................. 135

Figure 5.12. DEPLA field test procedure ..................................................................... 136

Figure 5.13. Image capture from ROV camera showing the follower retrieval line at the

seabed ............................................................................................................................ 137

Page 22: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

xx

Figure 5.14. Projectile rotations during free-fall through water and soil penetration: (a)

DEPLA, (b) IFFS and (c) DPA ..................................................................................... 140

Figure 5.15. Projectile accelerations during free-fall through water and soil penetration:

(a) DEPLA, (b) IFFS and (c) DPA ............................................................................... 143

Figure 5.16. Projectile velocity profiles corresponding to free-fall through water and soil

penetration: (a) DEPLA, (b) IFFS and (c) DPA ........................................................... 146

Figure 5.17. Comparison of IMU derived displacement measurements with those

obtained using a draw wire sensor ................................................................................ 147

Figure 5.18. DEPLA velocity profile derived from the IMU data measured at the Firth

of Clyde test site and corresponding theoretical profile ............................................... 151

Figure 6.1. DEPLA dimensions .................................................................................... 156

Figure 6.2. Typical undrained shear strength profiles .................................................. 160

Figure 6.3. Self-propelled barge ................................................................................... 161

Figure 6.4. (a), (b) and (c). DEPLA field test procedure .............................................. 162

Figure 6.5. 1:7.2 scale DEPLA being lowered to drop height by the barge crane ........ 163

Figure 6.6. Image capture from under-water camera showing the follower retrieval line

at lakebed ...................................................................................................................... 163

Figure 6.7. Accelerations during free-fall through water and soil penetration (1:12 scale)

....................................................................................................................................... 165

Figure 6.8. Dependence of DEPLA drag coefficient on Reynolds number .................. 169

Figure 6.9. Examples of measured and theoretical velocity profiles in water .............. 170

Figure 6.10. Forces acting on DEPLA and velocity profiles during soil penetration (a)

1:12 scale, (b) 1:7.2 scale, (c) 1:4.5 scale ..................................................................... 180

Figure 6.11. Dependence of tip embedment on impact velocity .................................. 181

Figure 7.1. DEPLA field test procedure ....................................................................... 190

Figure 7.2. Typical load-displacement response during follower extraction (1:12 scale

DEPLA) ........................................................................................................................ 192

Figure 7.3. Typical load-displacement responses of during keying and pullout .......... 194

Figure 7.4. Loss in anchor embedment during keying .................................................. 195

Figure 7.5. Influence of pullout angle on bearing capacity factor (1:12 scale DEPLA)

....................................................................................................................................... 198

Figure 7.6. Measured and predicted capacity for each anchor scale ............................. 200

Figure 8.1. Location of the field trials .......................................................................... 206

Figure 8.2. Index properties at the anchor trial locations in Firth of Clyde .................. 206

Page 23: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

xxi

Figure 8.3. Undrained shear strength profile at the anchor trial locations in Firth of

Clyde ............................................................................................................................. 207

Figure 8.4. Degradation of undrained shear strength with cycle number during piezoball

cyclic remoulding tests .................................................................................................. 207

Figure 8.5. Assembling the DEPLA on the deck of the RV Aora in preparation for

installation ..................................................................................................................... 210

Figure 8.6. An ROV video still showing the DEPLA installation site. These

observations were primarily intended to ensure no gross error in the embedment depth

established from the anchor instrumentation, but also provided evidence that the hole

created by the passage of the anchor remained open. ................................................... 210

Figure 8.7. Embedment trends for different scale DEPLAs in soils with different

strength gradients .......................................................................................................... 213

Figure 8.8. Harmonising embedment trends through the use of total energy ............... 213

Figure 8.9. Typical load-displacement responses during keying and pullout ............... 216

Figure 8.10. Experimental and numerical anchor capacity factors ............................... 217

Figure 8.11. Forces acting on the DEPLA during dynamic embedment in soil ........... 221

Figure 8.12. Determining drag coefficients Cd,al and Cd,l by optimising the fit between

measured and theoretical velocity profiles in water ...................................................... 223

Figure 8.13. Evaluation of the power law in quantifying strain rate dependency for

dynamic installation problems: (a) and (c) Test 12, (b) and (d) Test 2 ......................... 227

Figure 8.14. Measured and predicted velocity profiles ................................................. 227

Figure 8.15. Measured and predicted DEPLA follower tip embedments ..................... 228

Figure 8.16. DEPLA design chart for a range of seabed strength gradients and interface

friction ratios ................................................................................................................. 230

Figure 8.17. Flow chart for DEPLA design approach .................................................. 231

Page 24: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

xxii

NOTATION

The notation used in this thesis has been adopted with the aim of maintaining

consistency with previously published work. However, there may be instances where

different notation has been used for a specific parameter.

Roman

a resultant linear acceleration

ax linear acceleration coincident with the inertial frame x-axis

ay linear acceleration coincident with the inertial frame y-axis

az linear acceleration coincident with the inertial frame z-axis

A reference area

Abx acceleration measurement coincident with the body frame x-axis

Aby acceleration measurement coincident with the body frame y-axis

Abz acceleration measurement coincident with the body frame z-axis

Ap frontal area

As surface area

Ax acceleration measurement coincident with the inertial frame x-axis

Ay acceleration measurement coincident with the inertial frame y-axis

Az acceleration measurement coincident with the inertial frame z-axis

B plate width

ch horizontal coefficient of consolidation

cv vertical coefficient of consolidation

Ca added mass coefficient

Page 25: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

xxiii

Cd drag coefficient

Cd,a anchor drag coefficient

Cd,l mooring line and follower recovery line drag coefficient

Cm inertia coefficient

d diameter

deff effective diameter

dv padeye displacement

dt change in time

dT-bar T-bar diameter

D plate diameter

Dball ball diameter

Dcone cone diameter

Df follower diameter

Ds sleeve diameter

ecs void ratio on critical state line

E Young’s modulus

Etotal total energy

Fb soil buoyancy force

Fbear bearing resistance force

Fd fluid drag resistance force

Fd,a fluid drag resistance force acting on the anchor

Fd,l fluid drag resistance force acting on the mooring line and follower

recovery line

Ffrict frictional resistance force

Page 26: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

xxiv

Fi inertia resistance force

Fmax maximum force

Fn normal force

Fv,net net force

Fvf follower capacity

Fvp plate capacity

g gravitational acceleration

G shear modulus

Gs specific gravity

Hs sleeve height

k undrained shear strength gradient

K soil viscosity property (consistency parameter)

L length

Lf follower length

Ltip tip length

m mass

m' effective mass

ma added mass

mf follower mass

mp plate mass

M0 initial moment corresponding to zero net vertical load

n rate effect parameter

ncent centrifuge scale factor

n1 rate effect parameter

Page 27: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

xxv

Nball piezoball bearing capacity factor

Nc bearing capacity factor

Ncb bearing capacity factor under breakaway conditions and considering soil

self-weight

Nc,max maximum bearing capacity factor

Nc0 bearing capacity factor under breakaway conditions in weightless soil

Nkt piezocone bearing capacity factor

NP bearing capacity factor for a force normal to a pipeline

NT-bar T-bar bearing capacity factor

Pstag stagnation pressure

Pstat static pressure

p' mean effective strength

qin intact penetrometer tip resistance

qrem fully remoulded penetrometer tip resistance

Rbi direction cosine matrix (body frame to inertial frame)

Re Reynolds number of a Newtonian fluid

Renon-Newtonian Reynolds number of a non-Newtonian fluid

Rf strain rate function

Rx roll matrix

Ry pitch matrix

Rz yaw matrix

St soil sensitivity

s distance travelled in the direction of motion

s0 initial distance travelled in the direction of motion

Page 28: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

xxvi

sbz vertical distance travelled coincident with the body frame

su undrained shear strength

su,min minimum undrained shear strength at zero strain rate

su,p undrained shear strength at an embedment corresponding to peak plate

capacity (plate mid-height)

su,op operative undrained shear strength

sur remoulded undrained shear strength

su0 undrained shear strength at the initial anchor embedment depth

sz vertical distance travelled coincident with the inertial frame

sz0 initial vertical distance travelled coincident with the inertial frame

t time

T non-dimensional time factor

Tbi angular velocity transformation matrix (body frame to inertial frame)

v velocity

vbz velocity coincident with the body frame z-axis

vi impact velocity

vx velocity coincident with the inertial frame x-axis

vx0 initial velocity coincident with the inertial frame x-axis

vy velocity coincident with the inertial frame y-axis

vy0 initial velocity coincident with the inertial frame y-axis

vz velocity coincident with the inertial frame z-axis

vz0 initial velocity coincident with the inertial frame z-axis

vtheo theoretical velocity

V non-dimensional velocity

Page 29: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

xxvii

Va volume of anchor

Vs volume of soil

w water content

W dry weight

Ws submerged weight

x inertial frame x-axis

xb body frame x-axis

y inertial frame y-axis

yb body frame y-axis

z embedment depth or inertial frame z-axis

zb body frame z-axis

zchain chain length

zcr critical embedment depth

ze anchor tip embedment

ze,plate final plate embedment depth

ze,plate,0 initial plate embedment depth

zLC length of chain

zm embedment depth (model scale)

zp embedment depth (prototype scale)

zs separation depth

zslack length of chain slack

Page 30: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

xxviii

Greek

α adhesion factor

β strain rate parameter (power law)

γ unit weight of soil

γa unit weight of anchor

γw unit weight of water

γ' effective unit weight of soil

γ'a effective unit weight of anchor

γ̇ strain rate

γ̇ref

reference strain rate

η soil viscosity parameter

θ pitch angle coincident with the inertial frame

θ0 initial pitch angle coincident with the inertial frame

θb pitch angle coincident with the body frame

θb0 initial pitch angle coincident with the body frame

θacc pitch angle coincident with the inertial frame (derived from

accelerometer measurement)

θp plate inclination

θpull pullout angle (load inclination to the horizontal at the anchor padeye)

κ slope of swelling line

Δze difference between measured and predicted embedment depth

Δze,follower follower embedment depth loss

Δze,plate plate embedment depth loss

Δze,max maximum plate embedment depth loss

Page 31: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

xxix

λ strain rate parameter (semi-logarithmic law)

ν dynamic viscosity

μ resultant tilt angle (Chapter 5) or dynamic viscosity (Chapter 6)

μk kinetic friction coefficient

π ratio of a circle’s circumference to its diameter

ρ density

ρs soil density

ρstag stagnation pressure

ρstat static pressure

ρw water density

ς slope of normal consolidation line

σ'h horizontal effective stress

σ'v vertical effective stress

τ shear stress

τy shear yield stress

ψ yaw angle coincident with the inertial frame

ψ0 initial yaw angle coincident with the inertial frame

ψb yaw angle coincident with the body frame

ψb0 initial yaw angle coincident with the body frame

ωbx rotation rate about the body frame x-axis

ωby rotation rate about the body frame y-axis

ωbz rotation rate about the body frame z-axis

ωx rotation rate about the inertial frame x-axis

ωy rotation rate about the inertial frame y-axis

Page 32: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

xxx

ωz rotation rate about the inertial frame z-axis

ϕ roll angle coincident with the inertial frame

ϕ0 initial roll angle coincident with the inertial frame

ϕb roll angle coincident with the body frame

ϕb0 initial roll angle coincident with the body frame

ϕacc roll angle coincident with the inertial frame (derived from accelerometer

measurement)

ϕ' angle of internal friction

Abbreviations

ALE Arbitrary Lagrangian-Eulerian

ARM Acorn RISC Machines

AUSSI Australian Underwater Sediment Strength Instrument

AVTM Angular Velocity Transformation Matrix

BMMP Burying Mock Mine Body

CEL Coupled Eulerian-Lagrangian

CFD Computational Fluid Dynamics

CPT Cone Penetrometer Test

DCM Directional Cosine Matrix

DEPLA Dyanmically Embedded PLate Anchor

DoF Degrees of Freedom

DPA Deep Penetrating Anchor

ESP European Standard Penetrometer

FEP FAU Experimental Penetrometer

Page 33: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

xxxi

FFCPT Free-Falling Cone Penetrometer Test

FFP Free-Falling Penetrometer

FPSO Floating, Production, Storage and Offloading unit

FPU Floating Production Unit

GB Giga Byte

GME Great Meteor East

GOM Gulf of Mexico

IFFS Instrumented Free-Falling Sphere

IMU Inertial Measurement Unit

LDFE Large Deformation Finite Element

LIR Lance Insertion Rod

LL Liquid Limit

MEMS Micro Electro Mechanical System

MIP Marine Impact Penetrometer

MODU Mobile Offshore Drilling Unit

MSP Marine Sediment Penetrometer

NAP Nares Abyssal Plain

OCR Over Consolidation Ratio

PERP Photo Emitter Receiver Pair

PL Plastic Limit

PROBOS Proboscis

PVC Poly Vinyl Chloride

RV Research Vessel

STING Seabed Terminal Impact Naval Gauge

Page 34: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

xxxii

UWA University of Western Australia

VLA Vertical Loaded Anchor

XBP eXpendable Bottom Penetrometer

XDP eXpendale Doppler Penetrometer

Page 35: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

1

CHAPTER 1 INTRODUCTION

In the context of offshore hydrocarbon exploration and development, the advancement

of technology has changed the definition of deepwater over time. At present, waters

depths less than 300 m are considered ‘shallow’, depths ranging 300 to 1500 m are

regarded as ‘deep’ and depths varying from 1500 to 3000 m are classified as ‘ultra-

deep’ (Nixon et al. 2009). In 1947 the world’s first offshore platform out of sight of

land, Superior, was installed in the Gulf of Mexico, 30 km off the coast of Louisiana in

6 m of water (Aubeny et al. 2001). Since then the offshore oil and gas industry have

installed platforms in ever increasing water depths (Figure 1.1). During 2012 the

Pioneer FPSO was installed in the Gulf of Mexico 250 km off the coast of Louisiana in

a world record water depth of 2500 m (Figure 1.2) (Pipeline and Gas Journal 2012).

Figure 1.1. Worldwide progression of water depth capabilities for offshore drilling

and production (after FloaTEC 2012)

Global oil demand was 89.8 million barrels per day (mb/d) in 2012, with medium-term

forecasts projecting an increase of 7.6 % to 96.7 mb/d by 2018 (International Energy

Agency 2013). Long-term forecasts predict a 12.3 % increase in world oil demand to

99.7 mb/d by 2035. Over the same period global natural gas demand is expected to rise

Page 36: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

2

by 47.1 % to 5.0 trillion cubic metres (International Energy Agency 2012). Oil

production from offshore areas represented 30 % of global capacity in 2011 and is

expected to remain the same through to 2016. However, global oil production from

deepwater is predicted to increase from 22 % (6 % of global capacity) to 29 % (9 % of

global capacity) by 2016 (Figure 1.3) (International Energy Agency 2011).

Figure 1.2. BW Pioneer FPSO (www.shipspotting.com)

Figure 1.3. Global deepwater oil production capacity 2010 to 2016 (after

International Energy Agency 2011)

Over the past 15 years oil and gas production from shallow water regions has been in

decline. In the Gulf of Mexico deepwater oil production now exceeds shallow water

production (Figure 1.4) (Nixon et al. 2009). It is estimated that by 2020 over 50 % of

the Gulf of Mexico’s oil and gas production will come from ultra-deepwater, which can

Page 37: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

3

produce 45 % of the total projected output over the decade from 2011 to 2020 (Figure

1.5) (PFC Energy 2011).

Figure 1.4. Comparison of shallow and deepwater oil production in the Gulf of

Mexico (after Nixon et al. 2009)

Figure 1.5. Total projected oil and gas production in the Gulf of Mexico 2011 to

2020 (after PFC Energy 2011)

Increasing global demand and declining shallow water production requires the

exploration and development of deepwater and ultra-deepwater oil and gas reserves in

order to ensure future supply. There have been major deepwater discoveries in various

regions of the world including Gulf of Mexico, West Africa, Asia Pacific regions and

Page 38: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

4

Brazil, such as the Tupi oil field located in the Santos Basin off the Coast of Brazil in

water depths ranging from 1500 to 3000 m. This field was discovered in 2007 with

estimated recoverable reserves in the range of 5 to 8 billion barrels of oil (Offshore

Magazine 2008). Australia’s deepwater exploration and development activities include

the Enfield and Stybarrow oil fields located in water depths of 550 and 825 m

respectively, and several substantial gas fields in waters depths of up to 1500 m such as

Dionysus (1100 m), Geryon (1232 m), Io (1350 m) and Scarborough (1500 m)

(Department of Industry and Resources 2008).

In shallow water oil and gas development systems are typically bottom-founded

structures (e.g. fixed platforms and compliant towers) that are not viable in deepwater.

Instead, floating vessels are utilised such as Tension Leg Platforms (TLPs), Semi-

submersible floating and production units (Semi-FPUs), Floating Production, Storage

and Offloading units (FPSOs) and SPARs (Figure 1.6). Floating vessels are kept on

station with mooring lines anchored to the seabed. The oil and gas industry has

traditionally used a catenary mooring system in shallow water (Figure 1.7). However, in

deepwater operations, the weight and length of the catenary mooring line become

limiting factors in the design of the platform (Vryhof Anchors 2010). Therefore it is

becoming increasingly common for deepwater platforms to be moored using taut-leg

(Figure 1.7) and vertical mooring systems. These systems are designed to have high

angles of inclination with the mud-line (up to 90° in the case of vertical mooring

systems) and require anchors that can sustain both horizontal and vertical load

components.

Page 39: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

5

Figure 1.6. Offshore development systems (after FloaTEC 2012)

Figure 1.7. Catenary system and taut leg system mooring systems (after Vryhof

Anchors 2010)

Recent advancements in technology together with increasing demand for natural gas

have made Floating Liquefied Natural Gas (FLNG) developments feasible. FLNG

developments require significantly larger floating vessels than are currently available.

For example the Prelude FPSO will be the world’s first FLNG plant and largest floating

vessel with a length of 488 m and deadweight of 600,000 t (50 % greater than the

largest existing FPSO the Pozflar). Prelude will be permanently moored for up to 25

Page 40: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

6

years at the Browse Basin on the North West Shelf of Australia 200 km offshore in 250

m of water (Figure 1.8) (Offshore Magazine 2012). FLNG developments in deepwater

have also been proposed such as at the Carnarvon Basin on the North West Shelf of

Australia 220 km offshore in 900 m of water (Offshore Magazine 2013). Permanently

moored vessels of Prelude’s size and weight will significantly increase the loads that the

anchoring system must sustain.

Figure 1.8. Prelude FPSO and shuttle tanker (after Offshore Magazine 2012)

1.1 Anchoring systems

1.1.1 Anchor piles

Anchor piles typically comprise of a hollow steel tube with a mooring line attached at

the padeye (Figure 1.9). Installation generally involves either vibration or hammering.

Anchor piles can resist vertical and horizontal loads making them suitable for catenary,

taut-leg and vertical mooring systems. Installation costs for anchor pile are extremely

high and significantly increase as crane barges and pile driving equipment are required.

Anchor piles have been installed in water depths exceeding 1500 m (Offshore Magazine

2005).

Page 41: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

7

Figure 1.9. Anchor piles (after Eltaher et al. 2003)

1.1.2 Suction caissons

Suction caissons comprise of a large diameter steel cylindrical shell with a cover plate at

the top, an open bottom and a mooring line attached at the padeye (Figure 1.10). During

installation the caisson is lowered to the seabed, where it is allowed to penetrate under

self-weight. Water is then pumped from the interior of the caisson to create a pressure

differential which drives the system further into the soil. Suction caissons can resist

Page 42: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

8

vertical and horizontal loads enabling them to be employed for catenary, taut-leg and

vertical mooring systems. It can be difficult for suction caissons to penetrate hard layers

of soil in the seabed. The soil plug within the anchor may fail during installation in sand

overlain by clay and thin-walled caissons may buckle due to the relatively high suction

pressures required to penetrate the sand. Installation costs are relatively high as suction

caissons require considerable deck space during transport and heavy lift vessels for

installation. These costs increase significantly with water depth (Richardson 2008).

Suction caissons have been installed in water depths exceeding 2621 m (Offshore

Magazine 2010).

Figure 1.10. Suction caissons (www.delmarus.com)

1.1.3 Drag embedment anchors

Drag embedment anchors comprise of a bearing plate (referred to as a fluke) rigidly

connected to a shank (Figure 1.11). These anchors are designed to self-embed when

dragged along the seabed by a wire rope or chain and are easily recovered making then

suitable for short to medium term mooring applications. Although drag anchors achieve

high capacity to weight ratios there is a possibility of interference with existing subsea

infrastructure during installation and it is difficult to predict the embedment (Aubeny et

al. 2001). Drag embedment anchors are only suitable for catenary mooring systems as

they cannot resist vertical load.

Page 43: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

9

Figure 1.11. Drag embedment anchor (www.vryhof.com)

1.1.4 Vertical loaded anchors

Vertical loaded anchors (VLAs) include drag-in plate anchors and direct embedment

anchors. Similar to drag embedment anchors, drag-in plate anchors comprise of a fluke

and shank, and self-embed when dragged along the seabed (Figure 1.12). The shank is

triggered when the fluke reaches the design embedment depth enabling the anchor to

rotate to an orientation that is perpendicular to the direction of load (referred to as

keying). As with drag embedment there is a possibility of interference with existing

subsea infrastructure during installation and the final anchor embedment is difficult to

predict (Ehlers et al. 2004) and there is also uncertainty regarding the keying processes

(Cassidy et al. 2012).

Figure 1.12. Drag-in plate anchor (www.bruceanchor.co.uk)

Page 44: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

10

Direct embedment anchors comprise of a plate anchor connected to a mooring line

installed at the end of a follower by free-fall, driving, vibrating or suction. The suction

embedded plate anchor (SEPLA) (Figure 1.13) is a direct embedment anchor which

uses a suction caisson follower to install a vertically orientated plate anchor that is

slotted into its base (Dove et al. 1998). The suction caisson enables the plate anchor to

be installed to a known penetration depth. The caisson is lowered to the seabed, where it

is allowed to penetrate under self-weight. Water is then pumped from the interior of the

caisson to allow the plate anchor to reach the design embedment depth. The plate

anchor mooring line is then disengaged from the caisson and the pump flow is reversed

with water being forced back into the caisson causing it to move upwards while leaving

the plate anchor in place at the design embedment depth. The follower is recovered to

the deck of an anchor handling vessel (AHV) for reuse in future installations. When

sufficient load develops in the mooring line the plate anchor is keyed to an orientation

that is approximately normal to the direction of loading. Sufficient load may be applied

by the bollard pull of an AHV or may develop during the life cycle of the mooring

system if the necessary environmental loading conditions occur. The installation and

keying process of the SEPLA are summarised in Figure 1.14. SEPLAs are commonly

used for the mooring of mobile offshore drilling units (MODUs) in water depths

ranging from 381 to 2195 m (Brown et al. 2010) and has recently been considered as

permanent mooring for developments in West Africa (Wong et al. 2012)

The capacity to weight ratio of VLAs is much higher than that of anchor piles and

suction caissons. VLAs can sustain significant components of horizontal and vertical

loads allowing them to be used in catenary and taut-leg mooring systems. Uncertainty

exists regarding the embedment depth loss due to the keying process and the subsequent

capacity of the plate anchor.

Page 45: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

11

Figure 1.13. SEPLA (after Gaudin et al. 2006)

Figure 1.14. SEPLA concept: 1. Suction installation, 2. Caisson retrieval, 3. Anchor

keying, 4. Mobilised anchor (after Gaudin et al. 2006)

1.1.5 Dynamically installed anchors

Dynamically installed anchors are designed to be released from a designated drop height

above the seabed and penetrate to a target depth in the seabed by the kinetic energy

obtained through free-fall (Figure 1.15). Resistance to loading is predominantly

provided by the skin friction developed from the anchor-soil interface. Dynamically

installed anchors can withstand both horizontal and vertical load components enabling

their use in both catenary and taut-leg mooring systems. Three types of dynamically

installed anchor have been developed: the torpedo anchor (also referred to as torpedo

pile) (Figure 1.16), the Deep Penetrating Anchor, DPA, and the OMNI-Max anchor (see

Page 46: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

12

Section 2.3). Installation costs are relatively low compared to conventional deepwater

anchoring systems such as anchor piles and VLAs. Unlike conventional anchoring

systems, the costs of dynamically installed anchors are independent of water depth as no

external energy source or mechanical intervention is required during installation

(Richardson et al. 2005). Fabrication and handling costs are also relatively low due to

the simple design of the anchor. Their capacity is less dependent on soil shear strength

than conventional anchoring systems as greater penetration depths are achieved in soils

with relatively low shear strengths and vice versa. However this type of anchor may not

be suitable for use in soils with relatively high shear strengths where adequate

embedment depths are difficult to achieve. Uncertainty exists in relation to predicting

the final embedment depth and consequently the capacity of dynamically installed

anchors.

Figure 1.15 Deep penetrating anchor installation procedure (after Lieng et al.

2000)

Page 47: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

13

Figure 1.16. Torpedo anchor (after Araujo et al. 2004)

1.1.6 Dynamically embedded plate anchor

The dynamically embedded plate anchor (DEPLA) (Figure 1.17) is a hybrid anchor

concept that combines the installation advantages of dynamically installed anchors and

the capacity advantages of VLAs (O’Loughlin et al. 2013a). The DEPLA was

conceptualised at UWA’s Centre for Offshore Foundation Systems in 2004. Further

work was undertaken to validate and refine the concept during 2008 to 2014, as detailed

in this thesis. In 2013, the intellectual property rights were assigned under contract to

Vryhof Anchors of the Netherlands.

The DEPLA is rocket or dart shaped with a removable central shaft (follower) and a set

of four flukes attached to a sleeve (plate anchor). A stop cap at the upper end of the

follower prevents the follower from falling through the DEPLA sleeve and a shear pin

connects the flukes to the follower. After release from a designated height above the

seabed (typically <100 metres), the DEPLA penetrates to a target depth in the seabed by

the kinetic energy obtained through free-fall and the self-weight of the anchor. After

embedment the follower is retrieved for the next installation leaving the flukes and

sleeve vertically embedded in the seabed. These embedded anchor flukes and sleeve

constitute the load bearing element as a plate anchor. The plate anchor remains vertical

in the seabed until sufficient load develops in the mooring line to cause the plate anchor

Page 48: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

14

to rotate or ‘key’ to an orientation that is approximately perpendicular to the direction of

loading maximising the bearing capacity of the anchor. Sufficient load may be applied

by the bollard pull of an AHV or may develop during the life cycle of the mooring

system if the necessary environmental loading conditions occur. The installation and

keying process is summarised in Figure 1.18 and Figure 1.19 respectively.

The advantages of the DEPLA are:

Installation durations and hence costs are significantly reduced as no mechanical

intervention or external energy source is required.

Installation durations and costs are independent of water depth.

Plate anchor can withstand both horizontal and vertical loads enabling use in

catenary, taut-leg and vertical mooring systems.

Fabrication is simple and economical.

Reuse of follower ˗ full anchor spread transported to site on a single vessel.

Robust and compact design makes handling and installation simple and economical

using one anchor AHV and requiring no remotely operated vehicle (ROV).

Figure 1.17. Dynamically embedded plate anchor

Page 49: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

15

Figure 1.18. DEPLA free-fall installation

Page 50: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

16

Figure 1.19. DEPLA keying, from left to right: dynamic free-fall installation,

follower removed and recovered on the vessel deck for reuse leaving DEPLA

flukes embedded in soil, DEPLA flukes keyed into position

1.2 Research objectives

Installation costs of conventional anchoring systems such as driven piles, suction

caissons, drag embedment anchors and vertical load anchors are highly dependent on

water depth. It is evident that mooring systems for deepwater applications and FLNG

vessels require a reliable anchoring system that can withstand major components of

horizontal and vertical loads, while maintaining installation costs at a reasonable level.

Hence, the development of an anchoring system that can meet these requirements is

necessary.

The DEPLA offers the installation advantages of dynamically installed anchors (no

external energy source or mechanical operation required during installation) and the

capacity advantages of VLAs (sustain significant horizontal and vertical load

Page 51: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

17

components). Despite these advantages there are no current geotechnical performance

data for the DEPLA. Such data would evidently stimulate confidence in the potential of

DEPLAs as a cost effective and technically efficient anchor for deepwater applications

and FLNG vessels. Therefore, there is a clear requirement for an experimental study to

investigate the geotechnical behaviour of these anchors and provide ‘proof of concept’.

The overall objective of this thesis is to enhance the understanding of the geotechnical

behaviour of DEPLA. In order to achieve this overall objective five detailed objectives

have been identified:

Proof of concept - demonstrate feasibility through anchor deployment and retrieval.

Installation - quantifying anchor embedment for a given anchor drop height,

geometry and seabed strength profile.

Keying - quantifying the embedment depth loss of the plate anchor during keying.

Capacity - determination of anchor capacity for a given seabed strength profile.

Design tools - verification and calibration of design tools for the prediction of

anchor embedment, keying response and anchor capacity.

1.3 Research methodology

Reduced scale model DEPLAs will be modelled in UWA’s beam centrifuge.

Geotechnical field testing in an aquatic environment is generally a challenging and

expensive undertaking. Centrifuge testing offers a more convenient and economical

alternative to field testing. Testing parameters such as anchor drop height (and hence

impact velocity), anchor geometry and mass, and stress history of the soil are readily

controlled in the centrifuge environment. Centrifuge modelling provides a method for

achieving stress and strain similitude between the prototype and a reduced scale

laboratory model. Prototype body forces and stress conditions are replicated by

subjecting a model reduced in scale by a factor ncent, to a radial acceleration field equal

Page 52: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

18

to ncentg (where g is the acceleration due to the Earth’s gravitational acceleration). For a

given soil type, the vertical stress at a depth, zm, in a model subjected to a radial

acceleration field of ncentg, will be identical to that in prototype at a depth, zp = ncentzm.

This fundamental scaling law, together with dimensional analysis enables extrapolation

of model test results to prototype conditions.

The centrifuge testing program will investigate DEPLA installation, keying and

capacity, and provide early stage performance data, necessary to provide initial proof of

concept and justification for subsequent field testing. However, it is important to

validate the centrifuge findings and further qualify the DEPLA concept through reduced

scale field testing in an aquatic environment. Two test locations have been identified; (i)

Lower Lough Erne, which is an inland lake located in County Fermanagh, Northern

Ireland, and (ii) the Firth of Clyde which is located off the West coast of Scotland

between the mainland and the Isle of Cumbrae. Reduced scale model anchors will be

deployed at both locations to assess DEPLA installation, keying and capacity. These

anchors will be instrumented with a custom built six degree of freedom (6DoF) inertial

measurement unit (IMU) for measurement of anchor motion during installation. This

will represent the first reported use of a 6DoF IMU for a geotechnical application.

Hence, there is a requirement to develop a framework for interpretation of the IMU

data. The IMU data will be used to validate an embedment prediction model for

dynamically installed anchors based on strain rate enhanced shear resistance and fluid

mechanics drag resistance.

1.4 Thesis structure

The thesis is organised as a series of academic papers that have been either published or

submitted for publication in accordance with the policies of the University of Western

Australia regarding higher degrees by research. Chapters 2 and 3 comprise the literature

review. Chapters 4 to 8 are journal papers that have been published, accepted for

Page 53: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

19

publication or are under review. These chapters report the experimental development

and results of an experimental study that involved a centrifuge testing program and an

extensive field testing campaign. Details of the contents of each chapter are provided

below.

Chapter 2: This chapter reviews the literature related to dynamic penetration of

projectiles into clayey seabeds. The review details field and laboratory studies on

free-falling penetrometers and dynamically installed anchor systems. The soil

mechanics, fluid dynamics and unified frameworks for predicting the shear strain

rate dependence of soil strength are discussed. Analytical and numerical methods

developed for predicting the embedment depth of projectiles penetrating the seabed

are described.

Chapter 3: This chapter reviews the literature related to the capacity of plate

anchors in clay. Experimental and numerical studies on the keying behaviour and

capacity of plate anchors are discussed. The review considers the factors that

influence the embedment depth loss associated with the keying process and

subsequent plate anchor capacity.

Chapter 4: This chapter presents results of a centrifuge testing program on reduced

scale model anchors in kaolin clay, carried out to attain early stage anchor

performance data necessary to provide motivation for subsequent field testing

(described in Chapters 5 to 8). The tests assessed the installation and capacity of

DEPLAs. The impact velocity, initial embedment depth following dynamic

installation, follower extraction, embedment depth loss due to the keying process

and the subsequent plate anchor capacity are examined. The results are compared

with corresponding experimental and numerical results for other dynamically

installed anchors (outlined in Chapter 2) and plate anchors (outlined in Chapter 3).

The methodology and outcomes have been presented in a journal paper:

Page 54: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

20

O’Loughlin, C. D., Blake, A. P., Richardson, M. D., Randolph, M. F. and Gaudin,

C. (2014). Installation and capacity of dynamically embedded plate anchors as

assessed through centrifuge tests. Ocean Engineering, Vol. 88, pp. 204–213.

Chapter 5: This chapter describes the technical development required to enable

field testing, and presents motion data captured by an inertial measurement unit

(IMU) during field tests of DEPLAs, DPAs and an instrumented free-fall sphere. A

comprehensive framework for interpreting the motion data is described and

validated against direct measurements. The motion data is used to verify the

appropriateness of an embedment prediction model for dynamically installed

anchors (outlined in Chapter 2). The framework is subsequently adopted in the

analysis of DEPLA field data in Chapters 6 and 8. The methodology and outcomes

have been presented in a journal paper:

Blake, A. P., O’Loughlin, C. D., Morton, J. M., O’Beirne, C., Gaudin, C. and White

D. J. (2016). In situ measurement of the dynamic penetration of free-fall projectiles

in soft soils using a low cost inertial measurement unit. Geotechnical Testing

Journal, forthcoming

Chapters 6 and 7: These chapters present results from field tests on reduced scale

model DEPLAs conducted at Lough Erne, Northern Ireland, required to validate the

centrifuge findings and demonstrate DEPLA deployment and operation in an

aquatic environment. Chapter 6 focuses on the behaviour of DEPLAs during free-

fall in water and dynamic embedment in soil. The framework for interpreting

motion data captured by an IMU described and validated in Chapter 5 is applied to

the data reported in Chapter 6 to derive net resistance profiles and velocity profiles.

Theses profiles are used to determine the hydrodynamic properties of DEPLAs, and

Page 55: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

21

refine and calibrate an embedment prediction model for dynamically installed

anchors (outlined in Chapter 2).

Chapter 7 considers follower extraction, the embedment depth loss due to the keying

process and the subsequent plate anchor capacity. The results are compared with

corresponding experimental and numerical results for other plate anchors (outlined

in Chapter 3).

The methodology and outcomes of Chapters 6 and 7 have been presented in two

separate journal papers:

Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded

plate anchors as assessed through field tests. Canadian Geotechnical Journal, Vol.

52, No. 9, pp. 1270–1282 (Chapter 6).

Blake, A. P., O’Loughlin, C. D. and Gaudin, C. (2014). Capacity of dynamically

embedded plate anchors as assessed through field tests. Canadian Geotechnical

Journal, Vol. 52, No. 1, pp. 87–95 (Chapter 7).

Chapter 8: This chapter presents results from field tests on reduced scale model

DEPLAs conducted at the Firth of Clyde, off the West coast of Scotland, required to

demonstrate DEPLA deployment and operation in an offshore environment. Chapter

8 examines: (i) DEPLA behaviour during free-fall in water and dynamic embedment

in soil and (ii) the embedment depth loss due to the keying process and the

subsequent plate anchor capacity. The results from the Firth of Clyde tests are

considered in parallel with centrifuge data (from Chapter 4) and field data (from

Chapters 6 and 7) to assess the merit of anchor embedment and capacity models. A

simple design approach for the prediction of DEPLA capacity based on these

models is developed. The methodology and outcomes have been presented in a

journal paper:

Page 56: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

22

O’Loughlin, C. D., Blake, A. P. and Gaudin, C. (2016). Towards a simple design

procedure for dynamically embedded plate anchors. Géotechnique, under review.

Chapter 9: Summarises the main findings of the research.

Appendix I: Contains a conference paper which describes a series of centrifuge

tests carried out to investigate the potential anchor capacity increase due to

reconsolidation (setup) for directly embedded plate anchors following the keying

process:

Blake, A.P., O’Loughlin, C. D. and Gaudin, C. (2010). Setup following keying of

plate anchors assessed through centrifuge tests in kaolin clay. Proceedings of the

2nd International Symposium on Frontiers in Offshore Geotechnics, Perth,

Australia, 8–10 November 2010, pp. 705–710.

Appendix II: Contains a conference paper which presents early stage results from

the field testing campaign reported in this thesis:

Blake, A. P. and O’Loughlin, C. D. (2012). Field testing of a reduced scale

dynamically embedded plate anchor. Proceedings of the 7th International

Conference in Offshore Site Investigation and Geotechnics, London, United

Kingdom, 12–14 September 2012, pp. 621–628.

Appendix III: Contains a conference paper which provides a brief overview of the

body of work reported in this thesis:

O’Loughlin, C. D., Blake, A. P., Wang, D., Gaudin, C. and Randolph, M. F. (2013).

The dynamically embedded plate anchor: results from an experimental and

numerical study. Proceedings of the 32nd International Conference in Ocean,

Offshore and Artic Engineering, Nantes, France, 9–14 June 2013, OMAE2013-

11571.

Page 57: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

23

CHAPTER 2 DYNAMIC EMBEDMENT OF

PROJECTILES INTO THE SEABED

2.1 Introduction

As offshore clay deposits are typically characterised by an increase in shear strength

with depth, the capacity of dynamically installed anchors such as the DEPLA is

dependent on the embedment depth achieved following free-fall in water. Hence

accurate prediction of the embedment depth is necessary in order to successfully

estimate the subsequent anchor capacity. The dynamic embedment of projectiles into

the seabed has previously been considered in studies of free-falling penetrometers

(FFPs) (see Section 2.2) and dynamically installed anchoring systems (see Section 2.3).

Several of these studies have adopted analytical and numerical methods to predict

embedment depth (see Section 2.5). However the accuracy of these methods is

questionable due to uncertainties regarding fluid drag resistance (see Section 2.4) and

viscous strain rate effects (see Section 2.4.3).

2.2 Free-falling penetrometers

FFPs are expendable or retrievable probes that are designed to dynamically penetrate

the seabed following free-fall through water. FFPs have been considered for seabed

strength characterisation, naval mine countermeasure research, paleolimnology

applications and the disposal of high-level nuclear waste in ocean abyssal plain

formations. FFP systems that have been developed for seabed strength characterisation

include (shown in Figure 2.1): marine impact penetrometer (MIP, Dayal 1975), marine

sediment penetrometer (MSP, Colp et al. 1975), free-fall penetrometer (Denness et al.

1981), expendable bottom probe (XBP, Akal and Stoll 1995), free-fall Cone

penetrometer (FF-CPT, Mulukutla 2009), CPT lance (Stegmann 2007), Nimrod (Stark

et al. 2009a) and lance insertion retardation meter (LIRmeter, Stephan et al. 2012).

Page 58: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

24

(a) (b)

(c) (d) (e)

(f) (g) (h)

Figure 2.1. FFPs developed for seabed characterisation (a) MIP (after Dayal et al.

1975), (b) MSP (after Colp et al. 1975), (c) Free-fall penetrometer (after Denesse et

al. 1981), (d) XBP (after Stoll and Akal 1999), (e) FF-CPT (www.brooke-

ocean.com), (f) CPT-Lance (after Strak et al. 2009b), (g) Nimrod (after Strak et al.

2009b) and (h) LIRmeter (after Stephan et al. 2012)

Page 59: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

25

For naval mine countermeasure research the following FFP systems have been reported

(shown in Figure 2.2): expendable doppler penetrometer (XDP), seabed terminal impact

naval gauge (STING), electronic strength profiler (ESP), burying mock mine body

(BMMB), Australian underwater sediment strength instrument (AUSSI), FAU

Experimental penetrometer (FEP) and Proboscis (PROBOS) (Beard 1977, Lott and

Poeckert 1996, Stoll 2004). An FFP system was also developed for paleolimnology

research that involved the measurement of the relative hardness and consistency of lake-

bottom sediments (Spooner et al. 2004).

2.2.1 Field studies

Numerous field studies have been carried out to assess the feasibility of FFP systems

such as the MIP (Dayal 1975, Chari et al. 1978), MSP (Colp et al. 1975), XBP (Stoll

and Akal 1999, Strak and Wever 2009), FF-CPT (Furlong et al. 2006, Mulukutla 2009,

Mulukutla et al. 2011, Stegmann et al. 2006, Steiner et al. 2012, Steiner et al. 2014),

XDP (Beard 1977, 1981, 1985, Bowman et al. 1995, Douglas and Wapner 1996,

Thompson et al. 2002, Ortman 2008) and STING (Mulhearn et al. 1998, Mulhearn et al.

1999, Mulhearn 2003, Abelev et al. 2009a, 2009b). The field tests have taken place at

various locations including: the Baltic Sea, the Black Sea, the Gulf of Mexico, New

York Harbour and around the coasts of Italy, Spain, Germany, Turkey (Stoll and Akal

1999). The FFPs were deployed in water depths of 7 to 5430 m and achieved impact

velocities of 2.2 to 29.6 m/s. Embedment depths of 0.1 to 7.9 penetrometer lengths were

reported. Significantly lower embedment ratios were observed in sandy soil than

compared to soft clay which suggests limited applications for FFPs in granular

sediments. The results of the field studies indicated that the final embedment ratio of a

FFP was found to be dependent on impact velocity, FFP geometry, FFP mass and soil

type.

Page 60: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

26

(a)

(b) (c) (d)

(e) (f) (g)

Figure 2.2. FFPs developed for naval mine countermeasure studies (a) XDP

(www.sonatech.com), (b) STING (after Abelev et al. 2009b), (c) ESP (after

Mulhearn et al. 1998), (d) BMMB (after Chow 2013), (e) AUSSI (after Mulhearn et

al. 1999), (f) FEP (after Chow 2013) and (g) PROBOS (after Stoll et al. 2007)

Page 61: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

27

Studies have also been carried out to assess the feasibility of using FFPs as oceanic

waste carriers for the disposal of high-level nuclear waste in ocean abyssal plain

formations (Ove Arup and Partners 1982, Freeman et al. 1984, Freeman and Burdett

1986, Freeman et al. 1988, Murray 1988). Field tests of oceanic waste carriers were

conducted at the Great Meteor East (GME) radioactive waste disposal site in the eastern

Atlantic Ocean (Freeman et al. 1984, Freeman et al. 1988) and the Nares Abyssal Plain

(NAP) in the western Atlantic Ocean (Freeman and Burdett 1986). The field tests

involved various penetrometer types including the European standard penetrator (shown

in Figure 2.3). Results indicated that impact velocities of 30 to 68 m/s were achieved

resulting in tip embedments of 29 to 58 m (8.1 to 16.4 penetrometer lengths) in the soft

seabed sediments encountered at the test sites. Tests were also conducted off the coast

of Antibes in the Mediterranean Sea where the seabed sediment was much stiffer and

stronger than encountered at the GME and NAP test sites resulting in much lower tip

embedments of 9 to 15 m (3.5 to 4.2 penetrometer lengths) (Audibert et al. 2006).

Figure 2.3. European standard penetrometer (after Freeman et al. 1984)

Page 62: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

28

2.2.2 Laboratory studies

The application of FFP systems for seabed strength characterisation is not widespread

mainly due to uncertainties regarding how to account for strain rate effects (see Section

2.4.3) which are attributed to the high penetration rates achieved by FFPs, in order to

obtain reliable estimates of undrained shear strength. Laboratory studies on model FFPs

have been carried out at 1 g to investigate the possible dependence of strain rate effects

on soil strength (Dayal 1974, Levacher 1985, Akal and Stoll 1995, Chow and Airey

2014), FFP mass (Dayal 1974, Chow and Airey 2014), tip shape (Dayal 1974, Levacher

1985, Chow and Airey 2014), tip diameter (Chow and Airey 2014) and impact velocity

(Dayal 1974, Levacher 1985, Stoll et al. 2007, Chow and Airey 2014). The results of

the studies indicated that strain rate effects are independent of FFP mass and tip

diameter. Levacher (1985) found the influence of tip shape on strain rate effects to be

negligible whereas Dayal (1974) reported results to the contrary. Test results

demonstrated that strain rate effects increase as the undrained shear strength decreases

(Dayal 1974, Chow and Airey 2014). Studies revealed that strain rate effects increase

for tests involving impact velocities of up to vi = 6.1 m/s (Dayal 1974, Stoll et al. 2007)

and reach a limit at a critical velocity e.g. vi > 6.0 m/s (Levacher 1985) or vi ≈ 5 m/s

Chow and Airey et al. (2014).

A series of centrifuge tests on oceanic waste carriers were conducted on a 60 mm long,

6 mm diameter, 13 gram projectile at 100 g (6 m long, 0.6 m diameter, 13 t at prototype

scale) in normally consolidated kaolin clay (Poorooshasb and James 1989). The tests

investigated the embedment depth, deformation pattern and degree of hole closure

associated with dynamic installation of the projectile. Results suggested that projectile

tip embedments of at least 30.5 m (at prototype scale) or 5.1 penetrometer lengths were

achievable for an impact velocity of 40 m/s.

Page 63: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

29

2.3 Dynamically installed anchors

Three types of dynamically installed anchor have been developed: the torpedo anchor

(also referred to as the torpedo pile), the Deep Penetrating Anchor (DPA) and the

OMNI-Max anchor. Each type of anchor has been developed by a different

manufacturer and has slightly different features. These anchors are released from a

designated drop height above the seabed and penetrate to a target depth in the seabed by

the kinetic energy obtained through free-fall (Figure 1.18). Resistance to loading is

predominantly provided by the skin friction developed at the anchor-soil interface.

2.3.1 Torpedo anchor

In 1996 the torpedo anchor was proposed as an efficient anchoring system for flexible

risers and floating vessels in soft seabed sediments. Torpedo anchors comprise of a

tubular steel pile, with or without fins, with a conical tip and filled with scrap chain or

concrete. A mooring line is connected to an omni-directional padeye located at the top

of the anchor (Medeiros 2001, 2002).

Full scale field tests were conducted in the Campos Basin off the east coast of Brazil in

water depths ranging from 200 to 1000 m. Tests were carried out using 12 m long

(finless) torpedo anchors with a 0.762 m diameter and a dry mass of 40.8 t. The anchors

were released from a height of 30 m (2.5 anchor lengths) above the seabed and the

following tip embedments (ze) were reported:

29 m (2.4 anchor lengths) in normally consolidated clay;

13.5 m (1.1 anchor lengths) in overconsolidated clay;

15 m (1.3 anchor lengths) in uncemented calcareous sand;

22 m (1.8 anchor lengths) in normally consolidated clay overlain by a 13 m thick

fine sand layer.

Page 64: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

30

Over 1000 torpedo anchors have been installed in offshore Brazil for the mooring of

flexible risers, MODUs and FPSOs (Wilde 2009). During 2005 the P-50 FPSO became

the first floating facility to be permanently moored using torpedo anchors. The P-50 is

moored with eighteen 98 t torpedo anchors in 1240 m of water at the Albacora Leste

field in the Campos Basin. The anchors were 17 m long with a 1.07 m diameter and four

10 m × 0.9 m fins. (Figure 2.4) (Brandão et al. 2006).

Figure 2.4. 98 t torpedo anchor for the Albacora Leste field in the Campos Basin

(after Brandão et al. 2006)

2.3.2 Deep penetrating anchor

In 1999 the Deep Penetrating Anchor (DPA) was conceived as a cost effective

anchoring solution for mooring floating facilities in soft seabed sediments (Lieng et al.

1999). DPAs comprise of a dart shaped, thick walled steel cylinder with fins (flukes)

attached to its upper section. A mooring line is connected to a padeye located at the top

of the anchor (Figure 2.5).

During 2003 twelve 1:3 reduced scale model DPA field tests were carried out in water

depths exceeding 300 m at the Trondheim fjord which is situated in the west central part

of Norway. The model DPA was 4.4 m long with a dry mass of 2.8 t and instrumented

Page 65: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

31

with an accelerometer, two pore water pressure sensors, an inclinometer and a depth

sensor. The results of these field tests are yet to be published.

In 2008 twelve 1:3 reduced scale model DPA field tests were carried at the Troll Field

in the North Sea off the west coast of Norway. The DPAs were dropped from heights

of 15 m and 75 m (3.4 and 17 anchor lengths) achieving impact velocities of 13 m/s and

15 m/s respectively, corresponding to tip embedments of 7 m and 9 m (1.6 and 2 anchor

lengths) (Figure 2.6) (Strum et al. 2010).

Figure 2.5. Deep penetrating anchor (www.ngi.no)

Page 66: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

32

Figure 2.6. Measured and predicted impact velocities and tip embedment depths

for 1:4 scale model DPA test at Troll Field – grey band denotes range of measured

values (after Strum et al. 2011)

In 2009 two full scale DPAs were installed at the Gjøa Field in the North Sea off the

west coast of Norway in 360 m of water as part of the mooring system for the

TransOcean Searcher MODU. The anchors were 13 m long with four 1 m wide flukes

and a dry mass of 80 t. The DPAs were dropped from heights of 50 m and 75 m (1.2

and 1.7 anchor lengths) achieving impact velocities of 24.5 m/s and 27 m/s respectively,

corresponding to tip embedments of 24 m and 31 m (1.9 and 2.4 anchor lengths) (Lieng

et al. 2010).

2.3.3 OMNI-Max anchor

In 2005 the OMNI-Max anchor was proposed as an anchoring solution for mooring

floating facilities in soft seabed sediments (Zimmerman and Spikula 2005). The OMNI-

Max anchor is arrow shaped with a mooring padeye located at the end of a hinged load

arm which can rotate 360º about the longitudinal axis of the anchor enabling it to be

loaded in any direction (Figure 2.7). The anchors have retractable fluke fins which

allow the anchor to be tailored to rotate into the soil and begin diving at a preferred

Tip

em

bed

men

t d

epth

, z e

[m]

High estimateLow estimate

Impact velocity, vi [m/s]

Page 67: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

33

tension level. Once significant load is applied to the anchor it begins to rotate into the

soil until the lateral resistance of the lower fluke fins becomes equal to the lateral

resistance of the upper fluke fins. Once this level of loading is exceeded the anchor

performance is determined by its axial capacity and begins to penetrate deeper into the

seabed mobilising stronger soil and gaining additional capacity (Shelton et al. 2011).

Figure 2.7. OMNI-Max anchor (after Zimmerman et al. 2007)

During 2006 and 2007 full scale OMNI-Max anchor field tests were carried out at

Viosca Knoll in the Gulf of Mexico in 427 m of water using a 9 m long anchor with a 3

m fin span and dry mass of 34 t. The soil at the test site was soft to medium clay and the

anchor was released from a height of 76 m (8.4 anchor lengths) above the seabed. The

test results demonstrated the stability of the anchor during free-fall (Zimmerman 2007).

Over 160 full scale OMNI-Max anchors have been installed in the Gulf of Mexico for

the mooring of MODUs (Shelton et al. 2011).

2.3.4 Laboratory studies

Massey (2000) conducted physical modelling using 1:200 reduced scale model DPAs at

1 g in kaolin clay to investigate the relationship between impact velocity, embedment

depth and holding capacity. Prototype impact velocities of 20 to 25 m/s (Lieng et al.

1999) could not be achieved at 1 g.

Page 68: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

34

Beam centrifuge tests were conducted at the University of Western Australia (UWA) on

1:200 reduced scale model DPAs in kaolin clay (O’Loughlin et al. 2004a, Richardson

2008, O’Loughlin et al. 2009). Centrifuge modelling enables prototype stresses and

velocities to be replicated. The reduced scale model anchors varied in aspect ratio

(length/diameter), mass, fluke arrangement and tip geometry. Typically, impact

velocities achieved during the tests ranged from 10 to 30 m/s. Typical anchor tip

embedments were in the range 150 to 220 mm (30 to 44 m in prototype scale) or 2.1 to

2.9 anchor lengths. The results demonstrated that anchor embedment depth increases

with increasing impact velocity and mass (Figure 2.8), and decreases with increasing

aspect ratio (Length/diameter ratio) (Figure 2.9).

Figure 2.8. Dependence of anchor tip embedment depth on impact velocity and

mass (after O’Loughlin et al. 2013b)

Tip

em

be

dm

en

t d

ep

th,

z e(m

)

Page 69: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

35

Figure 2.9. Dependence of anchor tip embedment depth on aspect ratio (after

O’Loughlin et al. 2013b)

Parametric studies involving drum centrifuge tests at UWA were conducted to assess

the influence of anchor geometry (diameter and aspect ratio) and mass on the

embedment of DPAs (Richardson et al. 2006, Richardson 2008). Impact velocities of 9

to 20 m/s were achieved corresponding to embedment depths of 20 to 110 mm (4 to 22

m in prototype scale). The results demonstrated that anchor embedment depth increase

with increasing aspect ratio and hence increasing mass (Figure 2.10).

Figure 2.10. Dependence of anchor tip embedment depth on mass (after

Richardson et al. 2006)

Tip

em

be

dm

en

t d

ep

th,

z e(m

)

L

d

0

20

40

60

80

100

120

0 2 4 6 8 10 12 14 16 18 20 22

Tip

em

bd

emen

t d

epth

, ze(m

)

Impact velocity, vi (m/s)

Page 70: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

36

Laboratory tests were conducted at The University of Texas at Austin on 1:30 reduced

scale torpedo anchors at 1 g in kaolin clay to assess the accuracy of an embedment

depth prediction model (see Section 2.5.1) (Audibert et al. 2006, Gilbert et al. 2008).

The embedment depths predicted by the model were generally within +/- 10 % of the

measured (Figure 2.11).

Figure 2.11. Comparison of measured and predicted tip embedment depths for

1:30 reduced scale model torpedo anchors in kaolin clay (after Gilbert et al. 2008)

Fernandes et al. (2006) conducted tests on a 1:15 reduced scale model torpedo anchor in

a water tank test to investigate anchor drag coefficient and directional stability. The test

results suggested that the anchor fins and mooring line have a positive effect on

directional stability. The anchor drag coefficient was estimated using an extrapolating

mathematical model.

Cenac Π (2011) conducted free-fall and tow tank tests in water on a 1:24 reduced scale

model OMNI-Max anchors at Texas A&M University to investigate drag coefficients

and embedment depth. The results indicated that the model anchor achieved an average

embedment depth of 1.5 times the anchor length in an artificial mud mixture following

free-fall in water and that the embedment depth increased with an increase in impact

velocity (Figure 2.12).

0

300

600

900

1200

1500

0 300 600 900 1200 1500

Pre

dic

ted

tip

em

bed

men

t (m

m)

Measured tip embedment (mm)

Pre

dic

ted

tip

em

bed

men

t (i

n.)

Measured tip embedment (in.)

Page 71: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

37

Figure 2.12. Dependence of anchor tip embedment on impact velocity (after Cenac

Π et al. 2011)

A parametric study was carried out at Texas A&M University to assess the

hydrodynamic properties of OMNI-Max anchors (Shelton et al. 2011). The study

involved free-fall and tow tank tests in water, and wind tunnel tests on 1:24 and 1:15

reduced scale model OMNI-Max anchors. The hydrodynamic parameters obtained from

the tests were used in an embedment depth prediction model (see Section 2.5.1). The

model’s lower and upper bound embedment depth predictions of 12.8 m and 18.9 m

respectively, agreed favourably with field data for full scale OMNI-Max anchors

installed in the Campos Basin offshore Brazil (Figure 2.13).

Figure 2.13. Comparison of anchor embedment prediction model results and

OMNI-Max anchor field data (after Shelton et al. 2011)

Tip

em

be

dm

ent d

epth

, ze

(in

)

Impact velocity, vi (ft/s)

Tip

em

be

dm

en

t d

ep

th, z

e(ft

)

0

5

10

15

20

25

Tip

em

bed

ment

dep

th,

z e(m

)

Page 72: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

38

Hasanloo et al. (2012) carried out tests on 1:40 reduced scale model torpedo anchors

(Figure 2.14) in a water tank to assess the influence of anchor weight, aspect ratio, scale

ratio and fin size on anchor velocity during free-fall prior to impact with the seabed.

The test results indicated that the directional stability, density ratio, aspect ratio and

scale ratio had a positive impact on the free-fall velocity of the anchors, while the

frontal area and fin length had a negative impact (Figure 2.15).

A centrifuge study was carried out at UWA on a 1:200 reduced scale OMNI-Max

anchor (Figure 2.16) in overconsolidated kaolin clay and calcareous silt from the North

West Shelf of Australia (Gaudin et al. 2013). Impact velocities of 10.4 to 23 m/s were

measured in the tests in the overconsolidated kaolin clay and corresponding embedment

depths in the range 1.18 to 2 anchor lengths (ze = 10.7 to 18.1 m in prototype scale).

Significantly lower embedment depths of 1.14 to 1.46 anchor lengths (ze = 10.3 to 13.2

m in prototype scale) were achieved in the calcareous silt despite higher impact

velocities of 20.5 to 29.4 m/s (Figure 2.17). Gaudin et al. (2013) suggested that this was

due to the higher undrained shear strength and potentially dilatant behaviour of the

calcareous silt.

Figure 2.14. 1:40 reduced scale model torpedo anchor (after Hasanloo et al. 2012)

Page 73: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

39

(a) (b)

(c) (d)

Figure 2.15. Comparison between the drag coefficients of torpedo anchors with

different (a) densities, (b) aspect ratio, (c) scale ratios, (d) fin sizes (after Hasanloo

et al. 2012)

Figure 2.16. 1:200 reduced scale model OMNI-Max anchor (after Gaudin et al.

2013)

Density, ρ (kg/m3)

L/D = 9.375

L/D = 6.25

L/D = 12.5L

D

0.593*A

1.187*A

A

No fins

20 cm fins

10 cm fins

L

Page 74: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

40

Figure 2.17. Embedment depths achieved by 1:200 reduced scale OMNI-Max

anchors from centrifuge tests compared with field data (after Gaudin et al. 2013)

Hossain et al. (2014) reported results from a series of drum centrifuge tests on 1:200

reduced scale torpedo anchors (Figure 2.18) in lightly over consolidated kaolin clay and

calcareous silt at UWA. Impact velocities of 14.9 to 22 m/s were measured

corresponding to embedment depths of 1.2 to 2.1 anchor lengths (Figure 2.19). The

embedment was noted to be 1.6 times lower in the calcareous silt than in the clay for a

similar impact velocity and attributed to the significantly higher shear strength of the

silt. The cavity formed by the anchor during installation was observed to remain open in

the silt and fully covered in the clay.

Tip

em

be

dm

en

t d

ep

th,

z e(m

)

No

rmal

ise

d e

mb

ed

me

nt

de

pth

rat

io,

z e/L

Impact velocity, vi (m/s)

Page 75: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

41

Figure 2.18. 1:200 reduced scale torpedo anchors (after Hossain et al. 2014)

Figure 2.19. Embedment depths achieved by 1:200 reduced scale torpedo anchors

from centrifuge tests compared with field data (after Hossain et al. 2014)

2.4 Resistance forces acting on projectiles during dynamic embedment in

soil

Resistance forces (Figure 2.20) that have been considered in the prediction of the

dynamic embedment of projectiles in soil include: fluid drag (see Section 2.4.1), added

mass (see Section 2.4.2), strain enhanced soil resistance (bearing and frictional) and soil

buoyancy (see Section 2.5).

Impact velocity, vi (m/s)

Tip

em

be

dm

en

t d

ep

th, z

e(m

)

No

rmal

ise

d e

mb

ed

me

nt

de

pth

, ze/L

Page 76: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

42

Figure 2.20. Forces acting on a DEPLA during installation

2.4.1 Fluid drag resistance

As an object moves through a viscous fluid it is subjected to fluid drag resistance which

is attributed to: (i) friction drag (commonly referred to viscous drag) generated by

viscous stresses that develop in the boundary layer between the object and fluid, (ii)

pressure drag (also known as form drag and is independent of viscous effects) caused by

an adverse pressure gradient (related to the stagnation pressure defined in Equation

Water

Lakebed

Soil buoyancy

(Fb)

Fluke reverse

end bearing

(Fbear)

Fluke friction

(Ffrict)

Sleeve friction

(Ffrict)

Sleeve end bearing

(Fbear)

Fluke end bearing

(Fbear)

Follower Friction

(Ffrict)

Submerged weight

(Ws)

Follower tip end

bearing (Fbear)

Anchor fluid drag

(Fd,a

)

Follower recovery

line fluid drag

(Fd,l

)

Mooring line fluid

drag (Fd,l

)

Page 77: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

43

2.15) between the front and rear of the object which creates a force opposite to the

direction of motion (Hoerner 1965). The fluid drag resistance is generally expressed as

(Morison et al. 1950):

2

2

1AvCF dd Equation 2.1

where Cd is the drag coefficient of the object, ρ is the fluid density, A is the reference

area (e.g. frontal area of the object) and v is the object’s velocity. The drag coefficient is

dependent on the objects geometry, surface roughness and Reynolds number of the

associated flow.

For a Newtonian fluid (e.g. water) the relationship between shear stress and shear strain

is linear and the Reynolds number is defined as (Hoerner 1965):

DvRe Equation 2.2

where D is the projectile diameter and ν is the kinematic viscosity of the fluid.

During dynamic penetration of a projectile into the seabed, the projectile will transition

from the water into the soil. The viscosity and hence the Reynolds number of the

associated flow will vary from the water to the soil. However, it is commonly assumed

that the drag coefficient of a projectile is the same in both soil and water.

A relatively wide range of drag coefficient values have been reported for seabed

penetrating projectiles, which reflects the variation in projectile geometry and Reynolds

number:

True (1976) suggested a drag coefficient of Cd = 0.7 to be adopted to account for

fluid drag resistance in soil and water for a range of projectile velocities and

geometries.

Page 78: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

44

Back-analysis of results from field tests on oceanic waste carriers indicated drag

coefficients in the range Cd = 0.1 to 0.18 (Freeman et al. 1984, Freeman and Burdett

1986, Freeman and Murray 1988).

A Computational Fluid Dynamics (CFD) study by Øye (2000) on DPAs suggested a

drag coefficient of Cd = 0.63 for a four fluke anchor in water.

CFD analysis showed that drag coefficient decreases with increasing anchor aspect

ratio, L/d, from Cd = 0.35 for L/d = 1 (i.e. a sphere) to Cd = 0.23 for L/d ≥ 4

(representing Reynolds number values of approximately 1×106 to 7×10

7) for

anchors with no flukes (Richardson 2008).

Fernandes et al. (2006) extrapolated a drag coefficient of Cd = 0.33 from results of a

hydrodynamic study on a 1:15 reduced scale model torpedo anchor and noted that

the mooring line increased the total drag resistance.

Free-fall and tow tank tests on 1:15 reduced scale model OMNI-Max anchors

(Cenac Π 2011) indicated drag coefficients in the ranges: Cd = 0.46 to 0.83 and Cd =

0.7 to 1.12 from the free-fall and tow tank tests respectively and demonstrated the

dependence of drag coefficient on Reynolds number (Figure 2.21). Prototype scale

drag coefficients in the range Cd = 0.65 to 0.87 were estimated from the results. The

mooring line was found to increase the drag resistance by up to 13.5%. The drag

coefficient was observed to increase as the anchor fins were retracted.

Results from free-fall tests on 1:40 reduced scale model torpedo anchors (Hasanloo

et al. 2012) of various weight, aspect ratio, scale ratio and fin size indicated drag

coefficients in the range Cd = 0.2 to 1.2 (representing Reynolds number values of in

the range 4.8×105 to 2.16×10

6), and demonstrated the dependence of drag

coefficient on Reynolds number (Figure 2.21).

Page 79: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

45

(a)

(b)

Figure 2.21. Dependence of drag coefficient on Reynolds number for (a) torpedo

anchor (after Hasanloo et al .2012) and (b) OMNI-Max anchor (after Cenac Π

2011)

2.4.2 Added mass

An object moving in a fluid must accelerate some volume, Va, of the surrounding fluid

from rest to a velocity capable of displacing the fluid out of its path as the object and the

fluid cannot occupy the same physical space simultaneously. The reaction to the force

required to accelerate the volume of fluid creates an inertia force which is defined as

(Morison et al.1950):

aVCF ami Equation 2.3

where Cm is the inertia coefficient and a is the object’s instantaneous acceleration:

am CC 1 Equation 2.4

Cd

Re

Cd

Re

Page 80: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

46

and Ca is the added mass coefficient.

A simplified method for estimating the inertia force is to assume that a constant mass of

fluid (commonly referred to as ‘added mass’) is moving along with the object at the

same acceleration (Fernandes et al. 2006, Kunitaki et al. 2008):

amF ai Equation 2.5

where ma is the added mass, which is a function of the object’s shape and direction of

motion. For slender bodies such as FFPs and dynamically installed anchors the added

mass is considered to be negligible and the inertia force is ignored in motion analysis

(e.g. Beard et al. 1981, Shelton 2011).

2.4.3 Shear strain rate effects in clay

There are two distinct frameworks for predicting the shear strain rate dependence of soil

strength: a soil mechanics framework and a fluid dynamics framework. The soil

mechanics framework generally expresses the soil shear strength using a power or semi-

logarithmic law that accounts for strain rate effects (see Section 2.4.3.1). The fluid

dynamics framework treats the soil as a non-Newtonian fluid and applies fluid

mechanics principles to estimate the soil shear strength (see Section 2.4.3.2). A ‘unified

framework’ that considers both soil mechanics and fluid mechanics principles has

recently been proposed (see Section 2.4.3.3).

2.4.3.1 Soil mechanics framework

The undrained shear strength of soil is considered to be a function of the shear strain

rate, (Casagrande and Wilson 1951, Graham et al. 1983, Sheahan et al. 1996). For

variable rate penetrometers, FFPs and dynamically installed anchors the ‘operational’

shear strain rate is usually taken as the normalised penetration rate, v/d (where v is

velocity and d is diameter, O’Loughlin et al. 2013b). At very high shear strain rates

Page 81: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

47

viscous effects dominate the undrained shear strength of soil and the undrained shear

strength increases with increasing shear strain rate (Figure 2.22) (Sheahan et al. 1996,

Lunne and Anderson 2007, Jeong et al. 2009). The dependence of the operational

undrained shear strength, su,op, on shear strain rate is typically accounted for using a rate

function, Rf:

ufopu sRs , Equation 2.6

where su is the undrained shear strength at the reference shear strain-rate, .γref The rate

function is generally expressed as either a semi-logarithmic (e.g. Graham et al. 1983,

Chung et al. 2006) or power function (e.g. Biscontin and Pestana 2001, Lehane et al.

2009):

ref

fR

log1 Equation 2.7

ref

fR

Equation 2.8

where λ and β are strain-rate parameters in the respective formulations representing the

increase in undrained shear strength.

Studies of FFPs have shown that the rate effect of the shaft friction component of

penetrometer resistance is greater than that of the tip resistance component. It is

believed that the higher rate effect for the shaft friction is due to the higher strain rate at

the cylindrical shaft involving curved bands. The greater rate effects can be accounted

for in the power function by adopting a coefficient, n (Dayal and Allen 1975, Steiner et

al. 2014, Chow et al. 2014):

Page 82: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

48

ref

f nR

Equation 2.9

where n is taken as 1 for tip resistance (Zhu and Randolph 2011) and as a function of β

for estimating rate effects in shaft resistance (Einav and Randolph 2006) according to:

22 1

1 nn

n

Equation 2.10

The strain-rate parameters have been found to be dependent on strain rate range (e.g.

Sheahan et al. 1996, Jeong et al. 2009, O’Loughlin et al. 2013b), projectile geometry

(Dayal 1974, Levacher 1985, Ortman 2008) and soil properties such as: plasticity (e.g.

Bjerrum 1973, Gibson and Coyle 1968, Jeong et al. 2009), moisture content (Gibson

and Coyle 1968, Dayal 1974, Abelev and Valent 2009), stress history (Sheahan et al.

1996, Lehane et al. 2009, Chow and Airey 2013), anisotropy (Nakase and Kamei 1986,

Zhu and Yin 2000) and sensitivity (Yafrate and Dejong 2007).

Figure 2.22. Illustration of viscous effects for normally consolidated and lightly

overconsolidated soils – representative of typical deepwater seabed deposits (after

Lehane et al. 2009)

Page 83: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

49

O’Loughlin et al. (2013b) back-calculated values of λ and β from results of centrifuge

studies on dynamically installed anchors in normally consolidated kaolin clay

(O’Loughlin et al. 2004a, Richardson et al. 2006, Richardson 2008, O’Loughlin et al.

2009). The strain-rate parameters were typically in the ranges β = 0.06 to 0.17 and λ =

0.2 to 1.0 (indicating a 20 to 100% increase in soil shear strength per log cycle increase

in anchor velocity or penetration rate) (Figure 2.23). The strain rates (proportional to

v/d) in the centrifuge tests are on average 200 times equivalent strain rates in the field,

as the absolute velocities are comparable, but the anchor diameter is scaled by 1:200.

The value of v/d in the centrifuge tests were in the range 500 to 4250 s-1

compared with

v/d ≈ 25 s-1

for a typical dynamically installed anchor in the field (Randolph et al.

2011). Hence the strain-rate parameters representative of field conditions are likely to

be at the lower extreme of this range and would therefore be similar to parameters

deduced from variable rate penetrometer tests.

Gaudin et al. (2013) reported back-calculated strain rate parameters from centrifuge

tests on OMNI-Max anchors in overconsolidated kaolin clay and calcareous silt of β =

0.16 and 0.19 respectively (Figure 2.24). However a fluid drag resistance term was

excluded from the analyses.

Page 84: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

50

Figure 2.23. Back-analysed strain rate parameters for reduced scale mode DPAs

from centrifuge tests in kaolin clay (after O’Loughlin et al. 2013b)

Figure 2.24. Back-analysed strain rate parameters for a reduced scale mode

OMNI-Max anchor from centrifuge tests in kaolin clay and calcareous silt (after

Gaudin et al. 2013)

Value of λ deduced from variable rate penetrometer tests (e.g. Randolph and Hope

2004, Chung et al. 2006, Boylan et al. 2007, Young et al. 2011, Steiner et al. 2012),

shear vane tests (e.g. Biscontin and Pestena 2001, Schlue et al. 2010), tri-axial

compression tests (e.g. Sheahan et al. 1996, Chow and Airey 2013) and FFP tests

Impact velocity, vi (m/s)

Anc

hor

tip

embe

dmen

t, z

e(m

)

Nor

mal

ised

em

bedm

ent

rati

o, z

e/L

Page 85: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

51

(Aubeny and Shi 2006, Chow and Airey 2013) in clay generally lie in the range λ = 0.05

to 0.25. Variable rate penetrometers tests in kaolin clay indicated β values in the range β

= 0.06 to 0.08 (Lehane et al. 2009). β values derived from shear vane tests (e.g.

Biscontin and Pestena 2001, Peuchan and Mayne et al. 2009, Schlue et al. 2010) in clay

are generally in the range β = 0.05 to 0.12. The power law rate function has been

reported to provide a better fit to experimental data over several log cycles of shear

strain rate than the semi-logarithmic function (Biscontin and Pestena 2001, Peuchan and

Mayne 2007, Randolph et al. 2007 and O’Loughlin et al. 2013b).

The nominal shear strain-rates of triaxial compression tests and in situ shear vane tests

are approximately 3 × 10-6

s-1

and 2 × 10-3

s-1

respectively (Einav and Randolph 2006).

These rates are four and seven orders of magnitude lower than the typical rate of 25 s-1

associated with dynamically installed anchors in the field. Therefore it is difficult to

extrapolate strain-rate parameters from laboratory and in situ tests to be adopted in the

predicting the embedment depth of dynamically installed anchors into the seabed.

2.4.3.2 Fluid mechanics framework

In the fluid mechanics framework the soil is treated as a non-Newtonian fluid and its

stress–strain behaviour is represented by the Herschel-Bulkley model which combines

plastic and shear-thinning effects, and is expressed as (Deglo de Besses et al. 2003):

Ky Equation 2.11

where τ is the mobilised shear stress (essentially equivalent to the operative undrained

shear strength, su,op), τy is the yield stress, K represents the viscosity property of the soil

(often referred to as the consistency parameter), γ is the shear strain rate and δ is the

shear thinning index. Within this framework the resistance forces are generally

estimated using a form of Equation 2.1 with drag coefficient expressed in terms of the

non-Newtonian Reynolds number, Renon-Newtonian (also referred to as Johnson number),

Page 86: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

52

which gives the relative magnitude of pressure drag related resistance or stagnation

pressure to the mobilised shear stress of the soil, τ, and is expressed as (Randolph and

White 2012):

2

Rev

Newtoniannon Equation 2.12

2.4.3.3 Unified framework

The power law model in the soil mechanics framework and the Herschel-Bulkley model

in fluid dynamics framework both represent the variation in mobilised shear stress with

shear strain rate. The shear strain rate effects are predicted using a multiplicative

approach within the soil mechanics framework and an additive approach within the fluid

dynamics framework. Zhu and Randolph (2011) proposed a unified formulation that

originates from the power law model but included an additive term similar to the

Herschel-Bulkley model which can be used to predict the shear strain rate effects of a

soil like material and a fluid-like material. This formulation is referred to as an ‘additive

power-law model’ and is expressed as:

min,, 1 u

ref

opu ss

Equation 2.13

where su,op is the operative undrained shear strength, η represents the viscosity property

of the soil, γ is operational shear strain rate, refγ is the reference shear strain rate, η

represents the viscosity property of the soil, δ is the shear thinning index, su,min is the

minimum undrained shear strength at zero strain rate, while the reference undrained

shear strength is given by:

min,1 uu ss Equation 2.14

Page 87: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

53

Equation 2.13 is essentially equivalent to the Herschel-Buckley model with su,min = τy, β

= δ and Ks refu

min,

. The ‘additive power-law model’ focuses on the viscous

component of resistance attributed to strain rate effects but neglects the pressure drag

component of fluid drag resistance which is independent of viscous effects and related

to the stagnation pressure, Pstag:

2

2

1vPP statstag Equation 2.15

where Pstat is the static pressure and 1/2ρv2 term is the dynamic pressure.

However Zhu and Randolph (2011) noted that for high velocity, low soil strength

conditions (i.e. Renon-Newtonian > 10) the pressure drag component starts to dominate the

fluid drag resistance and needs to be accounted for separately from the viscous

component.

Randolph and White (2012) proposed to express the forces exerted by a submarine land

slide on a pipe using the power law from the soil mechanics framework and a fluid drag

resistance term with a fixed drag coefficient (i.e. independent of strain rate effects) for

high velocity low soil strength conditions. This superposition approach treats the forces

that arise from the shear strain rate dependent strength and pressure drag separately

rather than combining them into a single term. Randolph and White (2012) expressed

the total normal force, Fn, acting per unit length of the pipe as:

2

,2

1vDCDsNF pdpopupn Equation 2.16

where Cd is the drag coefficient, ρ is the fluid density, v is the flow velocity, Dp is the

pipe diameter, Np is the bearing factor for the force normal to the pipeline, su,op is the

operative undrained shear strength estimated using the power law. Randolph and White

Page 88: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

54

(2012) observed that the relative magnitude between the two components in Equation

2.16 are such that the fluid drag resistance term starts to dominate when the non-

Newtonian Reynolds number exceeds a boundary value of Renon-Newtonian ≈ 10. Sahdi et

al. (2014) demonstrated that for non-Reynolds numbers higher than the boundary value

the resistance associated with the pressure drag increases linearly with non-Newtonian

Reynolds number and can be accounted for using a constant mean drag coefficient.

Studies involving assessment of the impact forces exerted by high velocity submarine

slides on pipelines (e.g. Boukpeti et al. 2012, Randolph and White 2012, Sahdi et al.

2014) and dynamically installed anchors (e.g. O’Loughlin et al. 2004a, Fernandes et al.

2006, Audibert et al. 2006, Richardson et al. 2006, Richardson 2008, O’Loughlin et al.

2009, Shelton et al. 2011, O’Loughlin et al. 2013b) have adopted the superposition

approach.

2.5 Embedment depth prediction

2.5.1 Analytical modelling

True (1976) proposed a method for predicting the dynamic embedment of projectiles

penetrating in soil after free-fall in water which considered Newton’s second law of

motion and the forces acting on the projectile during dynamic embedment. The method

considers the static forces resisting projectile embedment (Figure 2.20), remoulding of

the soil on the sides of the projectile and fluid drag resistance as the projectile passes

through the soil:

dbearfricts FFFW

dt

zdm

2

2

Equation 2.17

where m is the projectile mass, z is the projectile tip embedment, t is time, Ws is the

submerged weight of the projectile (in soil), Ffrict is frictional resistance and Fbear is

bearing resistance and Fd is the fluid drag resistance (see Section 2.4).

Page 89: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

55

The dynamic embedment of a projectile in soil is calculated using a finite difference

approximation of Equation 2.17.

The frictional and bearing resistance terms are expressed as:

AsF ufrict Equation 2.18

AsNF ucbear Equation 2.19

where α is a friction ratio (of limiting shear stress to undrained shear strength), Nc is the

bearing capacity factor for the projectile tip or fins (front and reverse end bearing) and

su is the undrained shear strength averaged over the contact area, A. Typically, during

dynamic penetration of projectiles in soil, the frictional resistance generated along the

wall of the foundation is close to the remoulded shear strength of the clay and the

friction ratio is often expressed as (Andersen et al. 2005):

u

ru

t s

s

S

,1 Equation 2.20

where St is the soil sensitivity and su,r is the remoulded strength of the soil.

The estimation of Nc has been influenced by the use of either shallow foundation

bearing capacity solutions or quasi-static cone penetrometer test interpretation methods.

Typically a constant Nc value is adopted for the projectile tip such as Nc = 12 (Beard

1977, O’Loughlin et al. 2004a, Richardson 2008, O’Loughlin et al. 2009, O’Loughlin et

al. 2013b), 15 (Freeman and Schuttenhelm 1990), 10 (Mulhearn et al. 1998), 17 (Gilbert

et al. 2008) and 14 (Steiner et al. 2012, Steiner et al. 2014). Nc = 7.5 has been adopted

for the upper and lower end of dynamically installed fins (Richardson 2008, O’Loughlin

et al. 2013b) which is based on a solution for a deeply embedded strip footing

(Skempton 1951). Numerical studies tend to express Nc as a function of embedment

Page 90: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

56

ratio (e.g. Aubeny and Shi 2006, Abelev et al. 2009b). Nazem et al. (2012) expressed Nc

as a function of Rigidity Index, G/su (G is the shear modulus) in addition to the

embedment ratio.

Numerous studies on free-falling penetrometers (e.g. Beard 1981, Levacher 1985,

Mulhearn et al. 1998, Bowman et al. 2005, Aubeny and Shi 2006, Abelev et al. 2009b,

Strak et al. 2012, Chow 2013) and dynamically installed anchoring systems (e.g.

O’Loughlin et al. 2004a, Audibert et al. 2006, Richardson et al. 2005, Richardson et al.

2006, Gilbert et al. 2008, Richardson 2008, O’Loughlin et al. 2009, Shelton et al. 2011,

Gaudin et al. 2013, O’Loughlin et al. 2013b) have adopted True’s method for predicting

the embedment of projectiles penetrating the seabed, with slight variations on the

inclusion and formulation of forces acting on the projectile during penetration. Many

studies have included the shear strain rate function, Rf (see Section 2.4.3.1). Added

mass (see Section 2.4.2) has been considered in a limited number of studies (e.g.

Fernandes et al. 2006, Kunitaki 2008). Aubeny and Shi (2006) and O’Loughlin et al.

(2013b) assumed that a cylindrical void forms in a projectile’s wake during dynamic

embedment into soil and accounted for this with the inclusion of a soil buoyancy term,

Fb (included as a negative quantity on the right hand side of Equation 2.17):

sb VF ' Equation 2.21

where γ' is the submerged weight of the soil and Vs is the volume of soil displaced by

the projectile. This assumption is supported by radiographs of centrifuge clay samples

by Poorooshasb and James (1989) showing open pathways in the wake of dynamically

installed cylindrical projectiles (despite closed entrance craters).

2.5.2 Numerical modelling

Einav et al. (2004) conducted finite difference numerical analyses to model dynamic

penetration of DPAs through the soil stratum. The finite difference model employed a

Page 91: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

57

rigid-deformable contact formulation in an explicit time-marching large strain

Lagrangean analysis and incorporated a separate equation of motion to solve the

incremental changes in velocity of the DPA (Einav et al. 2004). Both rate-dependent

and rate-independent models were developed. The rate dependent model adopted a

semi-logarithmic rate function (Equation 2.7). The influence of fluid drag resistance and

added mass were not considered in the model. The results of the numerical analyses

showed good agreement with experimental results from centrifuge tests on DPAs

(O’Loughlin et al. 2004b).

Raie et al. (2009) describes the development of a computational procedure to predict the

embedment depth of torpedo anchors. The procedure uses a CFD model for the

evaluation of the resistance forces acting on torpedo anchors during installation. The

soil is modelled as a non-Newtonian Bingham plastic fluid with a non-zero shear stress

at zero strain-rate. The model is capable of simulating the anchor motion during free-fall

in water and soil penetration. In addition the model also predicts the pressure and shear

distributions on the soil-anchor interface and in the soil. Good agreement was achieved

between the CFD model, results from laboratory tests of a FFP (True 1976) and field

tests of a full scale torpedo anchor (Medeiros 2002).

Strum et al. (2010) carried out large deformation finite element (LDFE) analyses to

simulate the penetration of 1:3 reduced scale DPA in clay. Anchor penetration was

modelled in a simplified manner by means of quasi-static, implicit and updated

Lagrangian analysis employing a finite-slip contact formulation along the soil anchor

interface. The LDFE model simulates the stress and excess pore water pressure

distribution during installation. The model results showed that there is an increase in

normalised mean stress and radial stress at the mid-height of the anchor following

installation which is about half the value given by cavity expansion theory for the same

Page 92: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

58

problem. Results indicated that a considerable zone of remoulded clay forms along the

anchor after installation.

The Arbitrary Lagrangian Eulerian (ALE) method of finite element analyses has been

used to model the dynamic penetration of FFPs into soil (Nazem et al. 2001, Abelev et

al. 2009b, Carter et al. 2010, Nazem et al. 2012, Carter et al. 2013). The ALE method

eliminates severe mesh distortion due to large deformations by relocating nodal points

on all material boundaries followed by a static analysis. Carter et al. (2013) adopted the

ALE method to simulate the penetration of a 1,000 kg torpedo anchor into soft seabed

sediments. The soil behaviour was represented by an elastoplastic Tresca material

model with an associative flow rule. Rate dependency of the soil was modelled using a

semi-logarithmic rate function. The influence of fluid drag resistance was not

considered in the model. The model results indicated that the forces acting on the

anchor due to soil resistance during penetration increase approximately linearly with

depth for a linear shear strength gradient. It was also observed that these forces vary

with time and hence the deceleration of the torpedo is not constant but position

dependent.

Hossain et al. (2013) conducted LDFE analyses using Coupled Eulerian-Lagrangian

(CEL) to simulate the dynamic installation of torpedo anchors. Two interesting aspects

of the soil flow mechanism during anchor installation were revealed: (i) downward soil

flow concentrating around the advancing anchor reduced gradually with reducing

penetration velocity and more rapidly with increasing number of anchor fins and anchor

projected area, and (ii) mobilisation of an end bearing mechanism at the base of the

anchor as well as the fins with the latter reduced significantly for shorter fins.

Page 93: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

59

2.6 Summary

The dynamic embedment of projectiles into the seabed has previously been considered

in studies of FFPs and dynamically installed anchoring systems. Three types of

dynamically installed anchor have been developed: the torpedo anchor, the Deep

Penetrating Anchor, DPA and the OMNI-Max anchor. Full scale dynamically installed

anchors have been installed in the Campos Basin off the coast of Brazil (torpedo

anchors), in the North Sea off the coasts of Norway (DPAs) and in the Gulf of Mexico

(OMNI-Max anchors). Extensive centrifuge and 1 g modelling has been conducted on

dynamically installed anchors. With the exception of a few limited studies there is little

published field data on dynamically installed anchors. Furthermore a wide range of drag

coefficients and strain rate parameter values have been reported for dynamically

installed anchors from physical model tests.

An analytical method often referred to as ‘True’s method’ is commonly adopted to

predict the dynamic embedment of projectiles in soil that considers Newton’s second

law of motion, the static forces resisting projectile embedment, rate effects, soil

remoulding and fluid drag resistance. However, there have been significant variations

on the inclusion and formulation of forces considered in True’s method.

Hence, there is a clear requirement for an experimental study to ascertain the drag

coefficient and strain rate parameter values that best represent DEPLA behaviour during

free-fall in water and dynamic embedment in soil. These values will facilitate

refinement and calibration of True’s method for prediction of DEPLA embedment,

which is a prerequisite for estimating anchor capacity.

Page 94: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

60

CHAPTER 3 CAPACITY OF PLATE ANCHORS

3.1 Introduction

The capacity of direct embedment plate anchors such as the DEPLA is primarily

dependent on: (i) the surrounding shear strength of the soil, acknowledging that the

embedment depth loss due to the keying process may reduce the relevant value (see

Section 3.2), and (ii) the bearing capacity factor of the plate (see Section 3.3). There

have been numerous analytical, numerical and laboratory studies on (i) and (ii).

However, there are very little reported field data on the performance of plate anchors to

verify the findings of these studies.

3.2 Keying

During the keying process the anchor embedment depth reduces as the plate rotates.

Since offshore clay deposits are typically characterised by an increase in shear strength

with depth, a reduction in anchor embedment depth will correspond to a non-

recoverable loss in potential anchor capacity (Randolph et al. 2005, Gaudin et al. 2009,

Cassidy et al. 2012). Naval Civil Engineering Laboratory guidelines (NCEL 1985)

states that the embedment depth loss, Δze,plate, due to the keying process is

approximately twice the anchor height, B, in cohesive soils, and is a function of anchor

geometry, soil type, soil sensitivity and duration of time between installation and

keying. However results from onshore and offshore SEPLA field tests reported by

Wilde et al. (2001) indicated embedment depth loss in the range Δze,plate = 0.5 to 1.7B.

The embedment depth loss of plate anchors during the keying process has been

investigated through centrifuge model tests (O’Loughlin et al. 2006, Gaudin et al. 2006,

Song et al. 2006, Song et al. 2007, Gaudin et al. 2009, Gaudin et al. 2010), LDFE,

analyses (Song et al. 2005, Song et al., 2008a, Wang et al. 2008a, Song et al. 2009, Yu

Page 95: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

61

et al. 2009, Wang et al. 2011, Wang et al. 2013a, Wang et al. 2013b), small strain finite

element analyses (Tian et al. 2013) and plastic limit analyses (Cassidy et al. 2012, Yang

et al. 2012). These studies assessed factors that may influence the keying behaviour of

plate anchors such as: installation method, padeye eccentricity ratio, pullout angle, soil

properties, anchor thickness ratio, anchor submerged unit weight, anchor aspect ratio,

anchor roughness, anchor keying flap and padeye offset. The author believes that the

most influential of these factors is padeye eccentricity ratio, and as such great

consideration should be given to this factor during design, in order to achieve optimal

anchor performance. The effect of anchor keying flap and padeye offset on keying

behaviour are not considered here as these features are incompatible with the DEPLA

geometry currently proposed.

3.2.1 Installation method

Centrifuge tests on SEPLAs in kaolin clay showed that embedment depth loss is lower

for suction embedded anchors (Δze,plate = 0.9 to 1.3B, where B is plate width) than for

jacked-in anchors (Δze,plate = 1.3 to 1.5B) due to the reduced amount of soil disturbance

during suction installation (Gaudin et al. 2006). Considerably lower embedment depth

loss was observed by Song et al. (2007) during centrifuge tests in ‘synthetic’ uniform

clay for both a suction embedded anchor (Δze,plate = 0.23B) and a jacked-in anchor

(Δze,plate = 0.27B). Song et al. (2007) proposed that this is due to the difference in

strength profiles of the kaolin clay and synthetic clay.

3.2.2 Padeye eccentricity ratio

Centrifuge tests (O’Loughlin et al. 2006) and LDFE analyses (Wang et al. 2008a, Song

et al. 2009, Yu et al. 2009, Wang et al., 2011) indicated that the embedment depth loss

of plate anchors during keying is primarily a function of the padeye eccentricity ratio,

e/B (Figure 3.1). These studies suggested that embedment depth loss may be minimised

Page 96: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

62

by ensuring a minimum padeye eccentricity ratios of e/B = 1 (O’Loughlin et al. 2006)

and e/B = 0.5 (Wang et al. 2008a, Yu et al. 2009, Wang et al. 2011).

Figure 3.1. Effect of padeye eccentricity ratio on keying behaviour

3.2.3 Pullout angle

Centrifuge tests (Gaudin et al. 2006, Song et al. 2006, Gaudin et al. 2009) and LDFE

analyses (Song et al. 2005, Song et al. 2008a, Wang et al. 2008a, Song et al. 2009, Yu et

al. 2009, Wang et al. 2011) indicated that embedment depth loss during keying

decreases with decreasing pullout angle, θpull (load inclination to the horizontal at the

padeye) (Figure 3.2), as the anchor is required to rotate less in order to achieve an

orientation that is a perpendicular to the direction of loading despite a reduction in

eccentricity.

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6

No

rmal

ised

em

bem

den

t d

epth

lo

ss, Δ

z e,p

late/B

Padeye eccentricity ratio, e/B

SEPLA field test (Wilde et al. 2001)

Centrifuge test, square, t/B =0.05 (Song et al. 2005)

Centrifuge test, square, t/B =0.067 (O'Loughlin et al. 2006)

Centrifuge test, strip, t/B =0.067 (O'Loughline et al. 2006)

Centrifuge test, square, t/B = 0.05 (Song et al. 2006)

LDFE, square, t/B = 0.1 (Song et al. 2009)

LDFE, square, t/B = 0.067 (Song et al. 2009)

LDFE, square, t/B = 0.05 (Song et al. 2009)

LDFE, square, t/B = 0.05 (Yu et al. 2009)

LDFE, strip, t/B = 0.05 (Yu et al. 2009)

LDFE, square, t/B = 0.067 (Wang et al. 2011)

LDFE, strip, t/B = 0.067 (Wang et al. 2011)

Page 97: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

63

Figure 3.2. Influence of pullout angle on keying behaviour

3.2.4 Soil properties

The influence of soil strength on embedment depth loss has been explored through

LDFE analyses (Song et al. 2008a, Wang et al. 2008a, Song et al. 2009, Wang et al.

2011). Results showed that the keying behaviour of deeply embedded plate anchors is

effectively dependent of soil rigidity ratio, E/su (Figure 3.3) (Wang et al. 2008a, Wang

et al. 2011), and soil shear strength profile (Figure 3.4) (Song et al. 2008a, Wang et al.

2008a, Song et al. 2009, Wang et al. 2011). However embedment depth loss increases

with increasing soil strength ratio, su0/γ'B, (Figure 3.5) as the anchor rotates more easily

in lower strength soils thus loses less embedment depth (Wang et al. 2008a, Song et al.

2009, Wang et al. 2011). Wang et al. (2011) suggested a limiting value of su0/γ'B =

0.075 at which point embedment depth loss becomes independent of su0/γ'B.

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

0 20 40 60 80 100 120

No

rmal

ised

em

bed

men

t dep

th lo

ss, Δ

z e,p

late/B

Pullout angle, θpull ( )

Centrifuge test, square, e/B = 0.625 (Song et al. 2005)

Centrifuge test, square, e/B = 0.66 (Gaudin et al. 2006)

Centrifuge test, strip, e/B = 0.17 (O'Loughlin et al. 2006)

Centrifuge test, strip, e/B = 1 (O'Loughlin et al. 2006)

LDFE, square, e/B = 0.625 (Song et al. 2006)

Centrifuge test, strip, e/B = 0.25 (Gaudin et al. 2009)

Centrifuge test, strip, e/B = 1 (Gaudin et al. 2009)

LDFE, strip, e/B =0.625 (Song et al.2009)

LDFE, strip, e/B = 0.66 (Yu et al. 2009)

LDFE, square, e/B= 0.66 (Yu et al. 2009)

Page 98: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

64

Figure 3.3. Effect of soil rigidity ratio, E/su, on keying behaviour (after Wang et al.

2011)

Pla

te r

ota

tion, θ

p(

)

Normalised embedment depth loss, Δze,plate/B

Pla

te r

ota

tion, θ

p(

)

Normalised embedment depth loss, Δze,plate/B

Page 99: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

65

Figure 3.4. Influence of soil shear strength profile on keying behaviour (after

Wang et al. 2011)

Pla

te r

ota

tion, θ

p(

)P

late

rota

tion, θ

p(

)P

late

rota

tion, θ

p(

)

Normalised embedment depth loss, Δze,plate/B

Normalised embedment depth loss, Δze,plate/B

Normalised embedment depth loss, Δze,plate/B

Page 100: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

66

(a)

(b)

Figure 3.5. Effect of soil strength ratio on keying behaviour (a) after Song et al.

(2009) and (b) after Wang et al. (2011)

3.2.5 Anchor thickness ratio and submerged unit weight

LDFE analyses (Song et al. 2009, Wang et al. 2011) showed that embedment depth loss

decreases with increasing anchor thickness ratio, t/B, and that anchor thickness has a

two-fold effect on embedment depth loss: (i) The anchor submerged unit weight, γ'a,

increases with thickness thus the pullout force has to balance the increased anchor

weight during initial rotation which generates higher moment to rotate the anchor at its

initial embedment depth, and (ii) The anchor end bearing resistance increases with

thickness which decreases vertical movement of the anchor. Therefore thicker anchors

rotate more easily and lose less embedment depth (Figure 3.6 and Figure 3.7) (Song et

al. 2009, Wang et al. 2011).

Pla

te r

ota

tion, θ

p(

)

Normalised embedment depth loss, Δze,plate/B

Norm

alis

ed e

mbedm

ent

depth

loss

, Δ

z e,p

late

/B

Page 101: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

67

(a)

(b)

Figure 3.6. Effect of anchor thickness ratio on keying behaviour: (a) after Song et

al. (2009) and (b) after Wang et al. (2011)

Figure 3.7. Influence of anchor submerged unit weight on keying behaviour (after

Song et al. 2009)

Norm

alis

ed e

mbedm

ent

depth

loss

, Δ

z e,p

late

/BΔ

z e,p

late

/Δz e

,ma

xP

late

rota

tion, θ

p(

)

Normalised embedment depth loss, Δze,plate/B

Page 102: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

68

3.2.6 Anchor aspect ratio

The effect of anchor aspect ratio, L/B, on embedment depth loss during keying was

explored through centrifuge tests (O’Loughlin et al. 2006) and LDFE analyses (Yu et al.

2009, Wang et al. 2011). O’Loughlin et al. (2006) observed no discernible difference in

embedment depth loss between square and strip anchors. However LDFE analyses

suggested that the embedment depth loss reduces with increasing L/B (Wang et al.

2008a, Yu et al. 2009, Wang et al. 2011).

3.2.7 Anchor roughness

Wang et al. (2013b) assessed the effect of anchor roughness (frictional resistance) on

embedment loss using LDFE analyses. Results of the study suggest that the effect of

anchor roughness on embedment depth loss is combined with the effects of loading

eccentricity ratio, e/B, and thickness ratio, t/B, and that the effect of anchor roughness

becomes dominant for e/B ≤ 0.5 or for t/B = 0.07 (Figure 3.8). LDFE analyses by Song

et al. (2009) also indicated that the effect of anchor roughness is dominant for e/B ≤ 0.5

(Figure 3.9).

Figure 3.8. Combined effects of anchor roughness, thickness ratio and eccentricity

ratio on keying behaviour (after Wang et al. 2013b)

(Δz e

,ma

xw

ith α

= 0

.3)/

(Δz e

,ma

xw

ith α

= 1

)

Page 103: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

69

Figure 3.9. Influence of anchor roughness on the keying behaviour (after Song et

al. 2009)

3.2.8 Anchor geometry coefficient

Song et al. (2009) showed that the normalised embedment depth loss, Δze,plate/B, for a

vertical pullout may be expressed in terms of a non-dimensional anchor geometry

coefficient of the form (Figure 3.10):

1.0

0

0

3.0

,

up

platee

BsA

M

B

t

B

ef

B

z Equation 3.1

where is the e is the padeye eccentricity, t is the anchor thickness, M0 is the initial

moment corresponding to zero net vertical load, Ap is the projected area of the anchor

perpendicular to the direction of load, B is the anchor height and su0 is the soil shear

strength at the initial anchor embedment depth. The anchor geometry coefficient

showed good agreement with centrifuge test results (Gaudin et al. 2009) and LDFE

analyses results using the expression (Song et al. 2009):

1.0

0

0

3.0

, 15.0

up

platee

BsA

M

B

t

B

eB

z Equation 3.2

Song et al. (2009) recommended an anchor geometry coefficient of at least 0.3 or

conservatively 0.4 to minimise the normalised embedment depth loss to Δze,plate/B ≤ 0.5.

Pla

te r

ota

tion, θ

p(

)

Normalised embedment depth loss, Δze,plate/B

Page 104: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

70

The anchor geometry coefficient expression was expanded by Gaudin et al. (2010) to

include the load inclination at the padeye commonly referred to as pullout angle, θpull

(Figure 3.10):

21.0

0

0

3.0

,

2sin

15.0

pulluppull

platee

BsA

M

B

t

B

eB

z

Equation 3.3

(a)

(b)

Figure 3.10. Embedment depth loss expressed in terms of a non-dimensional

anchor coefficient: (a) after Song et al. (2009) and (b) after Gaudin et al. (2010)

(e/B)×(t/B)0.3×(M0/ApBsu0)0.1

Norm

alis

ed e

mbedm

ent

depth

loss

, Δ

z e,p

late

/B

2

2

1.0

0

3.0

sin

u

BspAoM

B

t

B

e

Norm

alis

ed e

mbedm

ent

depth

loss

, Δ

z e,p

late

/B

21.0

0

03.0

2sin

pupp BsA

M

B

t

B

e

Equation 3.3

Page 105: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

71

3.3 Capacity

The ultimate vertical pullout capacity of plate anchors is typically expressed in terms of

a bearing capacity factor or ‘break out’ factor (Das and Singh 1994, Merifield et al.

2003, Gaudin et al. 2006, Song et al. 2008b, Wang et al. 2010):

pup

netv

csA

FN

,

, Equation 3.4

where Fv,net is the ultimate vertical pullout capacity of plate anchors minus the

submerged weight of the anchor in soil, su,p is the undrained shear strength at the anchor

embedment depth (anchor mid-depth) corresponding with the ultimate capacity and Ap

is the projected area of the plate.

3.3.1 Failure mechanism

Plate anchors can be classified as ‘shallow’ or ‘deep’ depending on the failure

mechanism. An anchor is classified as shallow if the failure mechanism reaches the soil

surface at ultimate collapse or deep if the failure mechanism is localised shear around

the anchor and does not extend to the soil surface (see Figure 3.11). At the critical

embedment depth, zcr, the failure mechanism becomes localised and the bearing

capacity factor of the anchor no longer increases with embedment depth (Figure 3.11,

Merifield et al. 2003, Thorne et al. 2004, Song et al. 2006, Gaudin et al. 2006), as the

undrained shear strength is assumed to be independent of the mean normal stress

(Merifield et al. 2001, 2003, 2005).

Page 106: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

72

Figure 3.11. Shallow and deep anchor behaviour

3.3.2 Suction

Negative pore water pressure will usually develop below a plate anchor during loading

due to the low permeability of clays coupled with a high loading rate. This generates a

tensile force known as suction on the back face of the plate. Suction increases with the

differential pressure on the top and bottom faces of the plate (Shin et al. 1994). The

suction force developed between the anchor and soil is a function of the embedment

depth, soil permeability, undrained shear strength and loading rate (Merifield et al.

2001, Thorne et al. 2004, Song et al. 2008b). Following Rowe and Davis (1982) most

studies into the capacity of plate anchors in clay consider two distinct cases:

‘breakaway’ or ‘no breakaway’. For the breakaway or ‘vented’ case it is assumed that

interface between the soil and the back face of the plate cannot sustain sufficient suction

and either the soil separates from the anchor at the start of pullout (immediate

breakaway) if no initial stresses are considered or separates during pullout if initial

compressive stresses are considered. For the no breakaway or ‘attached’ case it is

z1z2

zcr

Nc1 Nc2 Nc

Nc

Nc

Nc2

Nc1

(d)

(a)

(b)

(c)

‘Deep’

B

B

Page 107: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

73

assumed that the interface between the soil and the back face of the plate can sustain

sufficient suction and the soil remains attached to the back of the plate throughout

pullout. The depth at which the anchor becomes vented is known as the ‘separation

depth’, zs.

3.3.3 Previous studies

A number of studies have proposed simple approximate methods to estimate the bearing

capacity factors of plate anchors in clay (Adam and Hayes 1967, Ali 1968, Meyerhof

and Adams 1968, Vesic 1971, Meyerhof 1973, David and Sunderland 1977,

Kupferman, 1971, Das 1978, Das 1980, Ranjan and Arora 1980, Das 1985a, Das 1985b,

Das and Puri 1989, Das et al. 1994). The majority of these studies estimate anchor

capacity through simple analytical solutions or empirical correlations based on

laboratory model test conducted under normal gravity (1 g). However small-scale 1 g

laboratory models cannot replicate the effect of in situ overburden pressure.

Rowe and Davis (1982) carried out small strain elasto-plastic finite element analyses to

estimate the bearing capacity factors of strip anchors in homogeneous clay. However no

ultimate capacity was achieved; instead, truncated capacities were derived using elastic

reasoning. The lower and upper bound bearing capacity factors for the no breakaway

condition were found to be Nc = 10.28 and 11.42 respectively.

Yu (2000) derived an expression for Nc based on accurate analytical solutions for cavity

expansion in cohesive-frictional soil. It was assumed that breakout occurs if the

boundary of the plastic zone predicted by cavity expansion theory was sufficiently close

to or on the ground surface (when the plastic flow was not confined by the outer elastic

zone).

Martin and Randolph (2001) carried out plasticity limit analysis of deep ultra-thin

circular plates in clay assuming no breakaway which produced exact solutions for

Page 108: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

74

bearing capacity factors of Nc = 13.11 and 12.42 for fully rough and smooth plates

respectively.

The bearing capacity factors of strip, circular and square anchors in homogeneous and

nonhomogeneous soils assuming no breakaway condition were investigated by

Merifield et al. (2001, 2003, 2005) through finite element formulations of limit analyses

based on rigid plastic soil response. Limiting bearing capacity factors for rough strip

anchors were found to be Nc = 10.47 (lower bound) and 11.16 (upper bound) with lower

bound values for rough circular and square anchors reported as Nc = 12.56 and 11.9

respectively. Results suggest that the Nc increases linearly with overburden pressure up

to a limiting value that corresponds with the transition from shallow to deep anchor

behaviour and that this value can be expressed as a function of overburden stress ratio,

γz/su (Figure 3.12).

Song et al. (2005) investigated the bearing capacity factors of strip anchors in uniform

clay through LDFE. For the attached condition the anchor reached a limiting bearing

capacity factor of Nc = 12 at a critical embedment ratio of zcr/B =3. No limiting bearing

capacity was reached for the vented condition even for zcr/B = 10 albeit in weightless

soil. The influence of plate angle, θp, was only found to be significant for embedment

ratio z/B ≤ 3 where Nc decreased with increasing pullout angle (measured relative to the

horizontal) (Figure 3.13). In Figure 3.12 and Figure 3.13 the point at which Nc starts to

reduce represents zcr.

Page 109: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

75

(a) Rough strip anchor – lower bound (after Merifield et al. 2001)

(b) Rough circular anchor – lower bound (after Merifield et al. 2003)

(c) Rough square anchor – lower bound (after Merifield et al. 2003)

Figure 3.12. Effect of overburden stress ratio on bearing capacity factor

γza/cu

zza

z/B = 5

z/B = 10

γz/cu

z

z/D = 1, 2, 3, 5

γz/cu

z

z/B = 5

z/B = 1

Page 110: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

76

(a) Attached plate

(b) Vented plate

Figure 3.13. Effect of pullout angle on bearing capacity factor (after Song et al.

2005)

Centrifuge tests on square SEPLAs in kaolin clay Gaudin et al. (2006) reported bearing

capacity factors in the range Nc = 12.3 to 13.5 for jacked-in anchors and Nc = 10.9 to

12.9 for suction installed anchors.

Song et al. (2008a) carried out small strain finite element analyses to assess the effect of

load inclination on the bearing capacity factor of a strip anchor in uniform and normally

consolidated clays. For the attached condition in normally consolidated clay the anchor

reached a limiting Nc = 11.6 at a normalised embedment depth, z/B = 6 (Figure 3.14a).

No limiting bearing capacity was reached for the vented condition even for z/B = 10

z/B

θp = 90 , 67.5 , 45 , 22.5 , 0

θp = 0 , 22.5 , 45 , 67.5 , 90

z/B

Page 111: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

77

(see Figure 3.14b). The soil strength gradient, k, was found to have no effect on Nc (see

Figure 3.15). Results indicated that su/γ'B influences Nc (Figure 3.16).

(a) Attached plate

(b) Vented plate

Figure 3.14. Effect of pullout angle on bearing capacity factor (after Song et al.

2008a)

Figure 3.15. Effect of soil strength gradient on bearing capacity factor (after Song

et al. 2008a)

z/B

NC, k = 1, θp = 90

NC, k = 1, θp = 45

NC, k = 1, θp = 0

Uniform, θp = 90

Uniform, θp = 45

Uniform, θp = 0

z/B

NC, k = 1, θp = 90

NC, k = 1, θp = 45

NC, k = 1, θp = 0

Uniform, θp = 90

Uniform, θp = 45

Uniform, θp = 0

z/B

k = 2, θp = 90

k = 2, θp = 45

k = 2, θp = 0

k = 1, θp = 90

k = 1, θp = 45

k = 1, θp = 0

Page 112: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

78

Figure 3.16. Influence of normalised strength ratio on bearing capacity factor

(after Song et al. 2008a)

Song et al. (2008b) investigated the bearing capacity factors of strip and circular plates

in uniform and normally consolidated through small strain and large strain deformation

finite element analyses. The bearing capacity factors assuming no breakaway condition

were found to be Nc = 11.6 (smooth) and 11.7 (rough) for deep strip anchors and Nc =

13.1 (smooth) and 13.7 (rough) for deep circular anchors. When the breakaway

condition was considered the soil remained attached to the back face of the plate for

deep anchors and the bearing capacity was the same as for the no breakaway condition

(Figure 3.17). The separation depth was found to be dependent on su/γ'B (Figure 3.18).

Wang et al. (2008b) showed through LDFE analyses that the bearing capacity of strip

anchors in normally consolidated clay decreases with increased loading rate for a given

initial embedment depth (Figure 3.19), until the loading rate approaches a critical value

which represents the undrained condition.

Yu et al. (2009) conducted 3-D LDFE analyses to assess the bearing capacity factor of

strip and square anchors in normally consolidated and uniform clay assuming the

attached condition. Results of the study showed that Nc increased with increased e/B

(Figure 3.20) and that θpull has minimal effect on Nc (Figure 3.21).

Uniform, su/γ'B = 0.89, θp = 90

Uniform, su/γ'B = 0.89, θp = 45

Uniform, su/γ'B = 0.89, θp = 0

Uniform, su/γ'B = 0.36, θp = 90

Uniform, su/γ'B = 0.36, θp = 45

Uniform, su/γ'B = 0.36, θp = 0

k = 1, θp = 90

k = 1, θp = 45

k = 1, θp = 0

z/B

Page 113: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

79

Figure 3.17. Bearing capacity factors and separation depth for continuous pullout

of rough circular plate anchors (after Song et al. 2008b)

Figure 3.18. Influence of normalised strength ratio on separation depth (Song et al.

2008b)

z/D

zcr/D = 0.3

zcr/D = 0.55

ze,p,initial/D = 0.5(A), su/γ'D = 0.065

ze,p,initial/D = 0.5(V), su/γ'D = 0.065

ze,p,initial/D = 1(A), su/γ'D = 0.065

ze,p,initial/D = 1(V), su/γ'D = 0.065

ze,p,initial/D = 2(A), su/γ'D = 0.065

ze,p,initial/D = 2(V), su/γ'D = 0.065

ze,p,initial/D = 1(V), su/γ'D = 0.039

zcr/D

su0/γ'D

Page 114: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

80

Figure 3.19. Influence of loading rate on bearing capacity factor (after Wang et al.

2008)

Nc

Nc

Nc

Nc

z = 5B

z = 3B

z = 2B

z = 1B

Page 115: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

81

Figure 3.20. Influence of padeye eccentricity ratio on bearing capacity factor (after

Yu et al. 2009)

Figure 3.21. Bearing capacity factor for various pullout angles (after Yu et al.

2009)

Wang et al. (2010) assessed the bearing capacity factors of circular, square, rectangular

and strip anchors in uniform clay through 3-D LDFE analyses. The bearing capacity for

a circular plate was found to be Nc = 14.33 achieved at z/B = 2 (Figure 3.22). Results

suggested that Nc increases with E/su, for square and circular anchors under immediate

breakaway conditions but that E/su has little effect on strip anchors (Figure 3.23).

Anchor roughness was shown to have negligible effect on bearing capacity factor.

Results indicated that the soil separation depth increases linearly with su/γ'B and is

independent of initial anchor embedment depth (Figure 3.24). Nc for rectangular

Δze,plate/B

dv/B

θpull = 45

θpull = 60

θpull = 90

Page 116: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

82

anchors was demonstrated to increase with increasing anchor aspect ratio, L/B, (Figure

3.25) which is in agreement with observations from soil model tests (Das 1978, 1980)

and lower bound analyses (Merifield et al. 2003).

Figure 3.22. Bearing capacity factors of plate anchors in uniform clay (after Wang

et al. 2010)

Figure 3.23. Effect of soil rigidity ratio on the bearing capacity factor (after Wang

et al. 2010)

z/B or z/D

Page 117: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

83

Figure 3.24. Dependence of separation depth on soil strength ratio (after Wang et

al. 2010)

Figure 3.25. Dependence of bearing capacity factor on anchor aspect ratio (after

Wang et al. 2010)

Yu (2011) carried out finite element analyses of strip anchors in non-uniform and

normally consolidated clay. For the attached condition the bearing capacity factor was

found to be Nc = 11.59 in uniform clay and Nc = 11.01 in normally consolidated clay

For attached anchors, or when γz/su, was large enough, Nc was observed to increase

with increasing pullout angle (Figure 3.26). Results showed that when γz/su was high Nc

for vented anchors was the same as those of attached anchors. The degree of soil non-

homogeneity factor, kB/su0, was found to have an approximately linear negative effect

zs/B

su/γ'B

z/B = 2

z/B = 1

z/B = 0.5

Page 118: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

84

on Nc (see Figure 3.27). Soil roughness was shown to have little effected on the Nc of

deeply embedded anchors (see Figure 3.28).

(a)

(b)

Figure 3.26. Bearing capacity factors of strip anchors in (a) uniform clay and (b)

normally consolidated clay (after Yu et al. 2011)

θp = 0 (horizontal), 22.5 , 45 , 67.5 , 90 (vertical)

z/B

Nc

Nc,max = 11.59

z/B

θp = 0 (horizontal), 22.5 , 45 , 67.5 , 90 (vertical)

Nc = 11.01 for z/B = 10, θp = 90

Nc

Nc,,ax = 11.59, θp = 0

Page 119: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

85

Figure 3.27. Effect of soil non-homogeneity on the ultimate bearing capacity of

strip anchors (after Yu et al .2011)

(a)

(b)

Figure 3.28. Influence of soil roughness on bearing capacity (after Yu et al. 2011)

z/B = 1

z/B = 1.5

z/B = 2z/B = 3~10

z

z/B

su = 0.1 + 3z kPa, θp = 90

su = 0.1 + 3z kPa, θp = 0

Uniform clay, θp = 90

Uniform clay, θp = 0

z/B

su = 0.1 + 3z kPa, θp = 90

su = 0.1 + 3z kPa, θp = 0

Uniform clay, θp = 90

Uniform clay, θp = 0

Page 120: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

86

Yang et al. (2012) conducted plastic limit analyses to examine the effect of the keying

flap on the capacity of rectangular SEPLAs in normally consolidated clay. The

theoretical bearing capacity factor achieved by a SEPLA with a flap (Nc = 12.79) was

higher than a SEPLA without a flap (Nc = 11.67) as the flap increases the bearing area.

Cassidy et al. (2012) implemented plastic limit analyses to predict the bearing capacity

factors of plate anchors using a plasticity model termed ‘chain and SEPLA plasticity

analysis’ (CASPA). The CASPA results showed good agreement with comparable

LDFE analyses and centrifuge tests. The study investigated the optimal vertical position

of the padeye to maximise Nc. Results indicated that the padeye offset results in two

opposing phenomenon: (i) A gain in capacity resulting from a reduced loss of

embedment (ii) A reduction in capacity due to the inclination of the anchor in respect of

the pullout direction. Cassidy et al. (2012) concluded that the extent to which the gain

resulting from the first aspect compensates the loss from the second aspect is a function

of the shear strength gradient and the initial embedment depth.

Zhang et al. (2012) carried out a parametric study to assess the undrained bearing

capacity of deeply embedded flat circular footings under general loading. The study

showed that Nc significantly increases with thickness ratio, with Nc for t/D = 1 being

58% greater than for t/D =0.05 (Figure 3.29). Zhang et al. (2012) suggests that this trend

is due to the increased friction along the vertical footing-soil interface as t/D increases

and that the increased t/D deters the soil from flowing around from below the footing

through the footing edge to the top forcing the soil to take a deeper and longer route to

flow around thus mobilising more soil in the mechanism which increases foundation

capacity.

Page 121: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

87

Figure 3.29. Influence of anchor thickness ratio on the bearing capacity factor of

deeply embedded circular footings (Zhang et al. 2012)

Chen et al. (2013) and Tho et al. (2013) investigated the bearing capacity factors of

square plate anchors through 3-D LDFE analyses. Results demonstrated that the bearing

capacity factor increases with γz/su, up to a limiting factor Nc = 13.1 in both uniform

and normally consolidated clay (Figure 3.30). It was found that Nc is independent of

kB/su0 (Figure 3.31) (Tho et al. 2013) and E/su0 (Figure 3.32) (Chen et al. 2013). Chen

et al. (2013) also showed that Nc increases with loading rate (Figure 3.33).

Figure 3.30. Bearing capacity factors as a function overburden pressure ratio

(after Chen et al. 2013)

Thickness ratio, t/D

z/B = 7

z/B = 6

z/B = 5

z/B = 4

z/B = 3

z/B = 2

z/B = 1

γz/su

Page 122: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

88

(a) kB/su,0 = 0

(b) kB/su,0 = 1

(c) kB/su,0 = ∞

Figure 3.31. Effect of soil non-homogeneity factor on bearing capacity factor (after

Tho et al. 2013)

z/B = 7

z/B = 6

z/B = 5

z/B = 4

z/B = 3

z/B = 2

z/B = 1

z/B = 7

z/B = 6

z/B = 5

z/B = 4

z/B = 3

z/B = 2

z/B = 1

γz/(su,0+kz)

z/B = 10

z/B = 9

z/B = 8

z/B = 7

z/B = 6

z/B = 5

z/B = 4

z/B = 3

z/B = 2

z/B = 1

Page 123: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

89

Figure 3.32. Effect of soil rigidity ratio on bearing capacity (after Chen et al. 2013)

(a) z/B = 5, E/su = 2000

(a) z/B = 5, E/su = 10000

Figure 3.33. Effect of loading rate on the bearing capacity factor of square anchors

(after Chen et al. 2013)

γz/su

z/B = 7, E/su = 500

z/B = 6, E/su = 500

z/B = 5, E/su = 500

z/B = 4, E/su = 500

z/B = 3, E/su = 500

z/B = 2, E/su = 500

z/B = 1, E/su = 500

Page 124: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

90

3.4 DEPLA capacity

Wang and O’Loughlin (2014) assessed the bearing capacity factors of circular,

diamond, square and rectangular DEPLAs in normally consolidated clay through 3-D

LDFE. Results of the numerical analyses were validated by benchmarking with existing

numerical and analytical solution for circular and rectangular plate anchors. The effect

of anchor embedment depth, anchor roughness, fluke thickness, plate inclination and

anchor geometry on Nc were investigated assuming the attached condition. Results

indicated that the limiting bearing capacity factor associated with deep behaviour

occurred at a normalised embedment depth, z/D = 2.5 for circular, diamond, square and

rectangular DEPLAs under the attached condition and subject to vertical load (Figure

3.34). The limiting bearing capacity factors were found to be Nc =14.9 for square and

circular DEPLAs, Nc = 15 for a diamond geometry and Nc = 15.4 to 13.6 for rectangular

DEPLAs with L/B = 0.5 and 2 respectively. The Nc for the rectangular DEPLA with

L/B = 0.5 was attributed to the geometrical difference related to the direction of the

sleeve and the relative height of the plate. Plate roughness (Figure 3.35) and fluke

thickness (Figure 3.36) were found to have a relatively small effect on Nc. Result

showed that plate inclination has a considerable effect on Nc (Figure 3.37), particularly

for shallow embedment ratios. Nc for a circular DEPLA at z/D = 1 (shallow) reduced by

23.4% as the plate inclination changed from vertical to horizontal, compared with a

1.3% reduction for z/D = 4 (deep). Analyses were also conducted where soil separation

was permitted, the results of which suggested that the bearing capacity factors of

DEPLAs approach the no breakaway values when z/D increases and su (relative to the

unit weight of the soil) decreases (Figure 3.38).

Page 125: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

91

Figure 3.34. Nc of DEPLAs as a function of embedment ratio

Figure 3.35. Influence of anchor roughness on NC of DEPLAs

Figure 3.36. Effect of fluke thickness on NC of DEPLAs

z/D or z/B

z/D

z/D

Page 126: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

92

Figure 3.37. Effect of plate inclination on Nc of DEPLAs

Figure 3.38. Effect of soil seperation on Nc of DEPLAs

3.5 Summary

Numerous laboratory, analytical and numerical studies have been conducted on directly

embedded plate anchors to investigate embedment depth loss due to the keying process

and bearing capacity factor of the plate. A wide range of embedment depth loss and

bearing capacity factors have been reported. With the exception of a study on SEPLAs

by Wilde et al. (2001) there is currently no published field data for vertical load

anchors. Hence, there is a clear requirement for an experimental study to provide the

first DEPLA performance data, and determine the keying behaviour and capacity of

these anchors, along with enhancing the currently limited field database for directly

embedded plate anchors.

θp

θp= 0 θp= 30 θp= 45 θp= 60 θp= 90

z/D

Breakaway, γ'z/su = 4.5

Page 127: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

93

CHAPTER 4 INSTALLATION AND CAPACITY OF

DYNAMICALLY EMBEDDED PLATE

ANCHORS AS ASSESSED THROUGH

CENTRIFUGE TESTS

4.1 Abstract

The dynamically embedded plate anchor (DEPLA) is a rocket or dart shaped anchor that

comprises a removable central shaft and a set of four flukes. Similar to other

dynamically installed anchors, the DEPLA penetrates to a target depth in soft seabed

sediments by the kinetic energy obtained through free-fall in water and the self-weight

of the anchor. In this chapter DEPLA performance was assessed through a series of

beam centrifuge tests conducted at 200 times earth’s gravity. The results show that the

DEPLA exhibits similar behaviour to other dynamically installed anchors during

installation, with tip embedments of 1.6 to 2.8 times the anchor length. After anchor

installation the central shaft of the DEPLA, termed a follower, is retrieved and reused

for the next installation, leaving the DEPLA flukes vertically embedded in the soil. The

load-displacement response during follower retrieval is of interest, with mobilisation of

frictional and bearing resistance occurring at different rates. The load required to extract

the DEPLA follower is typically less than three times its dry weight. The vertically

embedded DEPLA flukes constitute the load bearing element as a circular or square

plate. The keying and pullout response of this anchor plate is similar to other vertically

embedded plate anchors, with an initial stiff response as the anchor begins to rotate,

followed by a softer response as the rotation angle increases, and a final stiff response as

the effective eccentricity of the padeye reduces and anchor capacity is fully mobilised.

For the padeye eccentricity ratios considered (0.38 to 0.63 times the plate breadth or

diameter), the loss in plate anchor embedment is between 0.50 and 0.66 times the

Page 128: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

94

corresponding plate breadth or diameter. Finally, the bearing capacity factors

determined experimentally are typically in the range 14.2 to 15.8 and are higher than

numerical solutions for flat circular and square plates. This is considered to be due to

the cruciform fluke arrangement which ensures that the failure surface extends to the

edge of the orthogonal flukes and mobilises more soil in the failure mechanism.

4.2 Introduction

Floating oil and gas installations in deep water require anchoring systems that can

withstand high components of vertical load, whilst being economical to install. Deep

water installations are typically moored using either (drag embedded) vertically loaded

plate anchors (VLAs) or suction caissons, although a number of other anchoring

systems such as suction embedded plate anchors and dynamically installed anchors

(Brandão et al. 2006, Zimmerman et al. 2009) have been used with effect in recent

years. Predicting plate anchor capacity for a given embedment is generally

straightforward as it is a function of the local undrained shear strength, the projected

area of the plate and a dimensionless bearing capacity factor, which for deeply

embedded plates is typically in the range 11.7 to 14.3 depending on plate geometry and

roughness (Merifield et al. 2003, Song et al. 2005, Song et al. 2008b). However

predicting the anchor installation trajectory and hence final embedment depth is more

challenging and requires accurate shear strength data over a wide seabed footprint

(Ehlers et al. 2004). Suction caissons are advantageous in this regard as embedment is

monitored and controlled, although successful prediction of the ultimate holding

capacity must account for a number of factors including the caisson geometry and

padeye location, loading angle, time effects and the integrity of the internal seal

provided by the soil plug (Randolph et al. 2011). Several of the disadvantages of these

two anchor types have been mitigated by the suction embedded plate anchor (SEPLA)

described by Wilde et al. (2001). The SEPLA employs a suction caisson to install a

Page 129: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

95

vertically oriented plate anchor. After installation the follower is removed (and reused)

and the embedded plate is rotated or keyed so that it becomes normal or near normal to

the load applied by the mooring chain.

Anchor installation operations become increasingly more complex, time-consuming and

hence costly as water depth increases. However, this is not the case with dynamically

installed anchors as no external energy source or mechanical operation is required

during installation. Dynamically installed anchors are rocket or torpedo shaped and are

designed so that, after release from a designated height above the seafloor, they will

penetrate to a target depth in the seabed by the kinetic energy gained during free-fall.

Centrifuge model tests indicate that in normally consolidated clay, expected penetration

depths are 2 to 3 times the anchor length and expected anchoring capacities are 3 to 5

times the anchor dry weight (O’Loughlin et al. 2004b). In contrast, expected anchoring

capacities for vertically loaded plate anchors are typically 30 to 40 times the anchor dry

weight in very soft clay.

A new anchoring system, termed the dynamically embedded plate anchor (DEPLA)

combines the advantages of dynamically installed anchors and vertically loaded anchors

in much the same way as the SEPLA combines the advantages of suction caissons with

VLAs. The DEPLA is a rocket or dart shaped anchor that comprises a removable central

shaft or ‘follower’ that may be fully or partially solid and a set of four flukes (see Figure

1.17). A stop cap at the upper end of the follower prevents it from falling through the

DEPLA sleeve and a shear pin connects the flukes to the follower. As with other

dynamically installed anchors the DEPLA penetrates to a target depth in the seabed by

the kinetic energy obtained through free-fall. After embedment the follower line is

tensioned, which causes the shear pin to part allowing the follower to be retrieved for

the next installation, whilst leaving the anchor flukes vertically embedded in the seabed.

Page 130: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

96

The embedded anchor flukes constitute the load bearing element as a plate anchor. A

mooring line attached to the embedded flukes is then tensioned, causing the flukes to

rotate or ‘key’ to an orientation that is normal or near normal to the direction of loading.

In this way, the maximum projected area is presented to the direction of loading,

ensuring that maximum anchor capacity is achievable through bearing resistance. The

installation and keying processes are shown schematically on Figure 1.18 and Figure

1.19 respectively.

This chapter provides DEPLA performance data through a series of centrifuge tests that

were designed to quantify expected embedment depths and plate anchor capacities.

4.3 Centrifuge testing program

The centrifuge tests were carried out at 200 g using the fixed beam centrifuge at UWA.

The UWA beam centrifuge is a 1.8 m radius Acutronic centrifuge with a maximum

payload of 200 kg at 200 g (Randolph et al. 1991). All parameters presented in this

chapter are expressed in model scale. Table 4.1 presents the scaling laws required for

conversion from model scale to prototype scale.

Table 4.1. Centrifuge scaling laws (after Scholfield 1980, Taylor 1995)

Parameter Scaling relationship

(model/prototype)

Acceleration ncent

Length 1/ncent

Area 1/ncent2

Volume 1/ncent3

Mass 1/ncent3

Stress 1

Strain 1

Force 1/ncent2

Velocity 1

Density 1

Time (consolidation) 1/ncent2

Page 131: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

97

4.3.1 Model anchors

The model DEPLAs used in the centrifuge tests comprised a set of DEPLA flukes

(which form the plate anchor after the follower is retrieved), a DEPLA follower and

anchor lines connected to the follower and plate anchor padeyes. An interchangeable

modular design for the follower (ellipsoidal tip, shaft, padeye) allowed various steel and

aluminium sections to be combined to form different overall follower lengths and

masses, although the follower diameter and ellipsoidal tip length remained constant at

Df = 6.0 mm and Ltip = 11.4 mm (Figure 4.1). The overall length of the follower (and

hence height of the anchor) used in the centrifuge tests was either 51.5 mm, 76.0 mm or

101.5 mm, which at 200 g corresponds to equivalent prototype heights of 10.3 m, 15.2

m and 20.3 m respectively.

The flukes were fabricated from 0.8 mm thick steel and the sleeve from 0.75 mm thick

copper (rather than steel) to permit soldering of the flukes to the sleeve without

inducing high temperatures. The outer diameter of the sleeve is 7.8 mm, which, together

with the sleeve wall thickness of 0.75 mm, gave a nominal clearance of 0.15 mm

between the 6.0 mm diameter follower and the sleeve. The DEPLA flukes that formed

the plate anchor element were either circular or square, with the latter oriented in a

diamond shape on the DEPLA sleeve (Figure 4.1), so as to reduce drag and to increase

the eccentricity of the anchor padeye (load attachment point). The diameter, D, or

breadth, B, of the DEPLA flukes varied between 16 mm and 37.6 mm, which

correspond to 3.2 m to 7.5 m in equivalent prototype scale. The plate anchor padeye

was located at an eccentricity, e, from the central axis of the plate, such that the

eccentricity ratio was in the range e/D = 0.38 to 0.46 for the circular plates and e/B =

0.63 for the square plate. Details of the six model anchors employed for the centrifuge

tests are summarised in Table 4.2 and shown on Figure 4.1.

Page 132: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

98

Figure 4.1. 1:200 reduced scale DEPLAs

Both the follower and plate anchor lines were modelled using uncoated stainless steel

wire of 0.90 mm diameter (180 mm in prototype scale) and a tensile capacity of 660 N.

This line was selected due to its flexibility, high tensile strength and resistance to

stretching. The follower line was attached to the padeye located at the top of the

follower, whereas the plate anchor line was connected to the plate anchor padeye

located on one of the anchor flukes. The other extremity of each anchor line was

connected via a 1.7 kN load cell to the actuator that effected the follower recovery and

the plate anchor pullout.

Table 4.2. Model anchors

Ref. Flukes Lf

(mm)

Lf/Df Hs

(mm)

D or

B

(mm)

D/Lf

or

B/Lf

e

(mm)

e/B

or

e/D

Mass

(g)

Follo-

-wer

Flukes

(plate)

Total

A1 Circular 76.00 12.67 28.97 30.00 0.39 13.50 0.45 10.46 11.00 21.46

A2 Circular 76.00 12.67 21.21 22.60 0.35 9.80 0.43 10.46 6.09 16.55

A3 Circular 76.00 12.67 36.78 37.60 0.49 17.30 0.46 10.46 16.12 26.58

A4 Circular 51.50 8.58 12.81 16.00 0.31 6.00 0.38 9.29 2.91 12.20

A5 Circular 101.5 16.92 28.97 30.00 0.30 13.50 0.45 12.04 11.00 23.04

A6 Square 76.00 12.67 30.00 26.73 0.35 16.78 0.63 10.46 10.75 21.21

Page 133: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

99

4.3.2 Soil properties

Normally consolidated clay samples were prepared by mixing commercially available

kaolin clay powder with water to form a slurry at 120 % (i.e. twice the liquid limit, see

Table 4.3) and subsequently de-aired under a vacuum close to 100 kPa. The clay slurry

was carefully transferred into a rectangular strongbox, with internal dimensions 650 mm

long, 390 mm wide and 325 mm deep, before self-weight consolidation in the centrifuge

at 200 g for approximately four days to create a normally consolidated clay sample. A

10 mm sand layer at the base of the clay permitted drainage towards the base of the

sample and a 5 to 10 mm layer of water was maintained at the sample surface to ensure

saturation. Additional clay slurry was added to the sample during consolidation to

achieve the required sample height of 230 mm. Pore pressure transducers placed at

various depths within the sample allowed for the measurement of excess pore water

pressures and these, together with T-bar penetration tests, served to indicate the degree

of consolidation of the sample.

Table 4.3. Kaolin clay properties (after Stewart 1992)

Liquid limit, LL (%) 61

Plastic limit, PL (%) 27

Specific gravity, Gs 2.6

Angle of internal friction, ' (°) 23

Effective soil unit weight, γ' (kN/m3) 6.5

Voids ratio at p' = 1 kPa on critical state line, ecs 2.14

Slope of normal consolidation line, ς 0.205

Slope of swelling line, κ 0.044

Coefficient of vertical consolidation, cv (m2/yr)

(at OCR = 1 and σ'v = 150 kPa)

4.5

Soil characterisation tests were performed using a T-bar penetrometer (Stewart and

Randolph 1991) from which a continuous profile of the undrained shear strength, su,

Page 134: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

100

was derived using the commonly adopted T-bar factor, NT-bar = 10.5. The T-bar tests

were conducted at a velocity of 1 mm/s, which corresponds to a dimensionless velocity,

V = vdT-bar/cv in the range 32 to 113 (where v is the velocity, dT-bar is the T-bar diameter

and cv is the vertical coefficient of consolidation, varying from 1.4 to 5 m2/yr for σ'v in

the range 10 to 300 kPa), thus ensuring undrained conditions (House et al. 2001).

T-bar penetration tests were conducted during the latter stages of consolidation and

before, during and after the anchor tests to observe any changes in strength with time.

Typical shear strength profiles from Samples 1 and 2 conducted before and after testing

are provided on Figure 4.2. Centrifuge strength data tend to deviate from the expected

linear profile for normally consolidated clays in the lower third of the sample. This

effect is due to the slight increase in radial acceleration level and effective unit weight

with increasing sample depth. The variation in radial acceleration and effective unit

weight has therefore been accounted for in the theoretical strength profile on Figure 4.2,

which was obtained using an undrained shear strength ratio, su/σ'v = 0.22. Although the

theoretical and measured profiles deviate beyond 140 mm, the theoretical profile

provides a good approximation of the strength over the range of interest for the DEPLA

plate (up to 103 mm).

Figure 4.2. Undrained shear strength profiles

0 20 40 60 80 100

200

180

160

140

120

100

80

60

40

20

0

Undrained shear strength, su (kPa)

Pen

etra

tio

n d

epth

, z

(mm

)

Sample 1, T-bar 1

Sample 1, T-bar 2

Sample 1, T-bar 3

Sample 2, T-bar 1

Sample 2, T-bar 2

Sample 2, T-bar 3

su/

v' = 0.22

Page 135: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

101

4.3.3 Experimental arrangement and testing procedures

The model DEPLAs were installed in-flight by allowing them to ‘free-fall’ through the

elevated acceleration field from a drop height of 300 mm. This approach has been

adopted for the installation of dynamically installed ‘torpedo piles’ in centrifuge tests

(Richardson et al. 2009, O’Loughlin et al. 2013b) and has resulted in anchor impact

velocities of the order of 30 m/s at 200 g, similar to that expected in the field. The

rotation of the centrifuge requires the anchors to be installed through a guide to prevent

lateral movement of the anchor during free-fall due to Coriolis acceleration. The anchor

installation guide (Figure 4.3) was fabricated from polyvinyl chloride (PVC) and

featured a 1.0 mm wide slot to accommodate one of the four DEPLA flukes and a 8.3

mm diameter channel machined into the front face of the guide to accommodate the

DEPLA sleeve (7.8 mm diameter). Two PVC rails with an internal circular profile of

8.3 mm diameter were attached to the guide with brackets to prevent the model DEPLA

falling away from the guide. The rails permitted an open slot over the length of the

guide, allowing unobstructed access for the follower and plate anchor lines. Anchor

release was achieved in-flight by a resistor which, when supplied with current, heated

and subsequently burned through an anchor release cord attached to a DEPLA fluke.

This eliminated the need for a shear pin, since the release cord connected to the DEPLA

fluke ensured that the follower and flukes stayed together prior to release. The anchor

velocity was measured using multiple photoemitter-receiver pairs (PERPs) positioned at

10 mm intervals along the guide. The DEPLA fluke travelling through the guide slot

during free-fall would temporarily interrupt each PERP signal and the time between the

trailing edges of consecutive PERP signals together with the fixed PERP spacing of 10

mm allowed the instantaneous velocity of the DEPLA to be determined at various

points above the clay surface.

Page 136: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

102

Figure 4.3. Anchor installation guide

The testing arrangement is shown on Figure 4.4a. Before each DEPLA installation the

model DEPLA was positioned within the installation guide such that the follower tip

was 300 mm above the clay surface (apart from one test with a 200 mm drop height).

The follower line was connected to the actuator and the plate anchor line was

temporarily connected to a fixed point at the base of the actuator with sufficient slack

such that it would not impede the flight of the DEPLA during installation. Each DEPLA

test was conducted at 200 g and comprised the following stages (summarised

schematically in Figure 4.4):

1. Anchor installation:

The model DEPLA was released by burning through the temporary release cord (as

described previously) and the DEPLA travelled vertically through the installation

guide before impacting the clay surface. After anchor release the fluke passing the

highest PERP in the installation guide triggered the high speed data acquisition and

PERP data were logged at 100 kHz.

2. Follower recovery:

The actuator was driven horizontally 43 mm (offset between the vertical axes of the

actuator and the installation guide) such that the vertical axis of the actuator was

directly above the anchor drop site. The follower was extracted by driving the

actuator vertically away from the sample surface at a rate of 0.3 mm/s; this rate was

PERPs

Rails

Bracket

Bracket

Rails

Bracket

Model

DEPLA

Page 137: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

103

selected so that the dimensionless velocity, V = vDf/cv (where v is the extraction

velocity, Df is the follower diameter and cv is the vertical coefficient of

consolidation quantified previously) exceeded 10, so that the response is primarily

undrained (House et al. 2001). Follower extraction was continued until video

footage confirmed that the follower was free from the clay surface.

3. Exchange anchor line:

Immediately after follower extraction the centrifuge was stopped for approximately

five minutes to disconnect the follower line from the load cell and to connect the

plate anchor line. The centrifuge was then restarted and a reconsolidation period of

just over 13 minutes (equivalent to one prototype year at 200 g) was permitted

before commencing the anchor pullout test. The DEPLA plate embedment depth

remains constant during this process.

4. Plate anchor keying and pullout:

The plate anchor keying and pullout were performed continuously at an actuator rate

of 0.1 mm/s such that V > 10 (for the lowest plate anchor diameter, D = 16 mm) to

ensure undrained conditions. Each anchor pullout was vertical, causing the plate

anchor to key from the initial vertical orientation to a final horizontal orientation,

before loading to failure. Pullout of the plate anchor was continued until the plate

was fully extracted from the clay.

The presence of a radial acceleration field limited anchor installation sites to the

longitudinal centreline of the beam centrifuge strongbox. To minimise interaction

effects, installation sites were located at least 20 DEPLA follower diameters (i.e. 120

mm) from rigid boundaries and adjacent installation sites were separated by at least 12

DEPLA follower diameters (i.e. 72 mm). In addition to the dynamic installation tests,

two further, installations were conducted in which the DEPLA was jacked at a constant

penetration rate of 1 mm/s at 1 g using a custom-built 6 mm diameter shaft connected to

Page 138: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

104

the follower padeye. The centrifuge was then accelerated to 200 g and the anchor

keying and pullout were conducted in accordance with Stage 4. As dynamic installation

is known to reduce short term anchor capacity by reducing the effective stresses in the

soil surrounding the embedded anchor (Richardson et al. 2009), these static installation

tests provided a basis for examining the effect of dynamic installation on the keying and

capacity of the plate anchor.

Figure 4.4. Experimental arrangement and testing procedure: (a) prior to

installation, (b) DEPLA embedment, (c) follower retrieval and (d) anchor keying

and pullout

4.4 Results and discussion

Table 4.4 (presented at the end of this chapter) summarises the main results from the ten

centrifuge tests considered here. The results are discussed in further detail in the

following sections.

(a) (b)

(c) (d)

Page 139: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

105

4.4.1 Impact velocity

Anchor velocities determined from the measured PERP data are shown on Figure 4.5

together with theoretical velocity profiles generated using Equation 2.17 albeit with the

frictional and bearing resistance terms omitted whilst accounting for the radially varying

acceleration field in the centrifuge. The theoretical velocity profiles have been adjusted

to fit the PERP data; this adjustment is necessary to account for friction along the

anchor-guide interface, which varies in accordance with the contact area between the

plate anchor and the guide. The reduction of the theoretical velocity for the three tests

shown in Figure 4.5 is in the range 8 to 16%, which is typical for the entire dataset. This

is consistent with a kinetic friction coefficient, μk = 0.17 to 0.36, between the anchor

and the guide, which is reasonable for a plastic-metal interface. As shown in Figure 4.5,

the impact velocity is taken from the adjusted velocity profile at the clay surface. The

impact velocities achieved in the centrifuge tests (for the 300 mm drop height) are in the

narrow range vi = 27.5 to 30.0 m/s, which is typical for a drop height of 300 mm

(Richardson 2008) and is at the upper end of the range from field experience of vi in the

range 24.5 to 27.0 m/s (Lieng et al. 2010) and 16 to 27 m/s (Brandão et al. 2006).

Figure 4.5. Velocity profiles during centrifuge free-fall

0 5 10 15 20 25 30 350

50

100

150

200

250

300

350

Velocity, v (m/s)

v/vtheor

= 84%

v/vtheor

= 89%

v/vtheor

= 92%

v/vtheor

= 93%

Hei

ght

above

sam

ple

(m

m)

Page 140: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

106

4.4.2 Embedment depth

The DEPLA tip embedment, ze was determined using the approach adopted by

Richardson (2008) (see Figure 4.6), where by ze is calculated from:

fslackLCchain Lzzzz e Equation 4.1

where zLC is the initial vertical position of the actuator load cell above the sample

surface, zchain is the length of the follower chain, Lf is the follower length and zslack is the

amount of slack removed from the follower chain prior to the onset of a significant

tensile load during follower extraction. The vertical installation guide located above the

centerline of the soil sample prevents the anchor tilting during the free-fall stage.

Although there is potential for the anchor to deviate slightly from a vertical trajectory

for the remaining one anchor length (on average) embedment, previous centrifuge

studies reported by O’Loughlin et al. (2009) report a maximum inclination of 3 for

dynamically installed anchors in normally consolidated kaolin clay. In the case of the

DEPLA, a 3° inclination would have little bearing on the capacity mobilisation of the

plate, other than a negligible error in determining the embedment depth using Equation

4.1.

Figure 4.6. Determining anchor tip embedment during centrifuge tests

Ziz,eze

Page 141: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

107

DEPLA tip embedments were in the range 144 to 186 mm, or 1.6 to 2.8 times the length

of the DEPLA follower, where the lower and upper bounds correspond to the longest

and shortest DEPLA followers respectively (51.5 mm and 101.5 mm). The majority of

the test data relate to the 76 mm follower (Lf/Df = 12.7), for which ze is in the range 144

to 162 mm, or 1.9 to 2.1 times the follower length. This is in good agreement with

reported field experience of 1.9 to 2.4 times the anchor length for a 79 tonne

dynamically installed anchor with Lf/Df = 10.8 and 4 wide flukes (Lieng et al. 2010).

4.4.3 Follower extraction

Figure 4.7 presents typical load versus displacement data for the extraction of the

DEPLA follower. The extraction response was characterised by a sharp increase in load

after the follower line slack had been taken up, towards an initial maximum (Peak 1)

followed by a sudden drop in load and a subsequent gradual increase towards a

secondary maximum (Peak 2) of lower magnitude than Peak 1. In some instances the

load increased rapidly towards a further maximum (Peak 3), similar to Peak 1, before

reducing suddenly again. The Peak 1 and Peak 2 response has been observed during

extraction of geometrically similar dynamically installed anchors (Richardson et al.

2009) and is considered to be due to the different rates of mobilisation for frictional and

bearing resistance as observed for suction caissons (Jeanjean et al. 2006). It appears that

high, but brittle, frictional resistance along the follower shaft (either against soil or the

DEPLA sleeve) is mobilised first (Peak 1), before a more gradual mobilisation of

bearing resistance at the follower padeye (Peak 2). The reason for Peak 3 and its

sporadic occurrence is not wholly understood, although it may be due to secondary

frictional resistance that occurs as the follower emerges from the upper end of the

DEPLA sleeve. The average displacement to Peak 1 was one follower diameter, with a

range of 0.7 to 1.3 times the follower diameter arising from uncertainties regarding the

Page 142: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

108

onset of follower movement. Interestingly the displacement between Peak 1 and Peak 3

is also approximately one follower diameter.

The magnitude of the Peak 1 capacity ranged from 2.2 to 3.2 times the dry weight of the

follower, with capacity increasing with follower length and hence frictional surface

area. These capacity ratios are lower than those reported for geometrically similar

dynamically installed anchors (e.g. 3 to 5, O’Loughlin et al. 2004b), mainly because the

DEPLA sleeve reduces the frictional surface area of the follower by between 30 and

50% depending on the model anchor configuration. Furthermore, minimal

reconsolidation was permitted before extraction of the follower (limited by the

minimum time required to adjust the actuator position and take up slack in the follower

line). Thus the measured follower extraction resistance represents short term capacity

that is likely to be 20 to 30% of the long term capacity (Richardson et al. 2009).

Figure 4.7. Follower load-displacement response

4.4.4 Plate anchor keying and capacity

Example plate anchor keying and capacity curves are provided on Figure 4.8 for a

square and circular plate anchor installed both dynamically and statically. The keying

0

10

20

30

40

50

60

70

80

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0

Peak 1

Peak 2

Peak 3

Foll

ow

er e

xtr

acti

on r

esis

tance

, F

vf (

N)

ze,follower

/Df

Page 143: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

109

and anchor pullout response is characterised by three distinct stages. After recovering

the slack in the plate anchor line, the load, Fvp, starts to increase as the anchor starts to

key (Stage 1). During Stage 2 Fvp begins to plateau and then reduce slightly. This

reduction was not observed by Gaudin et al. (2006) and Song et al. (2006) in their

respective centrifuge studies on square plates. The reason for this slight reduction is not

understood, but may be attributed to local softening of the soil as the plate rotates. As

keying progresses the effective eccentricity of the padeye reduces, which requires

higher padeye tension to maintain sufficient moment on the plate to continue rotation.

This increase in tension is shown by Stage 3, which continues until the projected plate

area reaches its maximum value, at which point a peak capacity is observed before

reducing again as the plate displaces towards the soil surface through soil of reducing

strength. It is noteworthy that the dynamic installation process does not appear to have a

discernible effect on either the plate anchor keying response or the magnitude of the

peak load.

Figure 4.8. Load-displacement response during anchor keying and pullout

0 10 20 30 40 500

50

100

150

200

250

300

A1: circular plate,

dynamically installed

A1: circular plate,

statically installed

A6: square plate,

dynamically installed

A6: square plate,

statically installed

Load

, F

vp (

N)

Padeye displacement, dv (mm)

1

2

3

Page 144: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

110

The anchor keying process gives rise to two effects. Firstly, the soil in the immediate

vicinity of the plate will be partially remoulded, and secondly, the anchor embedment

depth will reduce as the plate rotates and translates vertically (Randolph et al. 2005).

The loss of embedment due to the latter effect is more significant as any loss in

embedment will correspond to a non-recoverable loss in potential anchor capacity (for

offshore clay deposits that are typically characterised by an increasing strength profile

with depth). The loss in plate anchor embedment during keying is demonstrated by

Figure 4.9, which plots dimensionless load-displacement response for all tests. Loads

are normalised by the area of the plate, Ap, and the undrained shear strength at the

estimated anchor embedment at peak capacity, su,p, which was selected using the T-bar

test data most relevant to each DEPLA test. The vertical displacement of the anchor

padeye is normalised by the diameter, D, or breadth, B, of the plate anchor for circular

and square plates respectively. The occasional sharp drops in the response curves are

considered to be due to minor mechanical effects such as adjustments of the anchor line

at the padeye.

Figure 4.9. Dimensionless load-displacement response during keying and pullout

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.80

2

4

6

8

10

12

14

16

18

20

circular plate, dynamically installed circular plate, statically installed square plate, dynamically installed square plate, statically installed

Normalised padeye displacement, dv/B or d

v/D

Norm

alis

ed l

oad

, F

vp/A

ps u

,p (Fvp

-Ws)/(A

ps

u,p)

Page 145: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

111

In the main, the load-displacement responses in Figure 4.9 are typical of those observed

experimentally for plate anchors (e.g. Song et al. 2006, Gaudin et al. 2006), with an

initial stiff response as the anchor begins to rotate, followed by a softer response as the

rotation angle increases, and a final stiff response as the anchor capacity is fully

mobilised. The range of (Fvp-Ws)/Apsu,p values, adjusting the peak force by the

submerged weight of the anchor, WS, are also indicated in Figure 4.9, and are compared

with theoretical bearing capacity factors later in the chapter. The vertical displacement

of the anchor padeye, dv, is in the range 0.96 to 1.12 times the diameter for circular

plates and 1.16 to 1.26 times the breadth for square plates. However dv represents the

displacement of the anchor padeye rather than the loss in plate anchor embedment

which may be calculated as:

pullvplatee edz sin, Equation 4.2

where e is the padeye eccentricity and θpull is the load inclination (to the horizontal) at

the anchor padeye. Hence for a vertical pullout ∆ze,plate is in the range 0.50 to 0.68 times

the diameter for circular plates and 0.53 to 0.63 times the breadth for square plates. This

is shown on Figure 4.10 together with other centrifuge data for square, rectangular (L/B

= 2) and strip anchors. The horizontal axis on Figure 4.10 combines the two main

parameters that influence the keying response, padeye eccentricity and plate thickness

(and hence anchor weight). Also shown on Figure 4.10 are numerical data reported by

Wang et al. (2011), who showed that embedment loss increases with increasing shear

strength for a given eccentricity and anchor thickness to a limiting value at a normalised

shear strength, su/γ'aB ≈ 0.75. Wang et al. (2011) approximate this maximum loss in

embedment for square anchors as:

Page 146: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

112

15.12.0

maxe, 144.0

B

Δz

B

t

B

e Equation 4.3

Equation 4.3 is also shown on Figure 4.10 and represents the theoretical maximum loss

in embedment for a given anchor eccentricity and thickness. Experimental data for the

DEPLA square plates are in reasonable agreement with Equation 4.3, ranging between

94 and 113% of Δze,max calculated using Equation 4.3. Applying Equation 4.3 to the

circular DEPLA plate, but with D instead of B, reduces the agreement to 61 to 79%.

The poorer agreement for the circular plate data is not unexpected as Wang et al. (2011)

note that the maximum embedment loss differs for different plate shapes. Figure 4.10

indicates that the loss in DEPLA plate embedment during keying could be reduced if

either the eccentricity or anchor thickness was increased (relative to the plate diameter

or breadth), or if the soil was weaker (relative to the initial moment applied to the

anchor at the onset of keying). The most influential of these factors is the eccentricity

(O’Loughlin et al. 2006, Song et al. 2009, Wang et al. 2011) and the embedment loss

for the DEPLA should be lower if the padeye eccentricity was increased (albeit that this

would require a different plate geometry). Whilst an increase in plate thickness may be

attractive in that it should reduce embedment loss during keying and hence the loss in

potential anchor capacity, this may be offset by the cost implications of increasing the

thickness of the plate.

Page 147: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

113

Figure 4.10. Loss in anchor embedment during keying

Experimental bearing capacity factors for the DEPLA are calculated using:

pup

netv

csA

FN

,

, Equation 4.4

where Fv,net is the peak load minus the submerged weight of the plate anchor in soil, Ap

is the area of the plate and su,p is the undrained shear strength at the estimated anchor

embedment (anchor mid-depth) at peak capacity as defined previously. Experimentally

derived Nc factors are shown on Figure 4.11 and are typically in the range Nc = 14.2 to

15.8. The average experimental Nc = 15.0 for both circular and square plates is in good

agreement with Nc = 15.3 back analysed from reduced scale DEPLA field trials (Blake

and O’Loughlin 2012) and Nc = 14.9 for circular DEPLA determined numerically using

3D LDFE (Wang and O’Loughlin 2014). Corresponding solutions for rough flat plate

anchors are lower, e.g. Nc = 13.11 for an infinitely thin circular plate (Martin and

Randolph 2001), Nc = 13.70 for a circular plate with thickness equal to 5% of the

diameter (Song et al. 2008b) and Nc = 14.33 for a square plate with thickness equal to

5% of the diameter (Wang et al. 2010). However, as noted by Zhang et al. (2012) soil is

forced to take a deeper and longer route to flow around thicker plates, which mobilises

0.0 0.2 0.4 0.6 0.8 1.00.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

su/

'

aB =

Experimental: DEPLA, circular (exp) DEPLA, square (exp) O'Loughlin et al. (2006), strip (exp) O'Loughlin et al. (2006), L/B = 2 (exp) Gaudin et al. (2009), strip (exp) Song et al. (2009), square (exp)

Numerical, Wang et al. (2011):

su/

a'B = 0.01 s

u/

a'B = 0.015

su/

a'B = 0.025 s

u/

a'B = 0.035

su/

a'B = 0.05 s

u/

a'B = 0.075

su/

a'B = 0.1 s

u/

a'B = 0.2

z e,

pla

te/B

or

z e,p

late/D

[e/B×t/B0.2

]1.15

or [e/D×t/D0.2

]1.15

su/

'

aB = (fitted by Equation 4.3)

su/

'

aB =

Page 148: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

114

more soil in the failure mechanism and results in higher bearing capacity factors. The

DEPLA cruciform fluke arrangement ensures that the failure surface extends to at least

half the plate diameter (D/2) from the plate, which should result in a higher bearing

capacity factor for the DEPLA compared with a flat plate. From a design perspective,

anchor capacity would be based on the (higher) strength at the initial plate location. The

average experimental capacity factor determined in this way is Nc = 11.8, 21% lower

than the average Nc = 15 based on the strength at the anchor embedment at peak

capacity.

The final plate embedment of the anchor (at peak anchor capacity) is in the range 1.8 to

5.8 times the plate diameter or breadth, with lower values corresponding with larger

plates and vice versa. The lower bound of this range is still sufficiently high to mobilise

a deep failure mechanism (Song and Hu 2005, Song et al. 2008b, Wang et al. 2010).

This is supported by the observation that the experimental Nc values do not reduce over

this range of normalised embedment, indicating that the DEPLA mobilises a deep

failure mechanism at relatively shallow plate embedment ratios.

Figure 4.11. Comparison of experimental and numerical bearing capacity factors

0 1 2 3 4 5 6 7 80

2

4

6

8

10

12

14

16

18

Ciruclar plates

Square plates

Martin and Randolph (2001)

Son and Hu (2005)

Song et al. (2008b)

Wang and O'Loughlin (2014)

Bea

ring c

apac

ity f

acto

r, N

c

ze,plate

/B or ze,plate

/D

Page 149: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

115

4.5 Conclusions

This chapter has presented centrifuge data from a series of beam centrifuge tests

undertaken to assess the penetration and capacity response of dynamically embedded

plate anchors in normally consolidated clay. The main findings from this study are

summarised in the following.

At mudline impact velocities of approximately 30 m/s, DEPLA tip embedment

depths are in the range 1.6 to 2.8 times the length of the DEPLA follower, which is

in good agreement with reported field experience of 1.9 to 2.4 times the anchor

length for a 79 tonne dynamically installed anchor with L/d = 10.8 and 4 wide

flukes (Lieng et al. 2010).

Extraction of the DEPLA follower is characterised by an initial sharp increase in

pullout resistance corresponding with frictional resistance along the follower shaft,

followed by reduction in soil (and mechanical) friction and then a more gradual

increase in pullout resistance corresponding with bearing resistance at the follower

padeye. In some instances a second frictional capacity peak was observed at

approximately one follower diameter after the first frictional capacity peak. This is

considered likely to arise from frictional resistance as the follower emerges from the

top of the DEPLA sleeve. Follower pullout resistances quantified using the initial

capacity peak were in the range 2.2 to 3.2 times the dry weight of the follower.

The anchor keying and pullout response is typical of that for plate anchors, with an

initial stiff response as the anchor begins to rotate, followed by a softer response as

the rotation angle increases, and a final stiff response as the effective eccentricity of

the padeye reduces and anchor capacity is fully mobilised. The dynamic installation

process does not appear to have a discernible effect on either the plate anchor keying

response or the magnitude of the peak load.

Page 150: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

116

The loss in plate anchor embedment during keying is in the range 0.50 to 0.68 times

the diameter for circular plates and 0.53 to 0.63 times the breadth for square plates.

The square plate embedment loss data are within 94 to 115% of the value calculated

using formulations proposed by Wang et al. (2011) for square plates, and may be

reduced further by increasing the padeye eccentricity.

Experimental bearing capacity factors are typically in the range Nc = 14.2 to 15.8,

with an average Nc = 15.0 for both square and circular plates, which is higher than

existing numerical solutions for rough flat circular and square plates. The higher

bearing capacity factor for the DEPLA is considered to be due to the cruciform fluke

arrangement which ensures that the failure surface extends to the edge of the

orthogonal flukes.

Page 151: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

117

Table 4.4. Summary of test results

Tes

t

no.

An

cho

r re

f

Inst

alla

tio

n

met

hod

Imp

act

velo

city

,

v i (

m/s

)

Tip

emb

edm

ent

dep

th, z

e

(mm

)

Fol

low

er

extr

acti

on

load

, Fvf

(N)

Init

ial

anch

or

emb

edm

ent,

z e,p

late

,0 (

mm

)

Nor

mal

ised

loss

in

pla

te

anch

or

emb

edm

ent,

∆z e

,pla

te/D

or

∆z e

,pla

te/B

Fin

al a

nch

or

emb

edm

ent,

z e,p

late

(m

m)

Pla

te

emb

edm

ent

rati

o,

z e,p

late

/D o

r

z e,p

late

/B

Un

dra

ined

shea

r

stre

ngt

h a

t

pea

k

cap

acit

y,

s u,p

(k

Pa)

Net

pla

te

anch

or

cap

acit

y,

Fv,

net

(N

)

Bea

rin

g

cap

acit

y

fact

or,

Nc

1 A

1

Dyn

amic

23

.87

155.

4 69

.99

95.4

0.

55

79.0

2.

63

21.5

23

8.44

15

.8

2 A

1

Dyn

amic

30

.04

161.

4 78

.37

101.

4 0.

59

83.7

2.

79

24.1

29

7.70

17

.3

3 A

6

Dyn

amic

27

.53

162.

4 61

.79

102.

9 0.

61

86.5

3.

24

23.5

23

3.31

14

.5

4 A

2

Dyn

amic

28

.50

156.

9 65

.12

93.0

0.

64

78.5

3.

47

23.5

13

7.78

14

.9

5 A

4

Dyn

amic

29

.17

145.

9 40

.84

102.

3 0.

57

93.1

5.

82

29.3

71

.10

14.8

6 A

5

Dyn

amic

29

.37

186.

4 75

.93

100.

9 0.

55

84.4

2.

81

24.5

22

0.62

12

.8

7 A

6

Dyn

amic

29

.40

144.

4 60

.74

84.9

0.

63

68.0

2.

54

20.0

20

4.34

14

.9

8 A

3

Dyn

amic

29

.17

150.

6 45

.55

94.5

0.

66

69.6

1.

85

21.7

37

8.51

15

.8

9 A

1

Sta

tic

- 16

0.6

- 10

0.5

0.50

85

.4

2.85

23

.0

229.

66

14.2

10

A6

S

tati

c -

157.

0 -

97.5

0.

53

83.4

3.

12

23.8

24

1.94

14

.9

Page 152: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

118

CHAPTER 5 IN SITU MEASUREMENT OF THE

DYNAMIC PENETRATION OF FREE-

FALL PROJECTILES IN SOFT SOILS

USING A LOW COST INERTIAL

MEASUREMENT UNIT

5.1 Abstract

Six degree-of-freedom motion data from projectiles free-falling through water and

embedding in soft soil are measured using a low-cost inertial measurement unit,

consisting of a tri-axis accelerometer and a three-component gyroscope. A

comprehensive framework for interpreting the measured data is described and the merit

of this framework is demonstrated by considering sample test data for free-falling

projectiles that gain velocity as they fall through water and self-embed in the underlying

soft clay. The chapter shows the importance of considering such motion data from an

appropriate reference frame by showing good agreement in embedment depth data

derived from the motion data with independent direct measurements. Motion data

derived from the inertial measurement unit are used to calibrate a predictive model for

calculating the final embedment depth of a dynamically installed anchor.

5.2 Introduction

An inertial measurement unit (IMU) is an electromechanical device that measures an

object’s six degree of freedom (6DoF) motion in three-dimensional space using a

combination of gyroscope and accelerometer sensors. The development of micro-

electro mechanical systems (MEMS) gyroscope and accelerometer technology has

significantly reduced the cost, size, weight and power consumption of IMUs, and

enhanced their robustness.

Page 153: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

119

MEMS accelerometers and gyroscopes are typically fabricated on single-crystal silicon

wafers using micromachining to etch defined patterns on a silicon substrate. These

patterns take the form of small proof masses that are free from the substrate and

surrounded by fixed plates. The proof mass is connected to a fixed frame by flexible

beams, effectively forming spring elements. Low-cost consumer grade MEMS

gyroscopes typically use vibrating mechanical elements to sense angular rotation rate.

During operation the proof mass is resonated with constant amplitude in the ‘drive

direction’ by an external sinusoidal electrostatic or electromagnetic force. Angular

rotation then induces a matched-frequency sinusoidal Coriolis force orthogonal to the

drive-mode oscillation and the axis of rotation. The Coriolis force deflects the proof

mass and plates connected to the proof mass move between the fixed plates in the sense-

mode. The operational principle for MEMS accelerometers is much simpler;

accelerations acting on the proof mass cause it to displace, and plates connected to the

proof mass move between fixed plates. For both sensors, the movement of the plates

cause a differential capacitance that is measured by integrated electronics and is output

as a voltage that is proportional to either the applied angular rotation rate (in the case of

MEMS gyroscopes) or acceleration (in the case of MEMS accelerometers). The

operational principles of the MEMS accelerometers and gyroscopes as described above

are shown schematically in Figure 5.1.

Page 154: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

120

(a)

(b)

Figure 5.1. Schematic representation of the operational principle of: (a) MEMS

accelerometers and (b) MEMS gyroscopes

Common applications of low-cost IMUs featuring MEMS technology include: inertial

navigation systems (e.g. remotely operated vehicles, autonomous underwater vehicles

and unmanned aerial vehicles), active safety systems (electronic stability control and

traction control in motor vehicles) and motion-activated user-interfaces (e.g.

smartphones, game controllers and tablet computers). The use of low-cost 6DoF IMUs

Page 155: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

121

for geotechnical applications has not been reported. However, MEMS accelerometers

have been used for in situ geotechnical applications to measure: inclinations in

boreholes (Bennett et al. 2009), soil displacement associated with rapid uplift of

footings (Levy and Richards 2012) and the motion of free-falling cone penetrometers

(e.g. Stegmann et al. 2006, Stephan et al. 2012, Steiner et al. 2014). In geotechnical

centrifuge modelling MEMS accelerometers have been used to measure: the

acceleration response of a free-falling projectiles in clay (O’Loughlin et al. 2014a,

Chow et al. 2014), earthquake accelerations (Cilingir and Madabhushi 2011, Stringer et

al. 2010) and rotation of structures during slow lateral cycling and dynamic shaking

(Allmond et al. 2014). Although, accelerometers are often used to measure the rotation

of objects at constant acceleration, they cannot distinguish rotation from linear

acceleration if the object’s orientation and acceleration is changing. However,

gyroscopes are unaffected by linear acceleration, and the rotation of accelerating objects

can be derived from their measurements. Hence the combination of accelerometer and

gyroscope measurements enables an object’s linear acceleration to be determined

relative to a reference frame that is not necessarily coincident with the reference frame

of the object. This becomes important for the applications considered in this chapter,

where dynamically installed anchors and a free-falling sphere (collectively referred to as

‘projectiles’ from this point forward) free fall through water and bury in the underlying

soil. As described later, the motion response of the projectile must be considered from

the appropriate reference frame. From the viewpoint of the hydrodynamic and

geotechnical resistances acting on the projectile during motion, it becomes important to

consider the projectile’s trajectory, whereas from a geotechnical design viewpoint the

final depth and orientation of the projectile relative to a fixed inertial frame of reference

(with an axis in the direction of Earth’s gravity) is important as this will dictate the local

Page 156: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

122

soil strength in the vicinity of the embedded projectile and (for the case of the anchors)

how this strength will be mobilised during loading.

This chapter describes a custom-design, low-cost MEMS based IMU and presents a

comprehensive framework for interpreting the IMU measurements (which are made in

the body frame of reference) so that they are coincident with a fixed inertial frame of

reference. The framework is implemented to establish rotation, acceleration and velocity

profiles for the projectiles during free-fall in water and embedment in soil. The final

projectile embedment depths established from the IMU data are compared with direct

measurements, and the merit of collecting motion data during dynamic penetration is

demonstrated by using such data to verify the appropriateness of an embedment

prediction model for dynamically installed anchors.

5.3 Free-falling projectiles

5.3.1 Deep penetrating anchor

The deep penetrating anchor (DPA) is a proprietary term for a dynamically-installed

anchor design. The DPA is designed so that, after release from a designated height

above the seafloor, it will penetrate to a target depth in the seabed using the kinetic

energy gained through free-fall. The DPA data considered here are from tests using a

1:20 reduced scale model anchor based on an idealised design proposed by Lieng et al.

(1999). The model DPA (see

Figure 5.2), was fabricated from mild steel and had an overall length of 750 mm, a shaft

diameter of 60 mm and a mass of 20.7 kg. The anchor had an ellipsoidal tip and

featured four clipped delta type flukes (separated by 90º in plan) with a forward swept

trailing edge. The anchor shaft was solid with the exception of a watertight cylindrical

void towards the top to house the IMU.

Page 157: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

123

Figure 5.2. Deep penetrating anchor

5.3.2 Dynamically embedded plate Anchor

The dynamically embedded plate anchor (DEPLA, O’Loughlin et al. 2013a) is an

anchoring system that combines the capacity advantages of vertically loaded anchors

with the installation advantages of dynamically installed anchors. The DEPLA

comprises a removable central shaft or ‘follower’ and a set of four flukes (see Figure

5.3). A stop cap at the upper end of the follower prevents it from falling through the

DEPLA sleeve and a shear pin connects the flukes to the follower. The DEPLA is

installed in a similar manner as the DPA, but after coming to rest in the seabed the

follower retriever line is tensioned, which causes the shear pin to part (if not already

broken during impact) allowing the follower to be retrieved for the next installation

whilst leaving the anchor flukes vertically embedded in the seabed. These embedded

anchor flukes constitute the load bearing element as a plate anchor.

In the tests considered here the DEPLA was modelled at a reduced scale of 1:4.5 and

fabricated from mild steel. The follower (and hence DEPLA) length was 2 m, the

follower diameter was 160 mm, the fluke (plate) diameter was 800 mm and the overall

mass was 388.6 kg. As with the DPA, the DEPLA follower was solid with the exception

of a cylindrical void at the top to house the IMU. The model DEPLA is shown in Figure

5.3.

Page 158: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

124

Figure 5.3. Dynamically embedded plate anchor

5.3.3 Instrumented free-falling sphere

The instrumented free-falling sphere (IFFS) has been proposed as an in situ

characterisation tool for soft soils (Morton and O’Loughlin 2012, O’Loughlin et al.

2014a). The IFFS is a steel sphere that dynamically embeds in soft soil in a manner

similar to dynamically installed anchors. IMU data measured during embedment in soil

can be used to estimate undrained shear strength. As such, the IFFS is conceptually

similar to a free fall cone penetrometer, but the simple spherical geometry of the IFFS is

beneficial as the projected area does not change with rotation and the bearing factor for

the ball is more tightly constrained than for the cone. The IFFS data considered here are

from tests using a 250 mm diameter mild steel sphere with a mass of 50.8 kg. The IFFS

was fabricated as two hemispheres (that could be bolted together) with an internal

vertically orientated cylindrical void to accommodate the IMU (see Figure 5.4)

Figure 5.4. Instrumented free-falling sphere

Page 159: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

125

5.3.4 Inertial measurement unit

The IMU was used to measure projectile accelerations and rotation rates during free-fall

in the water column and embedment in the soil. The IMU (see Figure 5.5) includes a 16

bit three component MEMS rate gyroscope (ITG 3200) and a 13 bit three-axis MEMS

accelerometer (ADXL 345). The gyroscope had a resolution of 0.07 °/s with a

measurement range of +/- 2000 °/s. The accelerometer had a resolution of 0.04 m/s2

with a measurement range of +/- 16 g. Data were logged by an mbed micro controller

with an ARM processor to a 2 GB SD card at 400 Hz. Internal batteries were capable of

powering the logger for up to 4 hours. The IMU was contained in a watertight

aluminium tube 185 mm long and 42 mm in diameter and was located in a void (with

the same dimensions) within the projectile. The IMU had a mass of approximately 0.5

kg (including the batteries).

The accelerometer and gyroscope are aligned with the body frame of the projectile and

the IMU as shown in Figure 5.6 (for the DEPLA). The body frame is a reference frame

with three orthogonal axes xb, yb and zb that are common to both the IMU and the

projectile and where the zb-axis is parallel to the direction of earth’s gravity when the

projectile is hanging vertically. The accelerometer measures accelerations Abx, Aby and

Abz in the body frame along these three axes. These accelerometer measurements

include a component of gravitational acceleration (depending on the orientation of the

accelerometer) and linear acceleration. The gyroscope measures angular velocities ωbx,

ωby and ωbz in the body frame about the same orthogonal axes. Accelerometers are often

used to measure the rotation of quasi-static objects but cannot distinguish rotation from

linear acceleration if an object is in motion. However gyroscopes are unaffected by

linear acceleration and the rotation of objects in motion can be derived from their

measurements.

Page 160: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

126

Figure 5.5. Inertial measurement unit

Figure 5.6. Body frame of reference

ωby

ωbx

ωbz

yb xb

zb

Aby Abx

Abz

Page 161: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

127

5.4 Interpretation of IMU measurements

As the body frame is not fixed in space, it is necessary to define an inertial frame,

defined here and used in this chapter, as a local fixed reference frame, with the z-axis

aligned in the direction of the Earth’s gravitational vector, and with undefined

orthogonal x- and y-axes, that are fixed at their orientation at the start of each test. If the

projectile pitches and/or rolls whilst in motion, the body frame will move out of

alignment with the inertial frame of reference and the rotation rates ωbx, ωby and ωbz and

accelerations Abx, Aby and Abz measured by the IMU will not be coincident with the

inertial frame (see Figure 5.7). As a consequence gravitational acceleration, g, and

linear acceleration, a, (required for velocity and translation calculations as described

later), components cannot be distinguished from the accelerometer measurements.

Hence the IMU measurements were ‘transformed’ from the body frame to the inertial

frame. This was accomplished using transformation matrices as described in the

following sections.

Figure 5.7. Body and inertial frames of reference

y

zb

μ

z

x

xb

yb

Page 162: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

128

5.4.1 Rotations

The body frame rotation rates ωbx, ωby, and ωbz were transformed from the body frame

to the inertial frame to correspond with rotation rates about the inertial frame ωx, ωy and

ωz using an angular velocity transformation matrix (AVTM),i

bT (Fossen 2011):

bz

by

bx

i

b

z

y

x

T

Equation 5.1

bbbb

bb

bbb

i

b

b

T

cos/coscos/sin0

sincos0

tancostansin1

Equation 5.2

where ϕb and θb are the current rotation angles about the body frame axes xb and yb

respectively established from numerical integration of ωbx and ωby:

t

bxbb dttt0

0 )()( Equation 5.3

t

bybb dttt0

0 )()( Equation 5.4

Similarly, the rotation angle ψb about the body frame axis zb was established by

numerical integration of ωbx:

t

bzbb dttt0

0 )()( Equation 5.5

Numerical integration of the angular velocities ωx, ωy and ωz derived from the AVTM

allowed the roll, ϕ, pitch, θ and yaw, ψ, rotations about the inertial frame axes z, y and z

respectively (Euler angles) to be established:

t

x dttt0

0 )()( Equation 5.6

Page 163: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

129

t

y dttt0

0 )()( Equation 5.7

t

z dttt0

0 )()( Equation 5.8

5.4.2 Accelerations

The accelerometer measurements Abx, Aby and Abz were converted to accelerations

coincident with the inertial frame Ax, Ay and Az using a direction cosine matrix (DCM),

i

bR (Nebot and Durrant-Whyte 1999, Jonkman 2007, King et al. 2008, Fossen 2011):

bz

by

bx

i

b

z

y

x

A

A

A

R

A

A

A

Equation 5.9

)()()( xyz

i

b RRRR Equation 5.10

The DCM relates the accelerations measured in the body frame to the inertial frame by

considering three successive rotations of yaw -ψ, pitch -θ, and roll, -ϕ, about the inertial

frame axes z, y and x respectively. These rotations are represented by the yaw Rz(-ψ),

pitch Ry(-θ) and roll Rx(-ϕ), matrices that are used to rotate the measured acceleration

vectors Abx, Aby and Abz in Euclidean vector space:

100

0cossin

0sincos

zR Equation 5.11

cos0sin

010

sin0cos

yR Equation 5.12

cossin0

sincos0

001

xR Equation 5.13

Page 164: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

130

Multiplication of the Rz(-ψ), Ry(-θ) and Rx(-ϕ) rotation matrices gives the DCM:

coscossincossin

sinsincossincossinsinsincoscossincos

sincoscossinsinsinsincossincoscoscosi

bR

Equation 5.14

The linear accelerations coincident with the inertial frame ax, ay and az were derived

from the transformed accelerometer measurements Ax, Ay and Az (Az is a negative

output, i.e. when the projectile is at rest, az = Az + g = 0) using the following expression:

(Stovall 1997, Noureldin et al. 2012):

gAz

Ay

Ax

a

a

a

z

y

x

0

0

Equation 5.15

The resultant linear acceleration, a (acceleration in the direction of motion), was

calculated as:

gAAAa zyx 222 Equation 5.16

5.4.3 Velocities and distances

The linear accelerations corresponding to the inertial frame ax, ay and az were

numerically integrated to establish the projectile velocities coincident with the inertial

frame vx, vy and vz during free-fall in the water column and embedment in the soil:

t

xxx dttavtv0

0 )()( Equation 5.17

t

yyy dttavtv0

)(0)( Equation 5.18

t

zzz dttavtv0

0 )()( Equation 5.19

Page 165: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

131

The resultant projectile velocity, v, was calculated using the following expression:

222

zyx vvvv Equation 5.20

The resultant projectile velocity, v, was numerically integrated to establish the distance

travelled by the projectile along its trajectory, s:

t

dttvsts0

0 )()( Equation 5.21

The distance travelled by the projectile along the inertial z axis sz (required to calculate

the vertical embedment depth of the projectile relative to the soil surface, ze), was

established by numerically integrating the vertical velocity, vz:

t

zzz dttvsts00

)()( Equation 5.22

5.4.4 Tilt angles

Following dynamic penetration the projectile is at rest in the soil and has no linear

acceleration. Under these conditions the accelerometer measurements can be used to

derive the final pitch, ϕacc, roll, θacc, (coincident with the inertial frame) and resultant

tilt, μ, (tilt relative to Earth’s gravitational vector, see Figure 5.7) angles using the

following expressions:

g

Aby

acc

1sin (King et al. 2008) Equation 5.23

g

Abxacc

1sin (King et al. 2008) Equation 5.24

g

Abz1cos (Stephan et al. 2012) Equation 5.25

Page 166: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

132

5.5 Test sites and soil properties

The IMU performance has been examined using projectile data from two sites. The

DEPLA data considered here relate to tests conducted in the Firth of Clyde which is

located off the West coast of Scotland between the mainland and the Isle of Cumbrae.

The DPA and IFFS data are from tests conducted in Lower Lough Erne, which is an

inland lake located in County Fermanagh, Northern Ireland. At Lough Erne the water

depths at the test locations varied between 3 and 20 m whereas at the Firth of Clyde test

locations the water depth was typically 50 m. Both test locations are shown in Figure

5.8.

The seabed at the DEPLA test locations in the Firth of Clyde is very soft with moisture

contents in the range 50 to 100% (close to the liquid limit). Consistency limits plot

above or on the A-line on the Casagrande plasticity chart, indicating aclay of

intermediate to high plasticity. The unit weight increases from about γ = 14 kN/m3 at

the mudline to about γ = 18 kN/m3 at about 3.5 m (limit of the sampling depth). Figure

5.9a shows profiles of undrained shear strength, su, with depth derived from piezocone

and piezoball tests, and calibrated using lab shear vane data and fall cone tests, to give

piezocone bearing factors Nkt = 17.8 (5 cm2 cone) and Nkt = 16.9 (10 cm

2 cone), and

piezoball bearing factors Nball = 11.5 (50 cm2 ball) and Nball = 12.2 (100 cm

2 ball). The

su profile is best idealised as su (kPa) = 2 + 2.8z over the upper z = 5 m of the

penetration profile, which is the depth of interest for the DEPLA tests. The sensitivity

ratio of the remoulded to intact soil restance is in the range 0.19 to 0.33 as assessed

from piezoball cyclic remoulding tests. This range is similar, but not identical to the

range of soil sensitivity, as the bearing factor for remoulded soil is greater than for intact

soil (Yafrate et al. 2009, Zhou and Randolph 2009).

The Lough Erne lakebed is very soft clay with moisture contents in the range 270 to

520%, typically about 1.5 times the liquid limit. The measured unit weight of the Lough

Page 167: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

133

Erne clay is only marginally higher than water at γ = 10.8 kN/m3. This is considered to

be due to the very high proportion of diatoms that are evident from scanning electron

microscopic images of the soil (e.g. see Colreavy et al. 2012) and which have an

enormous capacity to hold water in the intraskeletal pore space (Tanaka and Locat

1999). Colreavy et al. (2012) report data from piezoball penetration tests (using a 100

cm2 ball) at the Lough Erne site to depths of up to 8 m. Figure 5.9b shows su profiles

with depth, obtained from the net penetration resistance using Nball = 8.6, calibrated

using in situ shear vane data. The undrained shear strength profile is best idealised over

the depth of interest (0 to 2.2 m) as su (kPa) = 1.5z. Piezoball cyclic remoulding tests

show that the ratio of remoulded to intact soil resistance is in the narrow range 0.4 to

0.5, indicating a low sensitivity soil.

Figure 5.8. Test site locations

Lough Erne

Firth of Clyde

Page 168: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

134

(a) (b)

Figure 5.9. Undrained shear strength profiles: (a) Firth of Clyde and (b) Lough

Erne

5.6 Test procedure

Testing was conducted using the RV Aora, a 22 m research and survey vessel in Firth of

Clyde (Figure 5.10) and either a fixed vessel berthing jetty or a 15 m self-propelled

barge (Figure 5.11) in Lough Erne. The self-propelled barge was equipped with a 13

tonne winch and a 2 tonne crane, whereas the RV Aora was equipped with several

winches and an 8 tonne crane. The testing procedure for each site and projectile was

broadly similar (summarised schematically in Figure 5.12 for the DEPLA tests using the

RV Aora) and involved the following stages:

1. The IMU was powered up and secured in the projectile.

0 5 10 15 20

5.0

4.5

4.0

3.5

3.0

2.5

2.0

1.5

1.0

0.5

0.0

10 cm2 cone

5 cm2 cone

100 cm2 ball

50 cm2 ball

Lab vane

Fall cone

Undrained shear strength, su (kPa)

Pen

etra

tio

n d

epth

, z

(m)

su = 2 + 2.8z

0 1 2 3 4 5 6 7 8

8

7

6

5

4

3

2

1

0

su = 1.5z

z

Ball 1

Ball 2

Ball 3

Ball 4

Vane

Undrained shear strength, su (kPa)

Pen

etra

tion d

epth

, z z

(m)

Page 169: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

135

2. The projectile was lowered below the water surface to the desired drop height

above the mudline.

3. The projectile was released by opening a quick release shackle connecting the

projectile release/retrieval line to the crane, allowing the projectile to free-fall and

penetrate the soil.

4. The projectile tip embedment depth ze, was measured by sending a remotely

operated vehicle (ROV) (Firth of Clyde), or a drop camera (Lough Erne) to the

mudline to inspect markings on the projectile retrieval line (see Figure 5.13).

Figure 5.10. RV Aora

Figure 5.11. Self-propelled barge

Page 170: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

136

Figure 5.12. DEPLA field test procedure

Water

Seabed

(a)

Follower line

DEPLA

Water

(b)

Follower line

Water

(c)

Follower line

ROV tether

ROV

Plate line

Plate line

Plate line

RV Aora

Crane Pulley

Drum Winch

Pulley

Drum Winch

Crane

Quick release shackle

A-frame

A-frame

PulleyCrane

A-frame

Load cell

Pulley

RV Aora

RV Aora

Seabed

Seabed

Drop height

Embedment depth, ze

Page 171: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

137

Figure 5.13. Image capture from ROV camera showing the follower retrieval line

at the seabed

5.7 Results and discussion

The IMU data were interpreted within the framework described above, which can be

readily implemented in a spreadsheet application such as Microsoft Excel or

alternatively using numerical analysis software such as MATLAB.

5.7.1 Rotations

Rate gyroscopes are subject to an error known as bias drift where the zero rate output

drifts over time (Sharma 2007). However, the duration of a projectile drop never

exceeded 6.5 s, which is too short for any measurable bias drift to accumulate. This was

confirmed by comparing the zero rate outputs before the drop when the anchor was

hanging in the water with the zero rate outputs after the drop when the anchor was at

rest in the soil. No change was observed for any test.

Figure 5.14 shows typical rotation profiles during free-fall in water and embedment in

the lakebed for each of the three projectiles, released from drop heights of 17.69 m

(DEPLA), 5.95 m (IFFS), and 3 m (DPA). In Figure 5.14 ϕacc and θacc are rotations

relative to the inertial frame deduced from the horizontally orientated y- and x-axes

accelerometers using Equation 5.23 and Equation 5.24, ϕb, θb and ψb are rotations about

the body frame axes xb, yb and zb established using Equation 5.3, Equation 5.4 and

Page 172: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

138

Equation 5.5, and ϕ, θ and ψ are the pitch, roll and yaw rotations about the inertial frame

axes x, y and z derived using Equation 5.6, Equation 5.7 and Equation 5.8.

In Figure 5.14a, prior to release (time, t = 0 to 1.1 s) the DEPLA was swaying in the

water, suspended from the installation line, during which time rotations derived from

the accelerometer measurements (ϕacc and θacc) and from the gyroscope measurements

(ϕb and θb) were in broad agreement. During free-fall (t = 1.1 s to 3.59 s) rotations can

only be interpreted from the gyroscope measurements as the accelerometer

measurements include both acceleration and rotation components. The gyroscope

measurements indicate that rotations reached ϕb = 17.3 and θb = -8.3° when the anchor

came to rest in the lakebed at t = 4.2 s. There is a discrepancy of Δϕ = 1.7 and Δθ =

3.1° between the accelerometer and gyroscope measurements whilst the anchor is at

rest. However, when the anchor was at rest in the soil the ‘transformed’ rotations

derived from the gyroscope measurements (ϕ and θ) were in good agreement with

rotations derived from the accelerometer measurements, as both were coincident with

the inertial frame of reference.

Figure 5.14b shows that the IFFS rotated about all three axes during free-fall in water

and penetration in soil. Indeed, the non-zero ψb and ψ response started whilst the IFFS

was hanging in water, indicating that the IFFS started to spin before it was released.

After the IFFS came to rest in the soil there is a discrepancy of Δϕ = 4.1 and Δθ = 2.8

between the final accelerometer and gyroscope measurements. As with the DEPLA test,

the transformed rotations derived from the gyroscope measurements were in good

agreement with rotations derived from the accelerometer measurements. This highlights

the importance of using the AVTM to transform the angular velocities measured by the

gyroscope from the body frame to the inertial frame to establish rotations that relate to

the inertial frame.

Page 173: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

139

In contrast, rotations measured during the DPA free-fall and embedment phases (Figure

5.14c) were much lower than from the DEPLA and IFFS tests. Indeed, the rotation

appears to have only occurred before release (due to swaying and spinning in water) and

at the start of the free-fall phase, indicating that the DPA tends to self-correct and

become hydrodynamically stable during free-fall in water. As such the misalignment

between the body frame of the IMU (and hence the anchor) and the inertial frame of

reference in this case was negligible, with no discernible differences in the rotations

derived from the final accelerometer and gyroscope measurements when the anchor

came to rest in the soil. Hence transformation of rotations between the body frame and

the inertial frame may not be warranted in cases where the rotations are relatively small.

Page 174: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

140

(a)

(b)

(c)

Figure 5.14. Projectile rotations during free-fall through water and soil

penetration: (a) DEPLA, (b) IFFS and (c) DPA

Page 175: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

141

5.7.2 Accelerations

Figure 5.15 shows acceleration profiles for the same tests as shown in Figure 5.14. In

Figure 5.15 Abx, Aby and Abz are the accelerometer measurements and Ax, Ay and Az are

the transformed accelerometer measurements that are coincident with the inertial frame

(i.e. Aiz is the acceleration measurement in the direction of gravity). In Figure 5.15a the

DEPLA was initially hanging in the water experiencing only gravitational acceleration

with Ax = 0 (ax = 0), Ay = 0 (ay = 0) and Az = -9.81 m/s2 (i.e. az = 0, refer to Equation

5.15). Following release at t = 1.1 s the anchor began to free-fall in water with an abrupt

change in Az to -0.81 m/s2 (az = 9 m/s

2). From t = 1.1 to 3.59 s the anchor was in free-

fall through water and Az (and hence az) steadily reduced as the fluid drag resistance

increased with increasing anchor velocity. Impact with the mudline occurred at t = 3.59

s and is characterised by a rapid deceleration to a maximum value of approximately Az

= -41.6 m/s2 (az = -3.2g = -31.8 m/s

2). The anchor came to rest at t = 4.2 s before

rebounding slightly. This rebound has been reported in other studies involving free-fall

objects (e.g. Dayal and Allen 1973, Chow and Airey 2010, Morton and O’Loughlin

2012, O’Loughlin et al. 2014a), and is attributed to elastic rebound of the soil. The

importance of transforming the measured accelerations to the inertial frame using the

DCM is evident from the soil penetration phase where the magnitude of the peak

inertial frame deceleration, Az, is 3.7% lower than the peak body frame deceleration,

Abz. Furthermore, when the anchor was at rest the inertial frame accelerations Ax and

Ay, sensibly returned to zero and Az = -9.81 m/s2 (aiz = 0) in the absence of linear

acceleration, whereas the body frame accelerations, Abx and Aby are non-zero, and Abz ≠

-9.81 m/s2 due to anchor rotations causing misalignment between the body and inertial

frames.

The acceleration response of the IFFS (Figure 5.15b) is broadly similar to that of the

DEPLA, with the expected change in acceleration upon release and the subsequent

Page 176: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

142

reduction in acceleration due to increasing fluid drag resistance. Accelerations also

reduce markedly upon impact with the soil surface, although the absolute deceleration is

lower than for the DEPLA due to the lower soil strength at this site. The sudden

reduction in the accelerations along the z-axis during penetration in soil (evident in both

the body frame and the inertial frame accelerations) is considered to be due to changes

in the soil flow regime. This influences the magnitude of the drag resistance that

dominates at these very shallow embedment depths in very soft soil and at high

penetration velocities (Morton et al. 2015).

Figure 5.15c shows the acceleration response for the DPA test. The response is

qualitatively similar to those shown in Figure 5.15a and Figure 5.15b for the DEPLA

and the IFFS respectively, although there is negligible difference between the body

frame accelerations and the transformed inertial frame accelerations as rotations were

relatively small for this test.

Page 177: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

143

(a)

(b)

(c)

Figure 5.15. Projectile accelerations during free-fall through water and soil

penetration: (a) DEPLA, (b) IFFS and (c) DPA

Page 178: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

144

5.7.3 Velocity profiles

Figure 5.16 shows velocity profiles for free-fall in water and embedment into soil for

the three tests considered previously and shown in Figure 5.14 and Figure 5.15. The

velocity, vz, and distance, sz (i.e. depth), relative to the inertial frame were established

using Equation 5.19 and Equation 5.22. The velocity, vbz, and distance, sbz, were also

derived from Equation 5.19 and Equation 5.22, albeit with, abz = Abz + g, instead of az

and Az. vbz and sbz represent the values that would otherwise be used if the IMU

measurements were not corrected using the AVTM and DCM. The importance of

implementing the transformation matrices is demonstrated in Figure 5.16a where the

final embedment depth and impact velocity of the DEPLA are over estimated by 12%

and 7% respectively. This would correspond to an over prediction of the local undrained

shear strength (and hence capacity) at the mid-height of the DEPLA plate (following

installation but prior to keying) of 17% based on the final tip embedment of ze = 3.31 m

and the idealised strength profile, su (kPa) = 2 +2.8zz. Figure 5.16b indicates that the

embedment depth and impact velocity of the IFFS are over predicted by 27% and 10%

respectively. The over prediction for the IFFS is higher than for the DEPLA as the IFFS

rotations are higher (i.e. greater misalignment between the body- and inertial frames).

Figure 5.16a and Figure 5.16b also show that the velocity, vbz, established from the

integration of the body frame ‘linear’ acceleration, abz, does not return to zero despite

motion having ceased. This is because the body frame acceleration measurement, Abz,

(from which abz is derived) is not coincident with the inertial frame and does not return

to zero following installation (i.e. Abz > -9.81 m/s2). The DPA body frame and inertial

frame velocity profiles (Figure Figure 5.16c) are in excellent agreement as the rotations

are relatively low and the misalignment between the body frame and inertial frame is

negligible. Also shown on Figure 5.16 are direct measurements of the final embedment

depths based on mudline observations of markings on the retrieval line using a ROV in

Page 179: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

145

Firth of Clyde and an underwater drop camera in Lough Erne. Final embedment depths

derived from the IMU data are within 3.3% of the direct measurements, with differences

of Δze = 0.09 m (DEPLA), 0.06 m (IFFS) and 0.035 m (DPA). However, the direct

measurements are simply to confirm the lack of any gross error in the analysis, and have

a much lower accuracy than is possible from the IMU data. A more rigorous verification

of the IMU derived measurements was undertaken for a number of tests as described in

the following section.

5.7.4 Verification of the IMU derived measurements

Independent verification of the IMU-derived measurements of the projectile

displacement (Equation 5.22) was obtained by comparison with those obtained from a

draw wire sensor (also known as a string potentiometer) with a 10 m measurement

range. The draw wire sensor was connected between a fixed point on the deployment

platform and the free falling projectile (i.e. in parallel with the deployment and retrieval

line), and the data acquired using an independent 24-bit data acquisition system. Five

tests were undertaken using the IFFS projectile released from 0 to 4.8 m above the

lakebed.

Comparisons of displacements derived from the IMU measurements and the draw wire

sensor data are provided in Figure 5.17. The IMU-derived displacements are shown

both using the body reference frame and the inertial reference frame. This shows that

the inertial frame-derived displacements correctly remain constant when the projectile

comes to rest in the soil. In contrast, the body frame-derived displacements continue to

increase as the resultant linear acceleration, a, has not returned to zero due to the

rotation of the body (see also Figure 5.16). Importantly, excellent agreement is apparent

between the inertial frame displacements and those measured by the draw wire sensor

(within 1% of the measurement range), providing verification of the analysis approach

outlined here.

Page 180: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

146

(a)

(b)

(c)

Figure 5.16. Projectile velocity profiles corresponding to free-fall through water

and soil penetration: (a) DEPLA, (b) IFFS and (c) DPA

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14

25

20

15

10

5

0

vbz

= 10.9 m/s

vz = 10.2 m/s

sbz

3.72 m

sbz

= 18.03 m

Inertial frame

Body frame

Lakebed

sz = 3.31 m

Velocity, vbz

and vz (m/s)

Dis

tan

ce,

s bz

and

sz

(m)

sz = 17.69 m

direct measurement from lakebed = 3.4 m

0 1 2 3 4 5 6 7 8

10

9

8

7

6

5

4

3

2

1

0

vbz

= 7.4 m/s

vz = 6.7 m/s

sbz

2.66 m

sbz

= 6.26 m

Inertial frame

Body frame

Lakebed

sz = 2.10 m

Velocity, vbz

and vz (m/s)

Dis

tan

ce,

s bz

and

sz

(m)

sz = 5.95 m

direct measurment from lakebed = 2.17 m

0 1 2 3 4 5 6 7 8

6

5

4

3

2

1

0

sz s

bz = 1.72 m

vz v

bz = 6 m/s

Inertial frame

Body frame

Lakebed

Velocity, vbz

and vz (m/s)

Dis

tan

ce,

s bz a

nd

sz (

m)

sz s

bz = 3 m

direct measurment from lakebed = 1.755 m

Page 181: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

147

Figure 5.17. Comparison of IMU derived displacement measurements with those

obtained using a draw wire sensor

5.7.5 Example application of projectile IMU data

For the projectiles considered in the previous section, understanding the soil-structure

interaction at such high strain rates is crucial for predictive tools that calculate the final

embedment depth of the anchors (DEPLA and DPA, e.g. O’Loughlin et al. 2013b) or

estimate the undrained shear strength based on the interpreted inertial frame

accelerations (IFFS, e.g. O’Loughlin et al. 2014a, IFFS, Morton et al. 2015). This is

because those strain rates are up to seven orders of magnitude higher than used for

strength determination in a standard laboratory element test. It follows that motion data

such as those presented in Figure 5.14 and Figure 5.15 play an important role in the

validation and calibration of such predictive models. An example comparison is

provided in Figure 5.18. for the DEPLA, where the predictions are based on an

analytical model described in brief here, but in more detail by O’Loughlin et al.

(2013b). The model formulates conventional end bearing and frictional resistance acting

on the anchor during penetration in a manner similar to suction caisson or pile

installation, but scales these resistances to account for the well-known dependence of

undrained shear strength on strain rate (Casagrande and Wilson 1951, Graham et al.

1983, Sheahan et al. 1996), whilst also accounting for fluid drag resistance and the

Page 182: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

148

buoyant weight of the displaced soil. Consideration of these resistance components

leads to the following governing equation:

dbearfrictfbs FFFRFW

dt

sdm )(

2

2

Equation 5.26

where m is the anchor mass, s is the distance travelled by the projectile, t is time, Ws is

the submerged weight of the anchor in water, Ffrict is frictional resistance, Fbear is bearing

resistance, Fb is the buoyant weight of the displaced soil and Fd is fluid drag resistance,

formulated as:

2

2

1vACF fsd Equation 5.27

where ρs is the submerged density of the soil, Cd is the fluid drag coefficient, Af is

anchor frontal area and v is the instantaneous resultant anchor velocity. The inclusion of

fluid drag, Fd, is essential in situations where (non-Newtonian) very soft fluidised soil is

encountered at the surface of the seabed, and has been shown to be important for

assessing loading from a submarine slide runout on a pipeline (Boukpeti et al. 2012,

Randolph and White 2012, Sahdi et al. 2014). O’Loughlin et al. (2013b) and Blake and

O’Loughlin (2015) further showed that inertial drag is the dominant resistance acting on

a dynamically installed anchor in normally consolidated clay during initial embedment

and typically to about 30% of the penetration.

Frictional and bearing resistances are formulated as:

sufrict AsF Equation 5.28

fucbear AsNF Equation 5.29

Page 183: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

149

where α is an interface friction ratio (of limiting shear stress to undrained shear

strength), As is anchor shaft area, Nc is the bearing capacity factor for the projectile tip

or fluke, and su is the undrained shear strength averaged over the contact area, Af or As.

The reference undrained shear strength adopted in Equation 5.28 and Equation 5.29 is

the idealised profile shown in Figure 5.9a, which is enhanced using a power law strain

rate function (Biscontin and Pestena 2001, Peuchan and Mayne 2007, Randolph et al.

2007, O’Loughlin et al. 2013b) expressed as:

refref

fdv

dvnR

Equation 5.30

where β is the strain rate parameter, v/d is an approximation of the operational shear

strain rate, and the subscript ‘ref’ denotes the reference shear strain rate (v/Dball = 0.18 s-

1) associated with the measurement of the undrained shear strength. The factor n in

Equation 5.30 accounts for the greater rate effects reported for shaft resistance

compared to tip resistance (Dayal et al. 1975, Chow et al. 2014, Steiner et al. 2014) and

is taken as n = 1 for tip resistance (Zhu and Randolph 2011) and as a function of β

(adopted from Equation 8b in Einav and Randolph 2006) for estimating rate effects in

shaft resistance according to:

22 l

l nβ

nn Equation 5.31

where nl is for axial loading.

The predictions Figure 5.18 on were obtained using bearing capacity factors of Nc = 7.5

for the leading and trailing edges of the flukes (analogous to a deeply embedded strip

footing) and Nc = 12 for the follower tip, but not for the padeye as the hole formed by

the passage of the anchor was assumed to remain open. This is appropriate since ROV

Page 184: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

150

video capture of the drop sites (see Figure 5.13) showed an open crater and the

dimensionless strength ratio at the trailing end of the embedded DEPLA follower, su/γ'd

= 6.9 (where d is the diameter of the DEPLA sleeve and γ' is the effective unit weight of

the soil), which is sufficient to maintain an open cavity above the follower (Morton et

al. 2014). Values for the drag coefficient, Cd, were determined from the free-fall in

water phase of the tests, which gave an average Cd = 0.7 (Blake and O’Loughlin 2015).

The strain rate parameter was taken as β = 0.08, which is typical of that measured in

variable rate penetrometer testing (Low et al. 2008, Lehane et al. 2009) and

approximates to an 18% change per log cycle change in strain rate, typical of that

measured in laboratory testing (e.g. Vade and Campenella 1977, Graham et al. 1983,

Lefebvre and Leboeuf 1987). The interface friction ratio, α, was varied to obtain the

best match between the measured and predicted velocity profiles. The comparison

between these on Figure 5.18 indicates that the inclusion of a fluid-mechanics drag

resistance term is appropriate for projectiles penetrating soft clay at high velocities.

There is excellent agreement between the measured and predicted velocity profiles

using α = 0.27, which is within the range deduced from the cyclic piezoball remoulding

tests (0.19 to 0.33). In contrast, the best agreement that could be obtained without the

inclusion of inertial drag resistance required α = 0.38, which is inconsistent with results

from the cyclic piezoball remoulding tests and gave a much poorer match.

Page 185: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

151

Figure 5.18. DEPLA velocity profile derived from the IMU data measured at the

Firth of Clyde test site and corresponding theoretical profile

5.8 Conclusions

This chapter describes a fully self-contained low cost MEMS-based IMU consisting of a

tri-axis accelerometer and a three-component gyroscope, and considered sample data

captured by the IMU during field tests on dynamically installed projectiles. Such data

are important for understanding the soil-structure interactions that occur at the elevated

shear strain rates associated with dynamic penetration events. To the authors’

knowledge these data are the first reported use of a 6DoF IMU for a geotechnical

application.

A comprehensive framework for interpreting the IMU measurements so that they are

coincident with a fixed inertial frame of reference was described and implemented to

establish projectile rotations, accelerations and velocities during free-fall in water and

embedment in soil. It is often the final embedment depth of a dynamically embedded

projectile that is of interest. The chapter showed that embedments calculated from the

body frame acceleration measurements, rather than from accelerations transformed to an

inertial frame of reference, led to derived embedment depths that were in error by up to

27%. In contrast, embedment depths derived from IMU data interpreted from within an

0 1 2 3 4 5 6 7 8 9 10 11 12

4.0

3.5

3.0

2.5

2.0

1.5

1.0

0.5

0.0

IMU data

Cd = 0.7, = 0.27

Cd = 0, = 0.38

Resultant velocity, v (m/s)

Dis

tan

ce, s

(m)

Page 186: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

152

inertial frame of reference were shown to be in excellent agreement with independent

direct measurements.

The merit of collecting motion data during dynamic penetration events was

demonstrated by using the IMU data to validate an embedment prediction model based

on strain rate enhanced shear resistance and fluid mechanics drag resistance for

dynamically installed anchors. In this demonstration the inclusion of drag resistance

during embedment in soil was shown to be appropriate, as the measured and predicted

velocity profiles were in excellent agreement. In contrast, when drag resistance was

omitted an interface friction ratio inconsistent with the measured soil sensitivity was

required to match the final embedment depth, and as a consequence the overall

agreement between the measured and predicted profiles was much poorer.

In conclusion, the use of a reliable IMU with an appropriate interpretation framework is

required to successfully apply these projectile-based geotechnical devices.

Page 187: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

153

CHAPTER 6 INSTALLATION OF DYNAMICALLY

EMBEDDED PLATE ANCHORS AS

ASSESSED THROUGH FIELD TESTS

6.1 Abstract

A dynamically embedded plate anchor (DEPLA) is a rocket shaped anchor that

comprises a removable central shaft and a set of four flukes. The DEPLA penetrates to a

target depth in the seabed by the kinetic energy obtained through free-fall in water.

After embedment the central shaft is retrieved leaving the anchor flukes vertically

embedded in the seabed. The flukes constitute the load bearing element as a plate

anchor.

This chapter focuses on the dynamic installation of the DEPLA. Net resistance and

velocity profiles are derived from acceleration data measured by an inertial

measurement unit during DEPLA field tests, which are compared with corresponding

theoretical profiles based on strain-rate enhanced shear resistance and fluid mechanics

drag resistance. Comparison of the measured net resistance force profiles with the

model predictions shows fair agreement at 1:12 scale and good agreement at 1:7.2 and

1:4.5 scales. For all scales the embedment model predicts the final anchor embedment

depth to a high degree of accuracy.

6.2 Introduction

A new hybrid anchoring system referred to as the Dynamically Embedded PLate

Anchor (DEPLA), combines the capacity advantages of vertically loaded anchors

(Murff et al. 2005, Gaudin et al. 2006, Wong et al. 2012) with the installation

advantages of dynamically installed anchors (Medeiros et al. 2001, Zimmerman et al.

2009, Lieng et al. 2010). The DEPLA comprises a removable central shaft or ‘follower’

Page 188: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

154

that may be fully or partially solid and a set of four flukes (see Figure 1.17). A stop cap

at the upper end of the follower prevents the follower from falling through the DEPLA

sleeve and a shear pin connects the flukes to the follower. After release from a

designated height above the seabed (typically <100 metres), The DEPLA penetrates to a

target depth in the seabed by the kinetic energy obtained through free-fall and the self-

weight of the anchor. After embedment the follower line is tensioned, which causes the

shear pin to part (if not already broken during impact) allowing the follower to be

retrieved for the next installation, whilst leaving the anchor flukes vertically embedded

in the seabed. These embedded anchor flukes constitute the load bearing element as a

plate anchor. A mooring line attached to the embedded flukes is then tensioned, causing

the flukes to rotate or ‘key’ to an orientation that is perpendicular to the direction of

loading, ensuring that maximum anchor capacity is achieved through bearing resistance.

The installation and keying process are summarised in Figure 1.18 and Figure 1.19

respectively.

The holding capacity of a plate anchor such as the DEPLA is relatively straightforward

to determine for a given shear strength profile and a known embedment depth (Cassidy

et al. 2012). However predicting the final plate anchor embedment is challenging as it

firstly requires an assessment of the final DEPLA penetration depth after free-fall in

water, and secondly requires an assessment of the extent of embedment loss of the plate

anchor keying. The loss in embedment during keying has been investigated

experimentally for suction embedded plate anchors (e.g. O’Loughlin et al. 2006, Gaudin

et al. 2006, Gaudin et al. 2009, Wang et al. 2011, Cassidy et al. 2012, Yang et al. 2012),

the results of which are relevant to the DEPLA. Predicting the final embedment of the

DEPLA after free-fall is complicated by the very high penetration velocities, as this

significantly enhances the available shear resistance and for very soft normally

consolidated clay, introduces fluid drag resistance and inertia effects. As the velocity of

Page 189: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

155

dynamically installed anchors when they impact the seabed is typically between 10 and

30 m/s (Lieng et al. 2010), for a typical anchor with a shaft diameter of about 1 m, the

strain rates in the soil during installation are three orders of magnitude higher than that

in vane test (0.029 s-1

, Einav and Randolph 2006) and seven orders of magnitude higher

than the standard laboratory testing rate of 1%/hour (2.8 × 10-6

s-1). O’Loughlin et al.

(2013b) showed that, for anchor geometries similar to the DEPLA in normally

consolidated clay, fluid drag resistance dominates for embedment depths up to 0.84

times the anchor length. The importance of combining fluid mechanics drag resistance

with soil mechanics strain rate enhanced shear resistance has also been demonstrated for

other applications involving high displacement velocities (e.g. Zhu and Randolph 2011,

Randolph and White 2012, Randolph 2013, Sahdi et al. 2014). Failure to account for

strain rate enhanced shear resistance and fluid mechanics drag resistance during

dynamic embedment of anchors in clay has been shown to give prediction errors of the

order of 40% (O’Loughlin et al. 2004a).

A series of field tests were designed to provide a basis for understanding the behaviour

of the DEPLA during (i) free-fall and dynamic embedment in soil, and (ii) quantifying

the loss in embedment during keying and the subsequent plate anchor capacity. This

chapter considers the installation aspects of the field tests, whereas keying and anchor

capacity aspects of the field tests are considered in Chapter 7.

6.3 Experimental program

6.3.1 Instrumented model anchors

The DEPLA was modelled at reduced scales of 1:12, 1:7.2 and 1:4.5, and fabricated

from mild steel. The reduced scale infers a full scale anchor (or follower) length of 9 m,

which for a typical offshore seabed strength profile of 1.5 kPa/m would result in a plate

anchor capacity of the order of about 500 tonnes, suitable for temporary drilling

Page 190: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

156

operations. The DEPLA follower was fully solid with the exception of a cylindrical

void at the top of the follower to house the inertial measurement unit (described later).

A schematic of the model DEPLA and the notation used for the main dimensions are

shown in Figure 6.1. The dimensions and mass of each DEPLA are summarised in

Table 6.1. 12 mm diameter Dyneema SK75 was used for deploying the DEPLA (and

hence retrieving the follower) and as the mooring line. This rope was selected because

of its flexibility (compared with wire rope), high tensile strength and stiffness (<3.5%

extension at 50% of the maximum breaking load).

Figure 6.1. DEPLA dimensions

Ds

Hs

Lf

Df

Ltip

D

t

e

Page 191: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

157

Table 6.1. DEPLA anchors

1:12 scale 1:7.2 scale 1:4.5 scale

Follower length, Lf (mm) 750 1250 2000

Tip length, Ltip (mm) 50 83.3 133.3

Follower diameter, Df (mm) 60 100 160

Sleeve diameter, Ds (mm) 71 115 184

Sleeve height, Hs (mm) 290 487 779

Plate diameter, D (mm) 300 500 800

Fluke thickness, t (mm) 4.5 6.25 10

Padeye eccentricity, e (mm) 125 212.5 348

Eccentricity ratio, e/D (-) 0.42 0.43 0.44

Follower mass (kg) 14.8 73.7 297

Plate (flukes) mass (kg) 5.5 22.3 91.6

Total mass (kg) 20.3 96 388.6

6.3.2 Inertial Measurement Unit

An inertial measurement unit (IMU) housed within a void at the top of the DEPLA

follower was used to measure anchor acceleration and rotation rates during free-fall in

the water column and embedment in the soil. The IMU (see Figure 5.5) included two

micro-electro mechanical systems (MEMS) sensors; a 13 bit three-axis accelerometer

(ADXL 345) and a 16 bit three component rate gyroscope sensor (ITG 3200). The

accelerometer had a resolution of 0.04 m/s 2 with a measurement range of +/-16 g. The

rate gyroscope sensor had a resolution of 0.07 °/s with a measurement range of +/-

2000°/s. Data was logged by an mbed micro controller with an ARM processor to a 2

GB SD card at 400 Hz. Internal batteries were capable of powering the logger for up to

4 hours. The IMU was contained in a watertight aluminium tube 185 mm long and 42

mm in diameter and located in a void (with the same dimensions) at the top of the

follower. The accelerometer and rate gyroscope are aligned with the body-frame of the

DEPLA and the IMU as shown in Figure 5.6. The body frame is a reference frame with

three orthogonal axes xb, yb and zb that are common to both the IMU and the DEPLA

and where the zb-axis is parallel to the direction of earth’s gravity when the DEPLA is

Page 192: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

158

hanging vertically. The rate gyroscope measures angular velocities ωbx, ωby and ωbz in

the body-frame about three orthogonal axes xb, yb and zb. The accelerometer measures

accelerations Abx, Aby and Abz in the body-frame along the same orthogonal axes. These

accelerometer measurements include a component of gravitational acceleration

(depending on the orientation of the accelerometer) and linear acceleration.

Accelerometers are often used to measure the rotation of quasi-static objects but cannot

distinguish rotation from linear acceleration if an object is in motion. However

gyroscopes are unaffected by linear acceleration and the rotation of objects in motion

can be derived from their measurements. If the DEPLA pitches and/or rolls whilst in

motion the measured angular velocities and accelerations will correspond to a reference

frame that is not fixed in space. As a consequence the gravitational and linear

acceleration components cannot be distinguished from the accelerometer measurements.

Hence it is important to define an inertial frame of reference, and to transform the body

frame acceleration measurements Abx, Aby and Abz to accelerations that are coincident

with this inertial frame Ax, Ay and Az using rotation matrices as described in detail by

Blake et al. (2016). The inertial frame used here is a local fixed reference frame, with

the z-axis aligned in the direction of the Earth’s gravitational vector, and with undefined

orthogonal x- and y-axes, that are fixed at their orientation at the start of each test. The

linear accelerations ax, ay and az coincident with the inertial frame, which are required to

calculate velocity and displacement components (including the vertical embedment

depth of the anchor relative to the soil surface, ze, and the distance travelled by the

anchor along its direction of motion), can then be calculated using the following

expression:

gA

A

A

a

a

a

z

y

x

z

y

x

0

0

Equation 6.1

Page 193: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

159

where g is gravitational acceleration and Az is an negative output (i.e. when the anchor

is at rest az = Az + g = 0).

6.3.3 Test site and soil properties

The tests reported here were conducted in Lower Lough Erne, a glacial lake in Co.

Fermanagh, Northern Ireland in water depths between 3 and 12 m. The lakebed is very

soft clay with moisture contents in the range 270 to 520%, typically about 1.5 times the

liquid limit. The measured unit weight of the Lough Erne clay is only marginally higher

than water at γ = 10.8 kN/m3. This is considered to be due to the very high proportion of

diatoms that are evident from scanning electron microscopic images of the soil (e.g. see

Colreavy et al. 2012) and which have an enormous capacity to hold water in the

intraskeletal pore space (Tanaka and Locat 1999). However the water that rests within

this pore space is not considered to play a role in soil behaviour, and as such the

measured unit weight and other index properties that are expressed in terms of the

measured moisture content are not considered to be useful indicators of soil behaviour.

Colreavy et al. (2012) report results from piezoball and in situ shear vane tests

conducted at the test site to depths of up to 11 m. Figure 6.2 shows profiles of undrained

shear strength, su, with depth where su is derived from the net penetration resistance

using Nball = 8.6 (calibrated using in situ shear vane data). As will be shown later, the

1:12 scale anchor achieved maximum tip embedment depths of 1.76 m. Over this depth

su is best idealised as su = 1.5z (kPa), where z is depth, and this profile was adopted in

the analysis of the test data using this anchor. The 1:7.2 and the 1:4.5 scale anchors

achieved maximum tip embedment depths of 3.66 m and 6.94 m respectively. Over this

range of embedment su may be idealised as su = 0.45 + 0.9z (kPa) and this profile was

adopted in the analysis of the test data using the 1:7.2 and the 1:4.5 scale anchors. The

sensitivity of the soil is in the narrow range St = 2 to 2.5 as assessed from in situ vane

tests and piezoball cyclic remoulding tests.

Page 194: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

160

Figure 6.2. Typical undrained shear strength profiles

6.3.4 Testing setup and procedure

Tests were carried out from a fixed jetty (1:12 scale DEPLA only) and from the deck of

a 15 m self-propelled barge with a 13 t winch and 2 t crane (Figure 6.3). The testing

procedure (summarised schematically for the barge tests in Figure 6.4) was as follows:

1. The IMU was powered up and secured in the follower.

2. The DEPLA follower and flukes were attached using a shear pin and lowered below

the water surface to the desired drop height above the lakebed (see Figure 6.5).

0 1 2 3 4 5 6 7 8

8

7

6

5

4

3

2

1

0

su = 1.5z

Ball 1

Ball 2

Ball 3

Ball 4

Vane

Undrained shear strength, su (kPa)

Pen

etra

tio

n d

epth

, z

(m)

su = 0.45 + 0.9z

Page 195: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

161

3. The DEPLA was released by opening a quick release shackle connecting the

follower retrieval line to the crane, allowing the anchor to free-fall and penetrate the

lakebed.

4. The anchor tip embedment depth, ze, was measured by sending a camera to the

lakebed to inspect markings on the follower retrieval line (see Figure 6.6).

Figure 6.3. Self-propelled barge

Page 196: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

162

Figure 6.4. (a), (b) and (c). DEPLA field test procedure

hiz

, dro

p

z wat

er

Barge

Winch

Pulley

Water

Lakebed

(a)

Follower

line

DEPLA

Quick

release

shackle

Crane

Plate lineWat

er d

epth

Dro

p h

eigh

t

Winch

Pulley

Water

Lakebed

(b)

Follower

line

Barge

Crane

Plate line

Winch

Pulley

Water

Lakebed

(c)

Follower

line

Camera cable

Camera

Barge

Crane

Plate line

z iz,e

z e

Page 197: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

163

Figure 6.5. 1:7.2 scale DEPLA being lowered to drop height by the barge crane

Figure 6.6. Image capture from under-water camera showing the follower retrieval

line at lakebed

Page 198: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

164

6.4 Results and discussion

6.4.1 Accelerations

Figure 6.7 shows a typical acceleration during free-fall in water and embedment in the

lakebed for an installation using a 1:12 scale DEPLA released from a drop height of

6.86 m. In Figure 6.7 Abz is the accelerometer measurement and Az is the transformed

accelerometer measurement that is coincident with the inertial frame (i.e. Az is the

acceleration measurement in the direction of gravity). The DEPLA was initially hanging

in the water experiencing only gravitational acceleration with Az = -9.81 m/s2 (i.e. az =

0, refer to Equation 6.1). Following release at t = 1 s the anchor began to free-fall in

water with an abrupt change in Az to -0.92 m/s2 (a = 8.9 m/s

2). From t = 1 to 2.55 s the

anchor was in free-fall through water and Az (and hence az) steadily reduced as the fluid

drag resistance increased with increasing anchor velocity. Impact with the lakebed

occurred at t = 2.55 s and is characterised by a rapid deceleration to a maximum value

of approximately Az = -27.4 m/s2 (az = -1.8g = -17.6 m/s

2). The anchor came to rest at t

= 3.16 s before rebounding slightly. This rebound has been reported in other studies

involving free-fall objects (e.g. Dayal and Allen 1973, Chow and Airey 2010, Morton

and O’Loughlin 2012, O’Loughlin et al. 2014a), and is attributed to elastic rebound of

the soil. The importance of transforming the measured accelerations to the inertial frame

using the DCM described by Blake et al. (2016) is evident from the soil penetration

phase where the magnitude of the peak inertial frame deceleration Az, is 2.8% lower

than the peak body frame deceleration Abz. Furthermore, when the anchor was at rest

the inertial frame acceleration Az = -9.81 m/s2 (az = 0) in the absence of linear

acceleration, whereas the body frame acceleration Abz ≠ -9.81 m/s2 due to anchor

rotations causing misalignment between the body and inertial frames.

Page 199: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

165

Figure 6.7. Accelerations during free-fall through water and soil penetration (1:12

scale)

6.4.2 Free-fall in water

The linear accelerations corresponding to the inertial-frame ax, ay and az were

numerically integrated to establish the anchor velocities coincident with the inertial-

frame vx, vy and vz during free-fall in the water column and embedment in the soil:

t

xixx dttavtv00

)()( Equation 6.2

t

yyy dttavtv00

)()( Equation 6.3

t

zzz dttavtv00

)()( Equation 6.4

The resultant anchor velocity, v (velocity in the direction of motion), was calculated

using the following expression:

222

zyx vvvv Equation 6.5

The resultant anchor velocity, v, was numerically integrated to establish the distance

travelled by the anchor along its trajectory, s:

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0-35

-30

-25

-20

-15

-10

-5

0

penetration

Abz

or abz

Az or a

z

At restFree-fall in water

Acc

eler

om

eter

mea

sure

men

t, A

(m

/s2)

Time, t (s)

Hanging Soil

-25

-20

-15

-10

-5

0

5

Lin

ear

acce

lera

tion, a

= A

+ g

(m

/s2)

Page 200: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

166

t

dttvsts0

0 )()( Equation 6.6

The distance travelled by the projectile along the inertial z-axis, sz (required to calculate

the embedment depth of the anchor relative to the soil surface, ze) was established by

numerically integrating the vertical velocity, vz:

t

zzz dttvsts0

0 )()( Equation 6.7

The motion of the anchor during free-fall in water can be quantified by considering

Newton’s second law of motion and the forces acting on the anchor during free-fall

(Fernandes et al. 2006, Shelton et al. 2011, Hasanloo et al. 2012, Hossain et al. 2014):

ldads FFW

dt

sdmm ,,2

2

)'( Equation 6.8

where m is the anchor mass, m' is the added mass, z is the distance travelled by the

anchor along its trajectory, t is time, Ws is the component of the submerged anchor

weight of the anchor acting in the direction of motion, Fd,a is the fluid drag resistance

acting on the anchor and Fd,l is the fluid drag resistance acting on the mooring and

follower recovery lines. The added mass, m', in Equation 6.8 is the mass of the fluid that

is accelerated with the anchor. However for slender bodies such as the DEPLA, m' is

negligible and can be taken as zero (Beard 1981, Shelton et al. 2011).

As an object moves through a viscous fluid it is subjected to fluid drag resistance which

is attributed to; (i) friction drag (commonly referred to viscous drag) generated by

viscous stresses that develop in the boundary layer between the object and fluid, and (ii)

pressure drag (independent of viscous effects) caused by an adverse pressure gradient

(related to the stagnation pressure, Pstagnation = Pstatic + 1/2ρv2) between the front and rear

Page 201: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

167

of the object which creates a force opposite to the direction of motion (Hoerner 1965).

The fluid drag resistance terms are formulated as (Fernandes et al. 2006):

2

,,2

1vACF fwadad Equation 6.9

2

,,2

1vACF swldld Equation 6.10

where Cd,a is the drag coefficient of the anchor, Cd,l is the drag coefficient of the

mooring and follower recovery lines, ρw is the density of the water, Af is the frontal area

(of the follower, sleeve and flukes), As is the surface area of the mooring and follower

recovery lines in contact with water and v is the anchor (and hence line) velocity. In

cases where pressure drag (e.g. dynamically installed anchors such as the DEPLA), it is

common practice to use the frontal area, Af, to calculate the fluid drag resistance (e.g.

Fernandes et al. 2006). Conversely, as the drag acting on the mooring line is purely

frictional, the surface area, As, is adopted for the calculation of drag on the mooring

lines.

Figure 6.8 shows the dependence of anchor drag coefficient, Cd,a, on Reynolds number,

Re, for each anchor scale based on back analyses of the accelerations and velocities

derived from the IMU measurements using the following formulations:

2

2

,

,21

21

vA

vACmaWC

fw

swlds

ad

Equation 6.11

effwvdRe Equation 6.12

Page 202: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

168

where deff is the effective anchor diameter (deff = (4Af/π)0.5

), μ is the dynamic viscosity

of water taken as 1.307x10-3

Pa s at 10°C (Kestin et al. 1978) and a is the resultant

acceleration (acceleration in the direction of motion):

222

zyx aaaa Equation 6.13

The maximum Reynolds numbers achieved during free-fall in the water column are Re

= 0.33x106 (v = 4.9 m/s, 1:12 scale), Re = 0.68x10

6 (v = 6.3 m/s, 1:7.2 scale) and Re =

1.73x106 (v = 10.2 m/s, 1:4.5 scale). Figure 6.8 indicates that fluid drag resistance

acting on the anchor 1:4.5 scale anchor during free-fall in water may be approximated

using a constant mean drag coefficient of Cd,a = 0.7 for Re > 1.0x106 (v > 5.9 m/s). The

mean constant drag coefficient is representative of the fluid drag resistance at high

Reynolds numbers where the pressure drag component is dominant and the viscous drag

component is negligible. The 1:12 and 1:7.2 scale anchors did not reach Reynolds

numbers high enough to allow the constant mean drag coefficient to be back-analysed.

However, since the 1:12, 1:7.2 and 1:4.5 scale anchors are geometrically identical, Cd,a

= 0.7 may be inferred for all scales. This drag coefficient is at the upper range of values

typically reported for dynamically installed anchors and free-fall penetrometers; e.g. Cd

= 0.7 (True 1976), Cd,a = 0.63 (Øye 2000). The geometry of the mooring and follower

recovery lines is analogous to an ‘infinite cylinder’ and as such the fluid drag resistance

can be attributed exclusively to viscous effects (i.e. no pressure drag component). Due

to viscous effects Cd,l will vary at relatively low Re and become constant at higher Re.

However, it is not possible to isolate the viscous drag resistance through back analysis

of the IMU data. Instead, a constant Cd,l value is adopted which should reflect the fluid

drag resistance of the lines at high Re (and hence velocities). In the case of the anchor

penetrating into the lakebed following free-fall (considered later) the constant Cd,l value

should best represent the viscous drag of the lines during initial penetration where it is

Page 203: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

169

most prevalent. Back analysis of the IMU data suggest a Cd,l of 0.015 was necessary to

predict the increase in fluid drag resistance due to the paying out of the mooring and

follower recovery lines during installation.

Figure 6.8. Dependence of DEPLA drag coefficient on Reynolds number

Figure 6.9 shows typical velocity profiles for the 1:12, 1:7.2 and 1:4.5 anchor scales

during free-fall in water together with theoretical profiles based on a finite difference

approximation of Equation 6.8 using Cd,l = 0.015 and Cd,a = 0.7. This provides a good fit

to the 1:4.5 scale anchor velocity profiles. There is relatively poor agreement between

the theoretical and measured velocity profiles for the 1:12 and 1:7.2 anchor scales as the

constant mean coefficient does not account for the viscous component of fluid drag

resistance which dominates at low Reynolds numbers and hence velocities. However, as

described later, the dynamic penetration of the DEPLA in soil can be predicted by

considering a fluid drag resistance term with a constant mean drag coefficient (in the

form of Equation 6.9), and a soil mechanics term that accounts for the strain rate

enhanced shear resistance attributed to viscous effects.

0.0 5.0x105

1.0x106

1.5x106

2.0x106

0

1

2

3

4

5

1:12 scale 1:7.2 scale 1:4.5 scale

Cd,a

= 0.7

Re = 0.33x106

(v = 4.9 m/s, 1:12 scale)

Re = 1.73x106

(v = 10.2 m/s, 1.4.5 scale)

Reynolds number, Re

Re = 0.68x106

(v = 6.3 m/s, 1:7.2 scale)

An

cho

r d

rag

co

effi

cien

t, C

d,a

Page 204: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

170

Impact velocities (resultant velocity at point of contact with the lakebed) were in the

ranges: vi = 4.9 to 6.4 m/s (1:12 scale), vi = 2.7 to 8.2 m/s (1:7.2 scale) and vi = 8.5 to

10.5 m/s (1:4.5 scale) and were achieved for anchor drop heights in the ranges: hiz,drop =

3 to 8.2 m (1:12 scale), 0.5 to 6.2 m (1:7.2 scale), 6.1 to 9.1 m (1:4.5 scale). For the

larger drop heights (1:12 scale only) the DEPLA reached terminal velocity but then

slowed slightly due to the increasing fluid drag resistance afforded by the increasing

length of mooring and follower recovery lines that mobilised in the water.

Figure 6.9. Examples of measured and theoretical velocity profiles in water

6.4.3 Soil penetration

DEPLA tip embedment, ze (vertical distance travelled by the anchor relative to the soil

surface) was in the ranges: 1.6 to 1.8 m (2.1 to 2.4Lf, 1:12 scale), 2.8 to 3.7 m (2.2 to

2.9Lf, 1:7.2 scale) and 6.4 to 6.9 m (3.2 to 3.5Lf, 1:4.5). An additional test on the 1:12

scale DEPLA, in which the anchor was released from the soil surface (i.e. vi = 0)

achieved a tip embedment of 1.2 m (1.6Lf). These tip embedments are in good

agreement with reported field data experience of 1.9 to 2.4 times the anchor length for a

79 t dynamically installed anchor (LF/Df = 10.8), with a broadly similar geometry

0 2 4 6 8 10 12 14 16

12

10

8

6

4

2

0

1:4.5 scale

1:12 scale

Measured

Theoretical, Cd,l

= 0.015, Cd,a

= 0.7

Velocity, v (m/s)

No

rmal

ised

fre

e-fa

ll d

ista

nce

, s/

Lf

1:7.2 scale

Page 205: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

171

(Lieng et al. 2010). The results indicate that anchor embedment increases with

increasing impact velocity, which is sensible as the anchor possesses more kinetic

energy when it impacts the lake bed.

True (1976) proposed a method for predicting the dynamic penetration of projectiles

penetrating soil after free-fall in water. The method considers Newton’s second law of

motion and the net resistance on the projectile during penetration, where the resistance

to motion is due to fluid mechanics drag resistance (analogous to Equation 6.9) and soil

mechanics shear resistance.

Numerous studies on free-falling penetrometers (e.g. Beard 1981, Levacher 1985,

Aubeny and Shi 2006) and dynamically installed anchoring systems (e.g. O’Loughlin et

al. 2004a, Audibert et al. 2006, O’Loughlin et al. 2009, Shelton et al. 2011, Gaudin et

al. 2013, O’Loughlin et al. 2013b) have adopted True’s method for predicting the

embedment of projectiles penetrating the seabed, with slight variations on the inclusion

and formulation of the forces acting on the projectile during penetration. The governing

equation may be expressed as:

ldadbearfrictfbs FFFFRFW

dt

sdm .,2

2

)( Equation 6.14

where ze is the distance travelled by the anchor tip along its trajectory in the soil, Fb is

the buoyant weight of the displaced soil, Ffrict is frictional resistance and Fbear is bearing

resistance. The shear strain-rate function, Rf, is included to account for the well know

dependence of soil strength (and hence frictional and bearing resistance) on strain rate;

this is discussed in more detail later in the chapter. Fd,a and Fd,l account for drag

resistance on the anchor as it moves through the soil, and the lines as they move through

the soil and the water. Rf and Fd both relate to the high penetration velocity of the

anchor, but as described later in the chapter, Rf accounts for viscous effects arising from

Page 206: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

172

the very high strain rate in the soil, whereas Fd accounts for pressure drag which is

independent of viscosity. The added mass term, which is negligible for long slender

bodies and was ignored in Equation 6.8, is also ignored in Equation 6.14.

Frictional and bearing resistances are expressed as:

AsF ufrict Equation 6.15

AsNF ucbear Equation 6.16

where α is a friction ratio (of limiting shear stress to undrained shear strength), Nc is the

bearing capacity factor for the anchor tip, sleeve or fluke and su is the undrained shear

strength averaged over the contact area, A. Reverse end bearing at the top of the

DEPLA flukes and follower (accounted for in Equation 6.16) was treated in a similar

manner to that proposed by O’Loughlin et al. (2013b), where it would only be included

if the cavities formed by the passage of the anchor follower and flukes immediately

close during penetration. The cavity formed by the flukes was assumed to close owing

to their relatively low thickness and apparent plane strain conditions, whereas the

cylindrical cavity created by the passage of the follower would remain open during the

time taken for dynamic penetration, and collapse shortly after. This assumption requires

that the buoyant weight of the displaced soil, Fb, be calculated using a displaced volume

equal to the sum of the DEPLA volume and a cylindrical shaft (with diameter Ds)

extending from the top of the DEPLA follower to the soil surface.

The high penetration velocity of the anchor requires that the dependence of soil strength

on shear strain-rate is accounted for in Equation 6.14. This can be achieved by scaling a

reference value of soil strength using either semi-logarithmic or power functions

(Biscontin and Pestena 2001, Peuchan and Mayne 2007, Randolph et al. 2007 and

O’Loughlin et al. 2013b):

Page 207: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

173

refref

fdv

dvR log1log1

Equation 6.17

refref

fdv

dvR

Equation 6.18

where λ and β are strain-rate parameters in the semi-logarithmic and power functions

respectively, γ is the shear strain rate and refγ is the reference shear strain rate associated

with the reference value of undrained shear strength. As described by Einav and

Randolph (2006) and O’Loughlin et al. (2013b), γ will vary through the soil body, but

at any location the operational strain rate may be approximated by the normalised

velocity, v/d. The approximated reference value of operational strain rate, (v/d)ref, is

that associated with the measurement of the undrained shear strength. As shown on

Figure 6.2, the su measurements were made using a 113 mm piezoball penetrated at 20

mm/s (Colreavy et al. 2012), such that (v/d)ref = 0.18 s-1

, two orders of magnitude lower

that the maximum v/d values in the DEPLA tests.

It is worth noting that the effect of strain-rate on mobilised shear stress can also be

described from within a fluid mechanics framework by treating the soil as a non-

Newtonian fluid and using a viscoplastic model such as the Herschel-Bulkley model

(Zhu and Randolph 2011):

Ky Equation 6.19

where τ is the mobilised shear stress, τy is the yield stress, K represents the viscosity of

the soil (often referred to as the consistency parameter) and δ is the shear thinning

index.

Page 208: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

174

It may be argued that such an approach is appropriate in view of the very soft fluidised

soil encountered at the surface of the seabed. Indeed the loading from a submarine slide

runout on a pipeline is typically assessed using such an approach (Randolph and White

2012). However, as shown by Zhu and Randolph (2011) and Boukpeti et al. (2012), the

Herschel-Bulkley model resorts to the soil mechanics power law formulation (Equation

6.18) when τy = 0, β

refrefu, γsK and δ = β. Hence using either Equation 6.18 or

Equation 6.19 describes the viscous response of the soil attributed to strain rate effects.

However, both equations neglect the pressure drag component of fluid drag resistance

which is independent of viscous effects and related to the stagnation pressure (Zhu and

Randolph 2011, Randolph and White 2012, Sahdi et al. 2014). In the case of submarine

slide runout loading on a pipeline, pressure drag has been observed to dominate at non-

Newtonian Reynolds numbers, Renon-Newtonian > 3 (Sahdi et al. 2014) and Renon-Newtonian >

10 (Zhu and Randolph 2011, Randolph and White 2012), where the non-Newtonian

Reynolds number is defined as:

opu

Newtoniannons

v

,

2

Re

Equation 6.20

where su,op is the operative undrained shear strength.

Hence Renon-Newtonian > 3 to 10 corresponds to combinations of low strength and high

velocities, and as such the pressure drag component of fluid drag resistance needs to be

accounted for separately from the viscous component during dynamic penetration of

dynamically installed anchors such as the DEPLA. Therefore, an appropriate

embedment model for the DEPLA, accounting for effects arising from the high

penetration velocity, is to use Equation 6.18 to account for viscous effects and to use a

constant drag coefficient in Equation 6.9 (i.e. making it independent of viscous effects)

Page 209: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

175

to account for pressure drag on the anchor and to use a constant drag coefficient in

Equation 6.10 to account for viscous effects on the lines.

This ‘superposition’ approach is adopted in a finite difference approximation of

Equation 6.14 to predict the dynamic penetration of the DEPLA in soil. Predictions

were made for each anchor scale assuming idealised strength profiles, su (kPa) = 1.5z

for the 1:12 scale tests, and su (kPa) = 0.45 + 0.9z for the 1:7.2 and 1:4.5 scales (see

Figure 6.2), where z = s cos(μ), i.e. the vertical depth, which will be different from the

distance travelled by the anchor, s, if the anchor tilts by an angle, μ during installation.

Bearing resistance (Equation 6.16) was calculated using Nc = 12 and Nc = 7.5 for the tip

and flukes respectively (O’Loughlin et al. 2004b, O’Loughlin et al. 2009, Richardson et

al. 2009, Jeanjean et al. 2012, O’Loughlin et al. 2013b). The interface friction ratio, α,

was initially taken as the ratio of intact to remoulded penetration resistance from cyclic

piezoball tests, which gave qin/qrem = 0.4 to 0.5. As will be shown later, slight

adjustments beyond the lower and upper bound of this range were required to achieve

good agreement across all three anchor scales. The strain rate parameter, β, in the power

law (Equation 6.18) is typically in the range 0.05 to 0.17 (Jeong et al. 2009) and

increases with the order of magnitude difference between the operational strain rate and

the reference strain rate (Biscontin and Pestana, 2001, Peuchen and Mayne 2007,

O’Loughlin et al. 2013b). The maximum v/d values associated with the DEPLA tests

considered here is v/d = 67 s-1

, which is two orders of magnitude higher than (v/d)ref =

0.18 s-1

. This is similar to the range in v/d associated with variable rate penetrometer

tests, which tends to give β values close to the lower end of the range quoted above, e.g.

β = 0.06 to 0.08 (Lehane et al. 2009). β = 0.06 was adopted here, consistent with β =

0.055, which was shown by Biscontin and Pestana (2001) to satisfactorily describe the

increase in su over three orders of magnitude increase in peripheral shear vane velocity.

Drag resistance was modelled using Cd,a = 0.7 and Cd,l = 0.015 in Equation 6.9 and

Page 210: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

176

Equation 6.10 for the anchor and lines respectively (as deduced from data measured

during free-fall in water, see Figure 6.8). The density of soil rather than water was used

for the anchor, whereas both the density of soil and the density of water were used for

the lines depending on the current calculated anchor embedment depth which

determines if the line drag mobilised in water or soil. Cd,a = 0.7 (derived from the

acceleration data measured during anchor free-fall in water) was used to model anchor

embedment in the soil without adjusted for the viscosity of soil as it is representative of

the fluid drag resistance at high Reynolds numbers where the pressure drag component

is dominant and the viscous drag component is negligible.

Measured and predicted DEPLA velocity profiles are provided on Figure 6.10 for each

anchor scale for impact velocities of 5.6 m/s (1:12 scale), 7.5 m/s (1:7.2 scale) and 9.7

m/s (1:4.5 scale). These velocity profiles consider the anchor displacement in the

direction of motion, rather than anchor penetration depth, as the velocities are calculated

from the resultant acceleration (i.e. in the direction of anchor motion, Equation 6.13).

The predictions were obtained using the parameters introduced in the preceding

discussion, but by varying α to achieve agreement with the measured embedment depth.

This permitted a basis for examining the potential of the model to describe the

mechanisms taking during dynamic embedment. The best agreement was obtained for

each anchor scale using α = 0.46 (1:12 scale), 0.45 (1:7.2 scale) and 0.32 (1:4.5 scale).

These values are generally consistent with the range of intact to remoulded penetration

resistance ratios from cyclic piezoball tests, which gave α = 0.4 to 0.5.

The performance of the embedment model is better assessed from accelerations rather

than velocities as velocity is the integral of accelerations and by definition loses some of

the detail. To this end, Figure 6.10 also shows profiles of the calculated resistance

forces in Equation 6.14 together with the predicted and measured net resistance, Fnet.

Page 211: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

177

The predicted net resistance is the sum of the terms on the right hand side of Equation

6.14, whereas the measured net resistance is taken as the mass of the anchor multiplied

by the resultant acceleration, a.

For the 1:12 anchor scale the fluid drag resistance attributed to the mooring line and

follower recovery lines, Fd,l, is the dominant resistance up to s = 1.2Lf, whereas for the

1:7.2 and 1:4.5 anchor scale the fluid drag resistance acting on the anchor, Fd,a, is the

dominant resistance up to s = 0.8Lf (1:7.2 scale) and 1.2Lf (1:4.5 scale). This relative

reduction in Fd,l with increasing anchor scale is due to the constant line diameter used

for each anchor scale, which leads to reducing ratios of line surface area in contact with

the water, As (contributing to Fd,l) to anchor frontal area, Ap (contributing to Fd,a) as the

anchor scale increases.

The fluid drag resistance is dominant during shallow embedment as the anchor velocity

is high and the undrained shear strength of the soil is low (i.e. for high Renon-Newtonain).

This has also been demonstrated for debris flow on submarine pipelines (Zhu and

Randolph 2011, Randolph and White, 2012 Sahdi et al. 2014). As the anchor penetrates

deeper into the soil the anchor velocity reduces, decreasing the fluid drag resistance of

the anchor, Fd,a, and the mooring and follower recovery lines, Fd,l. At the same time the

strain rate enhanced shearing resistance increases due to the increasing undrained shear

strength and anchor contact area mobilised in the soil. This contact area is low to s =

0.6Lf at which point the flukes start to penetrate the soil. From s = 0.6Lf the circular

fluke geometry causes a non-linear increase in shear resistance until ze = Lf when the

DEPLA is fully submerged in soil and the strain rate enhanced bearing and frictional

resistances are approximately equal. From to s = 1.2Lf (1:12 and 1:4.5 scale) and 0.8Lf

(1:7.2 scale) the strain rate enhanced frictional resistance is the dominant resistance. The

power strain rate law using β = 0.06 increases both the frictional and bearing resistance

Page 212: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

178

by almost 50% (Rf = 1.46, 1:12 scale; Rf = 1.44, 1:7.2 scale; Rf = 1.42, 1:4.5 scale) over

the majority of the embedment. Soil buoyancy, Fb, increases with depth as the hole

formed by the advancing anchor is considered to remain open to the soil surface over

the period of time required for the anchor to come to rest in the soil (typically 0.6 to 1.4

s). Soil buoyancy never exceeds 8% of the strain enhanced bearing resistance since the

effective unit weight of the soil for the Lough Erne clay is very low.

For the 1:7.2 and 1:4.5 anchor scales the net resistance force is initially positive as the

anchor had not reached terminal velocity during free-fall in water and becomes negative

at s = 0.6Lf (1:7.2 scale) and s = 0.8Lf (1:4.5 scale) when sufficient resistance develops

to decelerate the anchors. Over the entire penetration depth good agreement between the

measurements and predictions is obtained for the 1:7.2 scale test and to a slightly lesser

extent the 1:4.5 scale test. Despite the less than perfect agreement between the measured

and predicted net resistance for the 1:12 scale test, the over prediction mainly balances

the under prediction such that the energy dissipated by the anchor during penetration is

approximately the same for both the predictions and the measurements. This has the

effect of predicting the final embedment depth adequately, as shown clearly by Figure

6.11, which shows the measured DEPLA tip embedments against impact velocity

together with the corresponding predictions obtained using Equation 6.14. The best

agreement with the measurements was obtained using the parameters adopted for the

predictions on Figure 6.11, with upper and lower bounds to the predictions obtained

using; α = 0.39 and α = 0.53 (1:12 scale), α = 0.42 and α = 0.48 (1:7.2 scale), α = 0.3

and α = 0.34 (1:4.5 scale). As discussed previously in relation to Figure 6.10, this range

in α is similar to that which would be inferred from cyclic full-flow penetrometer tests,

which gave qin/qrem = 0.4 to 0.5 for this soil. It should be noted that from a design

perspective, it is this value (or range) that would be used to calculate the embedment

depth. Adopting the median α = 0.45 would result in predictions that deviate by no more

Page 213: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

179

than 10% of the measurements. For practical design the impact velocity of DEPLA

following free-fall in water may be estimated using Equation 6.8 which could create

another source of error to the embedment depth predictions obtained using Equation

6.14. However, the excellent agreement between the velocity profile predicted using

Equation 6.8 and the measured data for the 1:4.5 scale anchor (see Figure 6.9) suggests

that this error should be small for full scale DEPLA.

Page 214: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

180

(a)

(b)

(c)

Figure 6.10. Forces acting on DEPLA and velocity profiles during soil penetration

(a) 1:12 scale, (b) 1:7.2 scale, (c) 1:4.5 scale

-500 -400 -300 -200 -100 0

1.8

1.6

1.4

1.2

1.0

0.8

0.6

0.4

0.2

0.0

Fd,l

Fd,a

RfFfrict

RfFbear

Fbear

Fb

Ffrict

Fnet (theoretical)

Measured

Theoretical

Penetration resistance (N)

Dis

tan

ce t

rav

elle

d b

y a

nch

or

tip

in

so

il, s

(m)

Fnet (measured)

0 1 2 3 4 5 6 7

1.8

1.6

1.4

1.2

1.0

0.8

0.6

0.4

0.2

0.0

Velocity, v (m/s)

-2000 -1500 -1000 -500 0 500

4.0

3.5

3.0

2.5

2.0

1.5

1.0

0.5

0.0

Fd,a

RfFfrict

RfFbear

Fbear

Fb

Ffrict

Fd,l

Fnet (theoretical)

Measured

Theoretical

Penetration resistance (N)

Dis

tan

ce t

rav

elle

d b

y a

nch

or

tip

in

so

il,

s (m

)

Fnet (measured)

0 1 2 3 4 5 6 7 8

4.0

3.5

3.0

2.5

2.0

1.5

1.0

0.5

0.0

Velocity, v (m/s)

-8000 -6000 -4000 -2000 0 2000

7.0

6.5

6.0

5.5

5.0

4.5

4.0

3.5

3.0

2.5

2.0

1.5

1.0

0.5

0.0Fd,l

Fd,a

RfFfrict

RfFbear

Fbear

Fb

Ffrict

Fnet (theoretical)

Measured

Theoretical

Penetration resistance (N)

Dis

tance

tra

vel

led b

y a

nch

or

tip i

n s

oil

, s

(m)

Fnet (measured)

2 4 6 8 10 12

7.0

6.5

6.0

5.5

5.0

4.5

4.0

3.5

3.0

2.5

2.0

1.5

1.0

0.5

0.0

Velocity, v (m/s)

Page 215: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

181

Figure 6.11. Dependence of tip embedment on impact velocity

6.5 Conclusions

The installation response of DEPLAs was assessed though an extensive field testing

campaign undertaken at a lake underlain by very soft clay, using 1:12, 1:7.2 and 1:4.5

reduced scale anchors. This chapter makes a particular contribution to the body of

literature on dynamic penetration problems by presenting and analysing acceleration

data measured by an inertial measurement unit during DEPLA installation. Such data

are rarely available, but are important for assessing the mechanisms associated with

dynamic penetration in soil and for the calibration and validation of embedment models

for dynamically installed anchors. Back analysis of the acceleration data measured

during anchor installation allowed the relationship between the anchor drag coefficient

and Reynolds number (in water), and profiles of net anchor resistance and velocity

during free-fall in water and soil penetration to be established. These profiles, together

with back analysed drag coefficients, were used to assess the merit of a simple anchor

embedment model formulated in terms of strain rate enhanced shear resistance and

strain rate independent fluid mechanics drag resistance. Qualitatively, the model

0 1 2 3 4 5 6 7 8 9 10 11 12

8

7

6

5

4

3

2

1

0

1:12 scale

1:7.2 scale

1:4.5 scale

Model - best fit

Model - lower/upper bound

Impact velocity, vi (m/s)

Tota

l dis

tance

tra

vel

led b

y a

nch

or

tip i

n s

oil

, s

(m)

= 0.53

= 0.48

= 0.34

= 0.46

= 0.45

= 0.32

= 0.39

= 0.42

= 0.3

Page 216: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

182

captures the dynamic response adequately and is capable of predicting the final

embedments to within 10% of the measurements using parameters that are familiar in

offshore geotechnical design, in conjunction with two less familiar parameters related to

the increase in penetration resistance with increasing anchor velocity. These ‘dynamic’

parameters account for the increase in geotechnical resistance due to strain-rate effects

and the increase in fluid mechanics resistance due to pressure drag.

In summary, an appropriate framework for describing the dynamic installation response

of dynamically installed anchors such as the DEPLA has been presented and was shown

to be in good agreement with measured field data. This framework has been shown to

be a suitable means of predicting anchor embedment, which is a prerequisite for

estimating anchor capacity.

Page 217: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

183

CHAPTER 7 CAPACITY OF DYNAMICALLY

EMBEDDED PLATE ANCHORS AS

ASSESSED THROUGH FIELD TESTS

7.1 Abstract

A dynamically embedded plate anchor (DEPLA) is a rocket shaped anchor that

penetrates to a target depth in the seabed by the kinetic energy obtained through free-fall

and by the anchor’s self-weight. After embedment the central shaft is retrieved leaving

the anchor flukes vertically embedded in the seabed. The flukes constitute the load

bearing element as a plate anchor. This chapter presents and considers field data on the

embedment depth loss due to the plate anchor keying process and the subsequent

bearing capacity factor of the plate anchor element. The loss in plate anchor

embedment was significantly higher than that reported from corresponding centrifuge

tests and is reflected in the larger padeye displacements required to mobilise peak

capacity in the field tests. Measured plate capacities and plate rotations during keying

indicate that the end of keying coincides with the peak anchor capacity. Experimental

bearing capacity factors are in the range Nc = 14.3 to 14.6 which are consistent with

those reported from corresponding centrifuge tests but appreciably higher than existing

solutions for vanishingly thin circular plates. The higher Nc for the DEPLA is

considered to be due to a combination of the cruciform fluke arrangement and the fluke

(or plate) thickness.

7.2 Introduction

Deep water floating oil and gas facilities such as SPAR platforms, tension leg platforms

and semi-submersibles are typically moored using taut-leg or vertical mooring systems.

These systems require anchors that can sustain significant components of vertical load

while maintaining installation costs at a reasonable level (Ehlers et al. 2004). Drag-in

Page 218: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

184

anchors and direct embedment anchors resist vertical loading through a plate oriented

such that its projected area is normal (or near-normal) to the direction of loading. A

drag-in plate anchor is designed to embed when pulled along the seabed by a wire rope

or chain. When the target installation load is reached, the anchor is triggered into its

mooring position by rotating or ‘keying’ the plate such that it becomes approximately

normal to the direction of the applied load. In the case of a direct embedment plate

anchor, a follower (such as a suction caisson) is used to install the plate to the design

embedment depth, before being recovered for reuse in future installations. The plate

remains vertical in the seabed until sufficient load develops in the mooring line to cause

the plate anchor to key to an orientation that is approximately perpendicular to the

direction of loading at the padeye, allowing the full bearing resistance of the plate to be

mobilised.

Dynamically installed anchors are rocket or torpedo shaped and are designed so that,

after release from a designated height above the seafloor, they will penetrate to a target

depth in the seabed by the kinetic energy gained during free-fall. The ease and speed of

installation makes dynamically installed anchors an attractive option in deep water.

However, their axial capacity to weight ratios are low (typically less than five times the

anchor dry weight, O’Loughlin et al. 2004a, Richardson et al. 2009), and very large

anchors are required for permanent installations. The dynamically embedded plate

anchor (DEPLA) is a hybrid anchor that combines the geotechnical efficiency of plate

anchors with the low installation costs of dynamically installed anchors. The DEPLA

comprises a removable central shaft or ‘follower’ and a set of four flukes (see Figure

1.17) arranged on a cylindrical sleeve and connected to the follower with a shear pin.

The DEPLA is installed in a similar manner to other dynamically installed anchors.

After the DEPLA has come to rest in the seabed, the follower retriever line is tensioned,

which causes the shear pin to part (if not already broken during impact) allowing the

Page 219: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

185

follower to be retrieved for the next installation, whilst leaving the anchor flukes

vertically embedded in the seabed. These embedded anchor flukes constitute the load

bearing element as a plate anchor. When sufficient load develops in the mooring line the

plate rotates to an orientation that is approximately perpendicular to the direction of

loading at the padeye, allowing full bearing resistance of the plate to be mobilised. The

DEPLA installation and keying processes are summarised in Figure 1.18 and

respectively Figure 1.19.

The holding capacity of a plate anchor such as the DEPLA is controlled by the

embedment depth achieved after dynamic installation, the loss in this embedment

during keying, and the geometry dependent anchor capacity factor. Approaches for

calculating the embedment of dynamically installed anchors generally considers the

forces acting on the anchor during penetration, whilst accounting for the enhancement

of the available soil strength due to the high penetration velocities and hence strain rates

(e.g. O’Loughlin et al. 2004a, Richardson et al. 2006, Audibert et al. 2006, O’Loughlin

et al. 2013b). The merit of this approach has been demonstrated by Blake and

O’Loughlin (2015) by predicting the final embedment depth for a series of model

DEPLA field tests to within 10% of the measurements. A series of field tests were

designed to provide a basis for understanding the behaviour of the DEPLA during (i)

free-fall and dynamic embedment in soil, and (ii) quantifying the loss in embedment

during keying and the subsequent plate anchor capacity. The dynamic embedment

aspects of the study are reported by Blake and O’Loughlin (2015). This chapter

considers the keying and capacity mobilisation aspects of the field tests.

Page 220: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

186

7.3 Experimental program

7.3.1 Model anchors and instrumentation

The model DEPLAs were manufactured from mild steel and modelled at reduced scales

of 1:12, 1:7.2 and 1:4.5. This reduced scale infers a full scale anchor with a follower

length of 9 m, a combined mass (follower and flukes) of 37 tonnes, which for a typical

offshore seabed strength profile of 1.5 kPa/m would result in a plate anchor capacity of

the order of about 500 tonnes, suitable for mooring mobile offshore drilling units. The

follower was solid, except for a 185 mm long, 42 mm diameter void at the top of the

follower, which housed an inertial measurement unit that measured accelerations and

rotations during free-fall and embedment (Blake at al. 2016).

A schematic of the model DEPLA and the notation used for the main dimensions are

shown in Figure 6.1, with the main dimensions and mass of each model DEPLA

summarised in Table 6.1. 12 mm diameter Dyneema SK75 rope was used to install the

DEPLA (and to retrieve the follower) and as the mooring line connected to one of the

DEPLA flukes. This rope was selected because it was more flexible and much lighter

than wire rope, whilst maintaining high tensile strength and stiffness (<3.5% extension

at 50% of the maximum breaking load).

Follower recovery loads and plate capacities were measured above the waterline using a

tension load cell connected in series with the winch cable and the Dyneema rope. The

displacement of the follower retrieval line and mooring line was measured using a draw

wire sensor with a range of 10 m, connected in parallel with the winch line. As

described later, in a number of tests the DEPLA was jacked in the lakebed. In these tests

the orientation of the plate during keying and pullout was determined from a tri-axis

micro-electro mechanical systems (MEMS) accelerometer with a range of +/- 3g on

Page 221: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

187

each axis (ADXL 335) that was potted with Urethane compound into an 8 mm thick

polyethylene housing and bolted to one of the DEPLA flukes.

7.3.2 Test site and soil properties

Tests were carried in Lower Lough Erne, a glacial lake located in Co. Fermanagh,

Northern Ireland, in water depths varying from 3 to 12 m. The lakebed soil is a very soft

clay with high moisture contents in the range 270 to 520% (1.2 to 1.7 times the liquid

limit) and high Atterberg limits with plastic limits of 130 to 180% and liquid limits of

250 to 315% (Colreavy et al. 2012). The unit weight of the soil is constant with depth

and is marginally higher than that of water at 10.8 kN/m3. The very high moisture

content and very low unit weight is considered to be due to the very high proportion of

diatoms that are evident from scanning electron microscopic images of the soil (e.g. see

Colreavy et al. 2012). Diatoms have an enormous capacity to hold water in the

intraskeletal pore space (Tanaka and Locat 1999). However this intraskeletal pore water

is not considered to play a role in soil behaviour, and as such the measured unit weight

and other index properties that are expressed in terms of the measured moisture content

are not considered to be useful indicators of soil behaviour.

Colreavy et al. (2012) reported results from piezoball and in situ shear vane tests

conducted at the anchor test sites to depths of up to 11 m. The undrained shear strength

profiles are shown on Figure 6.2 and were derived from the net penetration resistance

data using Nball = 8.6 (calibrated using in situ shear vane data). The analysis of the

DEPLA test data used the strength over the range of interest for each anchor scale. As

will be discussed later, the 1:12 scale DEPLA achieved maximum tip embedment

depths of ze = 2.2 m. Over this depth su is best idealised as su (kPa) = 1.5z and this

profile was adopted in the analysis of the test data using this anchor. The 1:7.2 and the

1:4.5 scale DEPLAs achieved maximum tip embedment depths of ze = 4.1 m and 7.5 m,

respectively. Over this range of embedment su is best idealised as su (kPa) = 0.45 + 0.9z

Page 222: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

188

and this profile was adopted in the analysis of the test data using the 1:7.2 and the 1:4.5

scale anchors. The sensitivity of the soil is in the narrow range St = 2 to 2.5 as assessed

from in situ vane tests and piezoball cyclic remoulding tests.

7.3.3 Testing setup and procedure

Tests were carried out from a fixed jetty (1:12 scale DEPLA only) and from the deck of

a 15 m self-propelled barge with a 13 t winch and 2 t crane (Figure 5.11). The testing

procedure (summarised schematically for the barge tests in Figure 7.1) was as follows:

1. The DEPLA was released from a designated drop height above the mudline

allowing it to free-fall and penetrate into the lakebed. The anchor tip embedment, ze,

was measured by sending an underwater camera to the mudline to inspect markings

on the follower retrieval line.

2. Following dynamic installation the load cell was connected in series with the winch

cable and follower retrieval line, and the draw wire sensor was connected in parallel

with the winch cable. The winch was used to extract the follower at approximately

10 mm/s for the jetty tests and approximately 30 mm/s for the barge tests. These

were the maximum winch velocities achievable with the winches used in the

respective tests. Movement of the follower relative to the stationary flukes causes

the shear pin to break, allowing the follower to be retrieved on the barge deck (or

the jetty).

3. The plate anchor mooring line was connected to the winch cable and extracted from

the lake bed at approximately 10 mm/s for the jetty tests and approximately 30

mm/s for the barge tests. This was continued until the DEPLA plate was retrieved

from the lake.

As the aim of the field tests was to measure vertical monotonic capacity, the winch

position was maintained directly above the anchor drop site for the DEPLA keying and

Page 223: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

189

capacity mobilisation stage (Figure 7.1e). As resistance data measured during extraction

of the follower and the plate are assessed relative to the undrained shear strength, it is

important that the strain rates in the soil during extraction are comparable to those

associated with the test used to quantify the soil strength. In this instance the in situ

strength was measured using a 113 mm piezoball penetrated at the standard 20 mm/s,

which is expected to give undrained conditions. The average strain rate may therefore

be approximated by v/Dball = 0.18 s-1

, which is comparable (within one order of

magnitude) to the range associated with extraction of the follower and the plate: v/Df =

0.06 to 0.51 s-1

and v/D = 0.01 to 0.09 s-1

for the three anchor scales.

Page 224: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

190

Figure 7.1. DEPLA field test procedure

hiz

, dro

p

z wat

er

Barge

Winch

Pulley

Water

Lakebed

(a)

Follower

line

DEPLA

Quick

release

shackle

Crane

Plate line

Wat

er d

epth

Dro

p h

eig

ht

Winch

Pulley

Water

Lakebed

(b)

Follower

line

Barge

Crane

Plate line

Winch

Pulley

Water

Lakebed

(c)

Follower

line

Camera cable

Camera

Barge

Crane

Plate line

z iz,e

z e

Winch

Pulley

Water

Lakebed

(d)

Follower

line

Follower

Barge

Crane

Plate

Plate line

Winch

Pulley

Water

Lakebed

(e)

Plate line

Barge

Crane

Follower

Page 225: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

191

7.4 Results and discussion

7.4.1 Follower extraction

DEPLA tip embedment, ze, was in the ranges: 1.495 to 2.19 m (2 to 2.9Lf, 1:12 scale),

2.56 to 4.1 m (2.1 to 3.3Lf, 1:7.2 scale) and 6.197 to 7.493 m (3.1 to 3.7Lf, 1:4.5 scale).

These data, together with other relevant data presented later, are summarised in Table

7.1 for each anchor scale. A typical load-displacement response of the 1:12 scale

follower extraction is shown in Figure 7.2. The extraction is characterised by a stiff

response to an initial peak at a displacement, Δze = 0.094 to 0.154 m, followed by a

sharp reduction as the follower mobilises progressively weaker soil as it becomes

shallower. An increase in resistance is consistently observed when the follower tip

begins to enter the base of the sleeve (Δze = 0.05 m), reducing just after the tip emerges

from the top of the sleeve (Δze = 0.8 m). This increase in resistance is likely to be due to

the development of suction at the follower tip as it enters the anchor sleeve (see Figure

7.2). Considering the aspect ratio of the sleeve, the soil is expected to plug, leaving a

gap underneath the follower tip where suction is generated. The maximum follower

extraction resistance, Fvf, was in the range 419 to 590 N and typically 2.9 to 4.1 times

the dry mass of the follower. This range reflects the range in embedment depths

achieved by varying the drop height and hence the range in undrained shear strength

mobilised during follower extraction. The follower extraction resistance is slightly

higher than measured in DEPLA centrifuge tests (2.2 to 3 times the follower dry mass;

O’Loughlin et al. 2014b). This reflects the higher tip embedments achieved in the field

tests, noting that the strength profiles and nondimensional reconsolidation times, T =

cht/Df2, are broadly similar between the centrifuge and field tests. As only a few minutes

elapsed between anchor installation and follower recovery, the follower resistance is

likely to be 20 to 30% of the potential follower extraction resistance following full

reconsolidation (Richardson et al. 2009).

Page 226: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

192

Table 7.1. Summary of DEPLA field test results

1:12 scale 1:7.2 scale 1:4.5 scale

Tip embedment, ze (m) 1.495–2.2 2.56–4.1 6.197–7.493

Tip embedment ratio, ze/Lf 2–2.9 2.1–3.3 3.1–3.7

Initial plate embedment, ze,plate,0 (m) 0.890–1.585 1.554–3.094 4.587–5.883

Initial plate embedment ratio, ze,plate/D 3–5.3 3.1–6.2 5.7–7.4

Plate anchor embedment loss, Δze,plate

(m) 0.419–0.543 0.484–0.643 1.150–1.338

Plate anchor embedment loss ratio,

ze,plate/D 1.4–1.8 1–1.3 1.4–1.7

Peak plate anchor resistance, Fvp (kN) 0.52–1.46 3.64–7.78 23.76–32.37

Peak plate anchor resistance ratio,

Fvp/W 6–31 19–42 31–42

Average anchor capacity factor, Nc 14.6 14.5 14.3

Figure 7.2. Typical load-displacement response during follower extraction (1:12

scale DEPLA)

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.80

100

200

300

400

500

600

ze = 2.010 m

ze = 1.976 m

ze = 2.040 m

ze = 1.990 m

ze = 2.190 m

ze = 1.945 m

Foll

ow

er e

xtr

acti

on r

esis

tance

, F

vf (

N)

Displacement, ze

Page 227: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

193

7.4.2 Plate keying and capacity

Initial plate embedment, ze,plate,0, was in the range: 0.890 to 1.585 m (3 to 5.3D, 1:12

scale), 1.554 to 3.094 m (3.1 to 6.2D, 1:7.2 scale) and 4.587 to 5.883 m (5.7 to 7.4D,

1:4.5 scale). The load-displacement responses of the 1:12, 1:7.2 and 1:4.5 scale plate

anchors during keying and pull-out are provided on Figure 7.3, together with the

centrifuge data reported in O’Loughlin et al. (2014b) for a geometrically similar

DEPLA tested in normally consolidated kaolin clay. The capacity of the plate is

expressed in terms of the dimensionless plate anchor capacity factor, Nc, defined as:

pup

netv

csA

FN

,

, Equation 7.1

where Fv,net is the peak load minus the submerged weight of the plate anchor in soil, Ws,

Ap is the area of the plate calculated as Ap = πD2/4 (plate orientation idealised as

perpendicular to the direction of load following keying) and sup is the undrained shear

strength at the anchor embedment corresponding with peak capacity. As mentioned in

Section 7.3.3 the piezoball penetrated at a strain rate (v/Dball = 0.18 s-1

) which is

expected to give undrained conditions. This strain rate is comparable (within one order

of magnitude) to the range associated with extraction of the plate for the three anchor

scales (v/D = 0.01 to 0.09 s-1

). Hence, experimental capacity factors may be determined

without correction for strain rate effects, particularly as the penetrometer strength will

include effects of strain softening not included in mobilisation of the peak anchor

capacity.

As the anchor load is adjusted by the submerged weight of the plate in soil, Nc, is

initially slightly negative (at zero displacement), but reaches a maximum anchor

capacity factor that is directly comparable with theoretical capacity factors, as discussed

later. The occasional sharp drops in the response curves are due to momentary pauses in

Page 228: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

194

the loading which were periodically required for resetting the draw wire sensor and load

cell on the mooring line.

Figure 7.3. Typical load-displacement responses of during keying and pullout

Also included on Figure 7.3 are the corresponding responses for a DEPLA (1:12 scale)

that was jacked vertically into the lakebed. The DEPLA was jacked-in to a tip

embedment of 1.855 m (ze,plate,0 = 1.25 m, 4.2D) typical of that achieved for a dynamic

installation with an impact velocity of ~7 m/s. Results for both dynamic and jacked

installations are similar, indicating that dynamic installation does not affect either the

peak capacity or the keying response. The response is qualitatively consistent with

centrifuge test data in normally consolidated kaolin clay, including the DEPLA

centrifuge data on Figure 7.3, and data from tests on square plate anchors installed in

the vertical orientation (e.g. Gaudin et al. 2006, Blake et al. 2010). However, the loss in

embedment during keying is much higher as discussed below.

The vertical displacement of the mooring line up to the peak load, dv, was determined

from the draw wire sensor measurements and was in the range 0.544 to 0.668 m (1:12

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0-2

0

2

4

6

8

10

12

14

16

18

1:12 scale 1:12 scale (jacked) 1:7.2 scale 1:4.5 scale DEPLA centrifuge

Bea

ring

cap

acit

y f

acto

r, N

c

Mooring line displacement, dv/D

Page 229: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

195

scale), 0.697 to 0.856 m (1:7.2 scale) and 1.508 to 1.696 m (1:4.5 scale). However, the

vertical displacement dv represents the displacement of the anchor padeye rather than

the loss in plate anchor embedment, ∆ze,plate, which may be calculated as:

pullvplatee edz sin, Equation 7.2

Hence ∆ze,plate was in the range 0.419 to 0.543 m (1.4 to 1.8D, 1:12 scale), 0.484 to

0.643 m (1D to 1.3D, 1:7.2 scale) and 1.150 to 1.338 m (1.4 to 1.7D, 1:4.5 scale). This

is shown on Figure 7.4 together with centrifuge data for square and circular DEPLAs

(O’Loughlin et al. 2014b), strip anchors (O’Loughlin et al. 2006, Gaudin et al. 2009)

and square plate anchors (Song et al. 2009, Blake et al. 2010). The horizontal axis on

Figure 7.4 combines the geometrical parameters that influence the keying response:

padeye eccentricity and plate thickness (and hence anchor weight).

Figure 7.4. Loss in anchor embedment during keying

0.0 0.2 0.4 0.6 0.8 1.0 1.20.0

0.5

1.0

1.5

2.0 1:12 scale DEPLA 1:7.2 scale DEPLA 1:4.5 scale DEPLA O'Loughlin et al. (2014), circular DEPLA (cent.) O'Loughlin et al. (2014), square DEPLA (cent.) O'Loughlin et al. (2006), strip O'Loughlin et al. (2006), L/B = 2 Gaudin et al. (2009), strip Song et al. (2009), square Blake et al. (2010), square

z e,

pla

te/D

or

z e,p

late/B

[e/D t/D0.2

]1.15

or [e/B t/B0.2

]1.15

Equation 7.3

Page 230: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

196

Also shown on Figure 7.4 is the expression:

15.1

2.0

max, 144.0

B

t

B

eB

ze Equation 7.3

where B is the plate breadth for square and strip anchors. Equation 7.3 was proposed

by Wang et al. (2011) as the limiting loss in embedment for square plate anchors and is

seen to provide a reasonable limit to the previously reported experimental data, but not

the DEPLA field data. It is evident from Figure 7.4 that despite the geometrical

consistency between the field and centrifuge DEPLA tests, the loss in embedment in the

field tests (∆ze,plate = 1D to 1.8D) is much higher than that in the centrifuge tests (∆ze,plate

= 0.5D to 0.7D). This disparity was explored further by considering the following

dimensionless groups which are considered to affect the keying response (Wang et al.

2011):

000

0,,

',,

',,

uu

a

u

uplatee

s

E

s

t

s

kD

D

s

D

t

D

ef

D

z

Equation 7.4

where su0 is the local undrained shear strength at the initial embedment depth of the

anchor (i.e. as distinct from sup in Equation 7.1, which is the local undrained shear

strength at the embedment depth corresponding with peak capacity), γ'a is the

submerged unit weight of the anchor in soil and E is Young’s modulus of the soil.

Equation 7.4 uses the DEPLA plate diameter, D, rather than the plate breadth, B, as

utilised in the original form of this equation in Wang et al. (2011) in their consideration

of rectangular and strip anchors.

Table 7.2 summarises the range of values for each group in Equation 7.4 for the DEPLA

centrifuge tests and the DEPLA field tests, with the exception of E/su0, which was not

measured in either the centrifuge tests or in the field tests, and which has a limited effect

Page 231: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

197

on the loss of embedment as demonstrated by Wang et al. (2011). There is broad

agreement for groups: e/D, t/D, kD/su0, γ'at/su0, but in the field tests su0/γ'D = 4.5 to 7.5,

much higher than su0/γ'D = 0.61 to 1.44 in the centrifuge tests. This is consistent with

numerical analyses reported by Wang et al. (2011), which demonstrated that higher

strength at the anchor embedment depth resulted in higher loss of embedment, as the

moment applied to the anchor by the mooring line at the point of zero net vertical load

becomes less significant in relation to the moment capacity.

Table 7.2 Comparison between the field and centrifuge tests of the non-

dimensional groups that govern keying behaviour

Group Field Centrifuge

e/D 0.42–0.44 0.38–0.36

t/D 0.01–0.02 0.02–0.05

su0/γ'D 4.5–7.5 0.61–1.44

kD/su0 0.16–0.25 0.16–0.38

γa't/su0 0.14–0.20 0.37–0.45

Figure 7.4 indicates that the loss in DEPLA plate embedment during keying could be

reduced if either the eccentricity or anchor thickness was increased (relative to the plate

diameter or breadth). The most influential of these factors is the eccentricity

(O’Loughlin et al. 2006, Song et al. 2009, Wang et al. 2011) and the embedment loss

for the DEPLA should be lower if the padeye eccentricity was increased (albeit that this

would require a different plate geometry).

The keying and capacity response of additional jacked installation tests are reported in

Figure 7.5, where the plate was loaded at pullout angles in the range θpull = 18.6° to

91.5° (to the horizontal). The plate rotation, θp (with respect to the horizontal), was

determined from the MEMS accelerometer attached to one of the DEPLA flukes. The

vertically loaded anchor test on Figure 7.5 (91.5° load inclination) is the same jacked

Page 232: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

198

installation test shown on Figure 7.3. The capacity curve for this vertically loaded

anchor test is characterised by an initial stiff response as the anchor initiates rotation (Nc

= 4.2 at dv = 0.137 m and θp = 8.2°), followed by a softer response as the rotation angle

increases with minimal increase in anchor resistance (Nc = 5.5 at dv = 0.262 m and θp =

35.1°), and a final stiff response towards a peak capacity factor, Nc = 14.8, at the end of

keying when the plate rotation, θp = 87° (i.e. approximately equal to θpull = 91.5°).

Figure 7.5. Influence of pullout angle on bearing capacity factor (1:12 scale

DEPLA)

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4

-2

0

2

4

6

8

10

12

14

16

Bea

rin

g c

apac

ity

fac

tor,

Nc

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.40

10

20

30

40

50

60

70

80

90

100

p

Mooring line displacement, dv (m)

Pla

te r

ota

tio

n,

p

pull

= 91.5

pull

= 47.9

pull

= 32.3

pull

= 18.6

Soil

Water

pull

Page 233: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

199

These stages become progressively less distinct as the load inclination reduces. For

instance the capacity curve for θpull = 18.6° is characterised by an initial soft response,

with a large padeye displacement required to initiate capacity mobilisation (Nc = 1.1 at

dv = 0.5 m and θp = 1.5°), followed by a stiffer response towards a peak capacity factor,

Nc = 10.5. The slow build up in rotation and capacity with line displacement indicates

that the initial soft response is mainly due to movement of the line through the soil

rather than movement of the plate. Hence the magnitude of Nc for the inclined loading

cases is overestimated as the measured capacity includes the resistance of the inclined

line in the soil. It is worth noting that Figure 7.5 shows that for each load inclination the

end of keying coincides with the peak anchor capacity at a plate rotation that is tolerably

equal to the load inclination. Figure 7.5 also indicates that Nc decreases with reducing

load inclination (for ze,plate/D < 4), as also shown numerically by Song et al. (2008) and

Yu et al. (2011) for strip anchors, and recently by Wang et al. (2014) for DEPLAs.

The peak plate anchor resistance, Fvp, was in the range: 0.52 to 1.46 kN (6 to 31W, 1:12

scale), 3.64 to 7.78 kN (19 to 42W, 1:7.2 scale) and 23.76 to 32.37 kN (31 to 42W, 1:4.5

scale), where W is the dry weight of the plate. As shown by Figure 7.6, Fvp increases

linearly with depth, reflecting the linear increase in su with depth on Figure 5.9b. Also

shown on Figure 7.6 are the predicted anchor capacities using Equation 7.1, where Nc

was varied to give the best overall agreement between the measurements and

predictions for each anchor scale. This resulted in Nc = 14.6 for the 1:12 scale anchor,

Nc = 14.5 for the 1:7.2 scale anchor, and Nc = 14.3 for the 1:4.5 scale anchor. This range

is in good agreement with capacity factors back analysed from DEPLA centrifuge test

data, which are in the range Nc = 14.2 to 15.8 with an average Nc = 15.0 (O’Loughlin et

al. 2014b), and capacity factors determined numerically which gave Nc = 14.9 (Wang et

al. 2014).

Page 234: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

200

The exact solution for a deeply embedded vanishingly thin circular plate is Nc = 12.42

for a smooth interface and Nc = 13.11 for a rough interface (Martin and Randolph

2001). However, finite element results reported by Wang et al. (2010) show that Nc

increases from 13.75 for a vanishingly thin rough circular plate (4.7% higher than the

exact solution) to Nc = 14.35 for a rough circular plate with t/B = 0.05. As noted by

Zhang et al. (2012), soil is forced to take a deeper and longer route to flow around

thicker plates, which mobilises more soil in the failure mechanism and results in higher

bearing capacity factors. This, together with the cruciform fluke arrangement on the

DEPLA sleeve, will result in a larger failure surface (and hence a higher capacity factor)

compared with a flat vanishingly thin circular plate.

Figure 7.6. Measured and predicted capacity for each anchor scale

7.5 Conclusions

A series of field tests were carried out to assess the installation process and holding

capacity of dynamically embedded plate anchors (DEPLAs) at a lake underlain by very

soft clay. This chapter considered the embedment depth loss due to the plate anchor

0 5 10 15 20 25 30 35

6

5

4

3

2

1

0

Nc = 14.3

Nc = 14.5

1:12 scale 1:7.2 scale 1:4.5 scale Equation 7.1

Plate capacity, Fvp

(kN)

No

rmal

ised

pla

te e

mb

edm

ent,

ze,

pla

te/D

Nc = 14.6

Page 235: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

201

keying process and the subsequent bearing capacity factor of the plate anchor element.

Results from tests on three different reduced scale DEPLAs (1:12, 1:7.2 and 1:4.5 scale)

were presented, along with supplementary data from centrifuge tests investigating the

influence of reconsolidation following anchor installation on the keying induced loss in

DEPLA. The main findings are summarised in the following:

The load-displacement response of the DEPLA during keying and pullout is

qualitatively consistent with that observed in DEPLA centrifuge tests and other plate

anchors installed in the vertical orientation. Results for both dynamic installed and

‘jacked-in’ DEPLAs are similar, indicating that dynamic installation does not affect

either the peak capacity or the keying response.

The loss in plate anchor embedment was in the range 1D to 1.8D, which is

considerably higher than the range 0.5D to 0.7D reported from centrifuge tests in

kaolin clay on DEPLAs with the same geometry. The higher embedment loss for the

field tests is reflected in the much softer load-displacement response compared with

the centrifuge tests, and may also be due to the difference in normalised strength,

which was su/γ'D = 4.5 to 7.5 in the field tests, compared with su/γ'D = 0.61 to 1.44

in the centrifuge tests.

Rotation data measured during keying and pullout of ‘jacked-in’ DEPLAs show that

the end of keying coincides with mobilisation of peak anchor capacity, and that the

bearing capacity factor decreases moderately with reducing load inclination (for

normalised plate embedment ratios less than 4).

Average experimental bearing capacity factors were in the range Nc = 14.3 to 14.6,

which are in good agreement with Nc = 15 as back analysed from corresponding

centrifuge data, but are larger than capacity factors for circular plates due to a

combination of the cruciform fluke arrangement and the fluke thickness.

Page 236: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

202

CHAPTER 8 TOWARDS A SIMPLE DESIGN

APPROACH FOR DYNAMICALLY

EMBEDDED PLATE ANCHORS

8.1 Abstract

A DEPLA is a rocket shaped anchor that comprises a removable central shaft and a set

of four flukes. The DEPLA penetrates to a target depth in the seabed by the kinetic

energy obtained through free-fall in water. After embedment the central shaft is

retrieved leaving the anchor flukes vertically embedded in the seabed. The flukes

constitute the load bearing element as a plate anchor.

This chapter considers the results from field trials on a 1:4.5 reduced scale DEPLA at a

site off the west coast of Scotland. The data gathered in these trials are considered in

parallel with DEPLA field and centrifuge data from earlier in the experimental

campaign to assess the merit of anchor embedment and capacity models. The anchor

embedment model is cast in terms of inertial drag resistance and strain rate enhanced

shear resistance. The latter is described using a power law with separate dependence for

bearing and frictional resistance. Back analysis of acceleration data captured during

dynamic embedment indicated that the power law is capable of describing the strain rate

dependence of undrained shear strength at the very high strain rates associated with the

dynamic embedment process. Back analysed peak bearing capacity factors from the

centrifuge and field data were compared with corresponding numerically derived

values. This comparison indicated the potential for soil tension to be lost at the

underside of the anchor plate at shallow embedment, resulting in a lower capacity

factor. A simple design chart, presented in terms of the total energy of the anchor at

impact with the mudline for a range of strength gradients and soil sensitivities, is shown

to be a useful approach for predicting anchor capacity.

Page 237: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

203

8.2 Introduction

The advent of floating liquid natural gas facilities will bring an associated increase in

mooring line loads, requiring either an upscale in current anchor technology such as the

suction caisson or more technically efficient and cost-effective anchor concepts.

Dynamically installed anchors such as the torpedo pile are attractive in this regard due

to the minimal installation times, although their relatively low capacity to weight ratio

means that they must be significantly upscaled to provide the required capacity. This

has obvious cost and operational implications. Vertically loaded plate anchors are

geotechnically efficient, offering a high capacity to weight ratio, but can be difficult to

install, particularly if they are drag embedded.

A relatively new anchor concept, referred to as the dynamically embedded plate anchor

(DEPLA), combines the installation advantages of dynamically installed anchors with

the capacity advantages of vertically loaded plate anchors. The DEPLA comprises a

removable central shaft or ‘follower’ and a set of four flukes arranged on a cylindrical

sleeve and connected to the follower with a shear pin (Figure 1.17). The DEPLA is

installed in a similar manner to other dynamically installed anchors, by releasing the

anchor from a height above the seafloor, chosen such that the anchor will impact the

seabed at a velocity approaching its terminal velocity and subsequently self-bury in the

ocean floor sediments. After the DEPLA has come to rest in the seabed, the follower

retriever line is tensioned. This causes the shear pin to part (if not already broken during

impact) allowing the follower to be retrieved for the next installation, whilst leaving the

anchor flukes vertically embedded in the seabed. These embedded anchor flukes

constitute the load bearing element as a plate anchor. When sufficient load develops in

the mooring line the plate rotates to an orientation that is approximately perpendicular

to the direction of loading at the padeye, allowing full bearing resistance of the plate to

Page 238: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

204

be mobilised. The DEPLA and its installation and keying processes are summarised in

Figure 1.18 and Figure 1.19 respectively.

Predicting the holding capacity of a plate anchor such as the DEPLA is relatively

straightforward for a given shear strength profile and a known embedment depth. Plate

anchor capacity prediction is based on bearing capacity factors which are well

established e.g. Martin and Randolph (2001), Song et al. (2008b), Wang et al. (2010),

Wang and O’Loughlin (2014). However predicting this final plate anchor embedment is

challenging as it firstly requires an assessment of the final DEPLA penetration depth

after free-fall in water, and secondly requires an assessment of the extent of embedment

loss of the plate anchor keying. The loss in embedment during keying has been

investigated experimentally for suction embedded plate anchors (e.g. O’Loughlin et al.

2006, Gaudin et al. 2006, Gaudin et al. 2009, Wang et al. 2011, Cassidy et al. 2012,

Yang et al. 2012), the results of which are relevant to the DEPLA. Predicting the final

embedment of the DEPLA after free-fall is complicated as the high penetration

velocities associated with dynamic installation process requires interpretative

frameworks that are capable of scaling the soil strength from nominally undrained

values to the strain rate enhanced values associated with the very high strain rates in the

soil, and that account for hydrodynamic effects, notably the pressure drag resistance that

dominates at shallow embedment in soft soils.

The DEPLA is attractive for facilities with higher mooring loads, as relative to current

technology such as suction caissons, the anchor will be much smaller and less expensive

to install. However, the anchor concept is recent and hasn’t yet been trialled at full

scale. Towards this eventual aim of qualifying the anchor at full scale, has been an

experimental phase of proof-of-concept centrifuge tests (O’Loughlin et al. 2014b) and

reduced scale field trials (Blake et al. 2014, Blake and O’Loughlin 2015). This chapter

Page 239: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

205

considers data from the final phase of reduced scale field trials, which focused on

DEPLA installation and capacity mobilisation at a soft seabed site off the west coast of

Scotland. Measurements in the trials include accelerations during free-fall in water and

embedment in soil, and pullout resistance as the DEPLA was loaded to failure and

subsequently retrieved to the deck of the installation vessel. These data, together with

previously reported centrifuge data (O’Loughlin et al. 2014b) and field data from the

onshore site (Blake et al. 2014, Blake and O’Loughlin 2015) are considered to

investigate the merit of frameworks for predicting anchor embedment depth after

dynamic installation and subsequent anchor capacity.

8.3 Field Trials

The field tests were conducted at the Firth of Clyde which is located off the West coast

of Scotland between the mainland and the Isle of Cumbrae. At the Firth of Clyde the

water depth was typically 50 m. The test location is shown in Figure 8.1.

Figure 8.2 summarises index properties for the seabed sediments at Firth of Clyde. The

soil is very soft with moisture contents in the range 50 to 100% (close to the liquid

limit). Consistency limits plot above or on the A-line on the Casagrande plasticity chart,

indicating a clay of intermediate to high plasticity. The unit weight increases from about

γ = 14 kN/m3 at the mudline to about γ = 18 kN/m

3 at about 3.5 m (limit of the sampling

depth). Figure 8.3 shows profiles of undrained shear strength, su, with depth derived

from piezocone and piezoball tests, and calibrated using lab shear vane data and fall

cone tests, to give piezocone bearing factors Nkt = 17.8 (5 cm2 cone) and Nkt = 16.9 (10

cm2 cone), and piezoball bearing factors Nball = 11.5 (50 cm

2 ball) and Nball = 12.2 (100

cm2 ball), similar to that suggested by Low et al. (2010). The su profile is best idealised

as su = 2 + 2.8z (kPa) over the upper 5 m of the penetration profile, which is the depth

of interest for the DEPLA tests. The piezoball tests including cyclic remoulding

episodes, where the piezoball was moved vertically by ± 0.5 m (4.4 to 6.25 piezoball

Page 240: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

206

diameters) at depths of 2.3 m and 4.3 m. The degradation in soil strength is plotted for

each cycle number (following the cycle numbering notation suggested by Randolph et

al. 2007) in Figure 8.4 and indicates that the sensitivity of the soil is in the range St =

2.4 to 4.6.

Figure 8.1. Location of the field trials

Figure 8.2. Index properties at the anchor trial locations in Firth of Clyde

Firth of Clyde

0 5 10 15 20

Bulk density

Dry density

Unit weight (kN/m3)

0 20 40 60 80 100

Moisture content, w Plastic Limit, PL Liquid Limit, LL

w, PL and LL (%)

6.0

5.5

5.0

4.5

4.0

3.5

3.0

2.5

2.0

1.5

1.0

0.5

0.00 20 40 60 80 100

SandSilt

Dep

th,

z (m

)

% clay, silt, sand

Clay

Page 241: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

207

Figure 8.3. Undrained shear strength profile at the anchor trial locations in Firth

of Clyde

Figure 8.4. Degradation of undrained shear strength with cycle number during

piezoball cyclic remoulding tests

0 5 10 15 20

5.0

4.5

4.0

3.5

3.0

2.5

2.0

1.5

1.0

0.5

0.0

10 cm2 cone

5 cm2 cone

100 cm2 ball

50 cm2 ball

Lab vane

Fall cone

Undrained shear strength, su (kPa)

Pen

etra

tio

n d

epth

, z

(m)

su = 2 + 2.8z

0 2 4 6 8 10 120.0

0.2

0.4

0.6

0.8

1.0

1.2

Test 1 (2 to 3 m)

Test 1 (4 to 5 m)

Test 2 (2 to 3 m)

Test 2 (4 to 5 m)

Test 3 (2 to 3 m)

Test 3 (4 to 5 m)

Cycle number, n

Deg

rad

atio

n f

acto

r, q

(n)/

qin

Page 242: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

208

8.4 Instrumented reduced scale DEPLA

The DEPLA was modelled at a reduced scale of 1:4.5, based on a full scale anchor with

a follower length of 9 m (appropriate for typical mobile drilling operations). It was

manufactured from mild steel with a total mass, m = 388.6 kg, of which the follower

mass, mf = 297 kg and the plate mass, mp = 91.6 kg. A schematic of the model DEPLA

and the notation used for the main dimensions are shown in Figure 6.1, with the main

dimensions of the anchor summarised in Table 6.1. The onshore trials reported by Blake

at al. (2014) and Blake and O’Loughlin (2015) tested this same 1:4.5 scale anchor and

two smaller anchors at scales of 1:7.2 and 1:12. As these data are also considered

throughout the chapter their dimensions are masses are also listed on Table 6.1. 12 mm

diameter Dyneema SK75 rope was used to install the DEPLA (and to retrieve the

follower) and as the mooring line connected to one of the DEPLA flukes.

The DEPLA follower was solid, except for a 185 mm long, 42 mm diameter void at the

top of the follower, which housed a custom-design, low-cost, six degree of freedom

Inertial Measurement Unit (IMU). The IMU includes a 16 bit three component MEMS

rate gyroscope (ITG 3200) and a 13 bit three-axis MEMS accelerometer (ADXL 345).

The gyroscope had a resolution of 0.07 °/s with a measurement range of +/- 2000 °/s.

The accelerometer had a resolution of 0.04 m/s2 with a measurement range of +/-16 g.

Data were logged by an mbed micro controller with an ARM processor to a 2 GB SD

card at 400 Hz. Internal batteries were capable of powering the logger for up to 4 hours.

The IMU was contained in a watertight aluminium tube 185 mm long and 42 mm in

diameter.

The IMU measurements were relative to the body-axis of the DEPLA which was free to

pitch and/or roll during free-fall in the water column and embedment in the soil. Hence,

it was necessary to convert these measurements to a fixed inertial frame of reference

using transformation matrices as described in detail by Blake et al. (2016).

Page 243: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

209

8.5 Test procedure

The field trials were conducted from the RV Aora, a 22 m research and survey vessel in

Firth of Clyde (see Figure 5.10) equipped with several winches and an 8 tonne crane.

The testing procedure involved the following stages:

1. The IMU was powered up and secured in the follower.

2. The DEPLA was assembled on the deck of the RV Aora by lowering the follower

through the flukes and connecting the follower and flukes with a shear pin (see

Figure 8.5)

3. The anchor was then lowered in the water to the required drop height and then

released by opening a quick release shackle connecting the follower retrieval line to

the crane, allowing the anchor to free-fall and penetrate the seabed.

4. The anchor tip embedment depth, ze, was measured by sending a remotely operated

vehicle to the seabed to inspect markings on the follower retrieval line (see Figure

8.6).

5. Following dynamic installation the load cell was connected in series with the winch

cable and follower retrieval line, and the winch was used to extract the follower at

approximately 30 mm/s. Movement of the follower relative to the stationary flukes

causes the shear pin to break, allowing the follower to be retrieved to the vessel

deck.

6. The plate anchor mooring line was then connected to the winch cable and a draw

wire sensor was connected in parallel with the winch cable to allow for monitoring

of the line displacement. The plate was loaded using the winch, which caused the

winch line to move at 30 mm/s (as for the follower retrieval) and effected the keying

and capacity mobilisation of the DEPLA plate. This was continued until the DEPLA

plate was retrieved from the seabed and recovered on the vessel deck.

Page 244: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

210

Figure 8.5. Assembling the DEPLA on the deck of the RV Aora in preparation for

installation

Figure 8.6. An ROV video still showing the DEPLA installation site. These

observations were primarily intended to ensure no gross error in the embedment

depth established from the anchor instrumentation, but also provided evidence

that the hole created by the passage of the anchor remained open.

8.6 Results and discussion

The experimental programme at Firth of Clyde included thirteen tests as summarised in

Table 8.1. These results are discussed in further detail in the following sections.

Page 245: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

211

Table 8.1. Summary of field test results at the Firth of Clyde

Test

no.

Anchor

release

height,

(m)

Anchor

impact

velocity,

vi (m/s)

Anchor tip

embedment,

ze (m)

Anchor

travel

distance,

sz (m)

Peak

anchor

capacity,

Fvp (kN)

Initial plate

embedment

depth,

ze,plate,0 (m)

Final plate

embedment

depth,

ze,p,final (m)

Plate

embedment

ratio,

ze,p,final/D

Anchor

capacity

factor,

Nc

1 42.508 12.9 4.010 4.19 31.2 2.4 1.589 2 9.4

2 40.552 12.4 3.685 3.8 32 2.075 1.264 1.6 11.2

3 48.426 11.5 3.292 3.46 29.5 1.682 0.871 1.1 12.9

4† 47.223 11.4 3.504 3.76 - - - - -

5# 46.464 12.3 3.874 4.22 - - - - -

6 49.3# - 3.7* - 21.9 2.090 1.279 1.6 7.6

7 47.25 11.8 3.326 3.62 25.6 1.716 0.905 1.1 10.9

8 48.797 12.1 3.386 4.15 29.9 1.776 0.965 1.2 12.3

9 47.824 11.7 3.611 3.86 23.8 2.001 1.19 1.5 8.6

10** 48.472 10.8 3.3* - 20.2 1.69 0.879 1.1 9.7

11 50# - 3.7* - 29.5 2.090 1.279 1.6 10.3

12 2.259 5.6 2.754 2.77 6.9 1.144 0.333 0.4 4.2

13 17.678 10.6 3.305 3.53 21.9 1.695 0.884 1.1 9.4

# Determined from RV Aora’s echo sounder measurements.

* Determined from ROV video capture. † ROV tangled in mooring follower, follower and plate recovered together.

# Load cell data acquisition system failed.

** IMU failed after impact with the lakebed.

8.6.1 Impact velocity and embedment

Anchor impact velocities achieved in the field tests were in the range 5.6 to 12.9 m/s,

corresponding to anchor release heights of 2.3 to 42.5 m. Corresponding anchor tip

embedments were in the range 2.75 to 4.01 m (up to 2 times the anchor length). These

results are shown on Figure 8.7 and compared with results from the onshore trials

(Blake and O’Loughlin 2015) conducted in a lake underlain by clay with a strength

gradient of about 1 kPa/m (i.e. almost 3 times lower than that at Firth of Clyde). Figure

8.7 shows a clear dependence of embedment depth on impact velocity, with similar

dependence for both datasets. Although the stronger soil at the Firth of Clyde site

reduces the embedment depth by about 50%, as will be shown later in the chapter the

anchor capacity is not significantly affected. As shown by O’Loughlin et al. (2013b),

comparisons of embedment trends between anchors of different geometries and mass,

installed in seabeds with different strength profiles can be simplified by considering the

Page 246: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

212

total energy of the anchor, defined as the sum of the kinetic and potential energy

(relative to the final embedment depth) of the anchor as it impacts the soil:

eitotal gzmmvE '2

1 2 Equation 8.1

where m' is the effective mass of the anchor (submerged in soil). The data considered

here are plotted in non-dimensional form on Figure 8.8, with tip embedments expressed

as ze/deff, (where the effective anchor diameter, deff accounts for the additional projected

area due to the anchor flukes) and total energy is expressed as Etotal/kdeff4 (removing the

influence of soil strength gradient, k and anchor diameter) together with previously

reported DEPLA field and centrifuge data (O’Loughlin et al. 2014b, Blake and

O’Loughlin 2015) and other available centrifuge and field data for dynamically installed

anchors. The data form a relatively tight band that can be conservatively fitted by the

following expression suggested by O’Loughlin et al. (2013b):

31

4

eff

total

eff

e

kd

E

d

z Equation 8.2

Where deff is calculated from:

f

eff

Ad

4 Equation 8.3

As shown by O’Loughlin et al. (2014b), Equation 8.2 has merit in obtaining a first order

approximation of the anchor scale required for a given mooring design, However, the

normalisation in Figure 8.8 has further benefits as demonstrated later in the chapter.

Page 247: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

213

Figure 8.7. Embedment trends for different scale DEPLAs in soils with different

strength gradients

Figure 8.8. Harmonising embedment trends through the use of total energy

0 2 4 6 8 10 12 14

4

3

2

1

0

k = 2.8 kPa/m (current study)

1:4.5 scale

k ~ 1 kPa/m (Blake and O'Loughlin 2015)

1:4.5 scale

1:7.2 scale

Impact velocity, vi (m/s)

Norm

alis

ed t

ip e

mbed

men

t, z

e/Lf

0 5 10 15 20 25 30 35 40

40

35

30

25

20

15

10

5

0

k = 2.8 kPa/m (field, current study)

1:4.5 scale DEPLA

k ~ 1 kPa/m (field, Blake and O'Loughlin 2015)

1:4.5 scale DEPLA

1:7.2 scale DEPLA

k ~ 1.5 kPa/m (centrifuge, O'Loughlin et al. 2014b)

1: 200 scale DEPLA

Other dynamically installed anchor/projectile data, k = 1 to 1.8 kPa/m

centrifuge (O'Loughlin et al. 2013b)

field (Lieng et al. 2010)

field (Brandão et al. 2006)

Etotal

/kdeff

4

z e/d

eff

× 103

3/1

4

eff

total

eff

e

kd

E

d

z

Page 248: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

214

8.6.2 Plate anchor capacity

After removing the DEPLA follower, the plate anchor mooring line was tensioned such

that the DEPLA plate was vertically loaded. This loading causes the DEPLA to rotate or

‘key’ in the soil, undergoing a loss of embedment in the process. This is shown clearly

by Figure 8.9, which plots the mooring line displacement, dv, as a function of the anchor

load, Fvp. Also shown on Figure 8.9 is an example response reported by Blake et al.

(2014) from the onshore site with k = 1 kPa/m. It is worth noting the similar capacity

for both sites despite the much shallower embedment from the Firth of Clyde tests

associated with the much higher strength gradient. This is because the capacity of

dynamically installed anchors such as the DEPLA is governed by the soil strength at the

final embedment depth. Embemdent will be higher in weaker soils and lower in stronger

soils but with approximately the same strength at the final plate embedment depth in

either case.

Although in principle the loss in embedment may be established from the mooring line

displacement (that was measured on the vessel deck), the elongation of the high-

modulus polyethylene (HMPE) rope makes attempts to derive the plate displacement

from the mooring line displacement erroneous and certainly unreliable. This is more

pronounced on Figure 8.9 for the results from Firth of Clyde than for the lake tests

reported by Blake et al. (2014) as the water depth, and hence elongation of the rope, is

much higher at Firth of Clyde. However, the loss in embedment during keying has been

extensively investigated for plate anchors (e.g. Song et al. 2006, O’Loughlin et al. 2006,

Gaudin et al. 2006) and as shown by O’Loughlin et al. (2014b) the loss of embedment

for a DEPLA is similar to that of a geometrically comparable plate and can be

calculated from:

Page 249: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

215

15.1

2.0

max, 144.0

D

t

D

eD

ze Equation 8.4

Hence for the 1:4.5 scale DEPLA considered here, the loss in embedment is calculated

to be Δze,max ~1D (or 0.8 m).

The DEPLA capacity may be expressed as an anchor capacity factor:

pup

netv

csA

FN

,

, Equation 8.5

where Fv,net is the peak load minus the submerged weight of the plate anchor in soil, Ap

is the area of the plate and su,p is the undrained shear strength at the estimated anchor

embedment at peak capacity, accounting for the loss of embedment calculated using

Equation 8.4. For determination of Nc using Equation 8.5 to be meaningful it is

important that the strain rates associated with mobilisation of anchor capacity are

comparable with those associated with measurement of su. As described earlier, su was

determined using 5 cm2 and 10 cm

2 piezocones and 50 cm

2 and 100 cm

2 piezoballs,

penetrated at the standard 20 mm/s. The average strain rate in the soil surrounding the

penetrometers may therefore be approximated by v/Dball = 0.25 to 0.18 s-1

and v/Dcone =

0.79 to 0.56 s-1

. This overall range is comparable (within one order of magnitude) to

v/D = 0.04 s-1

associated with loading of the DEPLA plate. Hence, in this instance

experimental capacity factors may be determined without correction for strain rate

effects, particularly as the penetrometer strength will include effects of strain softening

not included in mobilisation of the peak anchor capacity.

Page 250: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

216

Figure 8.9. Typical load-displacement responses during keying and pullout

Nc values for the Firth of Clyde trials are compared with other DEPLA centrifuge and

field data on Figure 8.10. Evidently the plate anchor embedment at peak capacity for the

current data are much shallower than for the other field and centrifuge data, due to the

much higher strength gradient at Firth of Clyde. Wang and O’Loughlin (2014) report

DEPLA capacity factors derived from large deformation finite element analyses for the

case where soil tension is permitted at the base of the DEPLA plate (the so-called ‘no

breakaway’ or ‘attached’ case) and for the case where soil tension is not permitted at the

base of the plate (the so-called ‘immediate breakaway’ or ‘vented’ case). The results

from these analyses are shown on Figure 8.10 for a fully rough interface and t/D =

0.027 (rather than t/D = 0.0125 as used in the field tests). Experimentally derived

DEPLA capacity factors at deep embedment (i.e. at z/D ≥ 2.5) have a mean Nc = 14.6,

and as shown by Figure 8.10, are in good agreement with the (deep) no-breakaway

capacity factor, Nc = 14.9. This limiting capacity factor is 13.7% higher than Nc =

13.11, which is the exact solution for a deeply embedded infinitely thin rough circular

plate (Martin and Randolph 2001), but only 3.5% higher than the numerically derived

Nc = 14.4 for a deeply embedded rough circular plate with the same t/D = 0.027 (Wang

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.00

5

10

15

20

25

30

35

Test 7

Test 8

Blake et al. (2014)

Pla

te c

apac

ity

, F

vp (

kN

)

Mooring line displacement, dv/D

Page 251: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

217

and O’Loughlin 2014). This suggests that the failure mechanism for a circular plate and

a circular DEPLA are broadly similar for the same plate thickness. For shallow

embedment, the experimentally derived DEPLA capacity factors are in the range Nc =

4.2 to 12.9. These values are much lower than Nc at deep embedment, which indicates

that breakaway may have occurred.

Figure 8.10. Experimental and numerical anchor capacity factors

Several studies including Das and Singh (1994), Merifield et al. (2001, 2003), Wang et

al. (2010) and Wang and O’Loughlin (2014) have shown that the immediate breakaway

capacity factor may be expressed as:

c

u

platee

ccb Ns

zNN

,

0

' Equation 8.6

where Nc0 is the immediate breakaway capacity factor in weightless soil (also shown on

Figure 8.10) and the dimensionless term, γ'z/su, captures the effect of overburden

pressure. Figure 8.10 shows that at deep embedment the experimentally derived

0 1 2 3 4 5 60

2

4

6

8

10

12

14

16

18

Breakaway,

weightless soil (Nc0

)

Current study

Field data, Blake et al. (2014)

Centrifuge data, O'Loughlin et al. (2014b)

Numerical solutions (Wang and O'Loughlin 2014)

ze,p,final

/D

Cap

acit

y f

acto

r, N

c, N

c0 a

nd N

cb

No breakaway

(Nc)

Breakaway (Ncb

), 'ze,p,fimal

/su = 1.5

(Equation 5)

Page 252: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

218

capacity factors are in quite good agreement with the numerical no-breakaway values,

whereas at shallower embedment (i.e. relating to the current dataset), the experimental

capacity factors lie within the bounds of the (soil weight adjusted) immediate

breakaway case (Equation 8.6) and the no-breakaway case. This observation appears

completely reasonable as at shallow embedment the failure mechanism extends to the

soil surface, such that the gap formed during breakaway may be filled with the free

water above the mudline. In contrast, at deep embedment where the mechanism is

localised, the formation of a gap beneath the DEPLA plate appears implausible as (for

this rate of loading) water would not be able to fill the void.

8.6.3 Embedment depth prediction model

For typical offshore seabed deposits that are generally characterised by increasing

strength with depth, reliability in the calculated anchor capacity firstly requires accurate

prediction of the final anchor embedment depth after dynamic installation. In practice

this can be confirmed by retrieving instrumentation of the type used in these tests from a

‘pod’ located on the exterior of the anchor (Lieng et al. 2010), but from a design

perspective the final embedment depth needs to be precalculated so that the anchor can

be scaled for the design mooring line load. As shown by Figure 8.8, a first order

estimate of the embedment depth for a given anchor geometry, mass and seabed

strength profile may be obtained using Equation 8.2. However, this approach does not

capture subtleties of the dynamic embedment process, including the influence of the

variation in soil strength with strain (i.e. sensitivity) or strain rate, or geometrical

aspects of the anchor that may influence the hydrodynamic response. A more rigorous

approach would be to consider the forces acting on the anchor during dynamic

embedment in soil and to calculate anchor velocities and displacements from the net

acceleration associated with the resultant force. Several studies (e.g. True 1976,

Levacher 1985, Beard 1981, Aubeny and Shi 2006, Audibert et al. 2006, O’Loughlin et

Page 253: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

219

al. 2004b, 2009, 2013b) have adopted such an approach, with variations on the inclusion

and formulation of the various forces acting on the anchor (or projectile) during

penetration. Figure 8.11 shows the forces acting on the DEPLA during embedment in

soil. Consideration of these forces leads to a governing equation of:

ldadbearfrictfbs FFFFRFW

dt

sdm .,2

2

)( Equation 8.7

where m is the anchor mass, z is depth , t is time, Ws is the submerged anchor weight (in

water), Fb is the buoyant weight of the displaced soil, Rf is a shear strain rate function,

Fbear is the bearing resistance, Ffrict is the frictional resistance and Fd is the inertial drag

resistance. The added mass, m', in Equation 8.7 is the ‘added’ mass of the soil that is

accelerated with the anchor. However for slender bodies such as the DEPLA, m' is

negligible and can be taken as zero (Beard 1981, Shelton et al. 2011). Frictional and

bearing resistances are formulated as

AsF ufrict Equation 8.8

AsNF ucbear Equation 8.9

where α is a friction ratio (of limiting shear stress to undrained shear strength), Nc is the

bearing capacity factor for the projectile tip or fluke, and su is the undrained shear

strength averaged over the contact area, A.

The inclusion of Fd in Equation 8.7 may appear controversial as this term is typically

associated with a fluid mechanics framework for Newtonian materials. However, it is

also justified for the (non-Newtonian) very soft fluidised soil typically encountered at

the surface of the seabed, and has been shown to be important for assessing loading

from a submarine slide runout on a pipeline (Boukpeti et al. 2012, Randolph and White

2012, Sahdi et al. 2014). O’Loughlin et al. (2013b) and Blake and O’Loughlin (2015)

Page 254: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

220

further showed that fluid drag is the dominant resistance acting on a dynamically

installed anchor in normally consolidated clay during initial embedment and typically to

about 30% of the penetration. Inertial drag resistance includes pressure drag created by

the pressure gradient between the rear and the front of the projectile and friction drag

generated by viscous stresses that develop in the boundary layer between the anchor and

the soil. For a free-falling anchor at the high non-Newtonian Reynolds numbers

associated with dynamic installation, the pressure drag component is due to the frontal

area of the anchor, whereas the friction drag component is mainly due to the contact

area of the trailing mooring and follower recovery lines. The fluid drag resistance terms

are formulated as (Fernandes et al. 2006):

swldfsadldadd ACACv

FFF ,,

2

,,2

Equation 8.10

where Cd,a is the drag coefficient of the anchor, Cd,l is the drag coefficient of the line, ρs

is the density of the soil, Af is the frontal area of the anchor, As is the surface area of the

line in contact with water and v is the anchor (and hence line) velocity. It is worth

noting that Equation 8.10 uses the density of water rather than soil for the drag on the

lines as for most applications most of the contact with the line will be in water,

particularly if the hole caused by the passage of the anchor remains open. ROV video

capture of the drop sites (see Figure 8.6) showed an open crater. Sabetamal (2014) used

the Arbitrary Lagrangian Eulerian method in finite element analyses to show that the

hole formed by the passage of the anchor would close immediately at a strength ratio,

su/γ'D = 3.6, but would stay open at a strength ratio, su/γ'D = 10.6. This is consistent

with the field data, which have an average strength ratio at the top of the follower,

su/γ'D = 7.3, and also with experimental work reported by Morton et al. (2014) that

shows that at this strength ratio, hole closure would be expected at a tip embedment of

3.85 m, which is the limit of the embedments achieved in the field tests (see Table 8.1).

Page 255: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

221

Figure 8.11. Forces acting on the DEPLA during dynamic embedment in soil

Water

Lakebed

Soil buoyancy

(Fb)

Fluke reverse

end bearing

(Fbear)

Fluke friction

(Ffrict)

Sleeve friction

(Ffrict)

Sleeve end bearing

(Fbear)

Fluke end bearing

(Fbear)

Follower Friction

(Ffrict)

Submerged weight

(Ws)

Follower tip end

bearing (Fbear)

Anchor fluid drag

(Fd,a

)

Follower recovery

line fluid drag

(Fd,l

)

Mooring line fluid

drag (Fd,l

)

Page 256: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

222

Drag coefficients for the anchor and the lines can be established from acceleration data

measured during free-fall in water using a reduced form of Equation 8.7 that does not

include the geotechnical resistance terms:

swldfsadsds ACACv

WFWdt

zdm ,,

2

2

2

2 Equation 8.11

where d2z/dt

2 is the measured acceleration. Equation 8.11 includes two unknowns, Cd,a

and Cd,l. Cd,a is best selected by optimising the fit between the theoretical and

experimental velocity profiles at depths less than the terminal velocity where pressure

drag dominates the resistance, whereas Cd,l may be obtained by considering the fit at

depths greater than the terminal velocity when additional drag resistance is due solely to

the increase in drag resistance on the lines. Example comparisons are provided in Figure

8.12, where the theoretical velocity profile was established from a finite difference

approximation of Equation 8.11 and the experimental velocity profile was established

from double numerical integration of the measured acceleration data (see Blake et al.

2016 for further details). In this instance Cd,a = 0.7 and Cd,l = 0.004 and 0.005 provide

the best fit to each dataset. The latter are lower than Cd,l = 0.015 reported by Blake and

O’Loughlin 2015) for the onshore site, and is considered to be due to the different

anchor deployment and line arrangement at each site.

The frictional resistance term in Equation 8.8 comprises friction on the follower, sleeve

and flukes, whereas the bearing resistance term in Equation 8.9 comprises bearing on

the follower tip and at the base and upper end of the flukes (but not at the upper end of

the follower due to the assumption that the cavity created by the passage of the

advancing anchor remains open).

Page 257: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

223

Figure 8.12. Determining drag coefficients Cd,al and Cd,l by optimising the fit

between measured and theoretical velocity profiles in water

Soil strength depends on the rate at which it is sheared. This has been quantified in

laboratory element tests reported by Vade and Campenella (1977), Graham et al.

(1983), Lefebvre and Leboeuf (1987) and others, where a 10 to 20% increase in

undrained shear strength was typically observed for every log cycle increase in shear

strain rate. Similar dependence has been observed in shear vane tests (e.g. Biscontin and

Pestana 2001, Peuchen and Mayne 2007) and in variable rate penetrometer tests (Chung

et al. 2006, Low et al. 2008, Lehane et al. 2009). This dependence of shear strength on

shear strain rate is generally formulated using either semi-logarithmic or power

functions (e.g. Biscontin and Pestana 2001), with the power function being more suited

to the higher orders of magnitude variation in shear strain rate associated with dynamic

installation problems. The power law function is expressed as:

ref

fR

Equation 8.12

where β is a strain rate parameter, γ is the strain rate and refγ is the reference strain rate

associated with the reference value of undrained shear strength. A similar formulation,

0 2 4 6 8 10 12 14

30

25

20

15

10

5

0

Measured

Equation 8.11

Velocity, v (m/s)

Norm

alis

ed f

ree-

fall

dis

tance

, z'

/Lf

Cd,l

= 0.005

Cd,l

= 0.004

Page 258: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

224

originating from fluid mechanics, is the so-called Herschel-Bulkley law. This law

describes the same form of strain rate dependence as the power law, but avoids the need

to limit Rf at strain rates lower than the reference strain rate by also including a limiting

yield stress, making it more suited to implementation in numerical models. During

dynamic penetration, the shear strain rate will vary through the soil body, but for

shearing around an axially loaded cylindrical object such as the DEPLA, it is reasonable

to assume that at any given location the operational strain rate will be proportional to

the normalised velocity, v/d, such that Equation 8.12 may be replaced by:

refref

fdv

dvnor

dvnR

Equation 8.13

If the reference strength is determined from laboratory element tests, refγ in Equation

8.13 is the strain rate employed in the element test, whereas if the reference strength is

determined from a penetrometer test (such as the cone, T-bar or ball), refγ may be

approximated by (v/d)ref, where v is the penetration velocity and d is the diameter of the

penetrometer. An important point to note is that in situ strengths determined from

penetrometer tests are typically similar in magnitude to those determined in the

laboratory (Low et al. 2010), despite a typical 5 orders of magnitude difference in the

strain rates associated with the laboratory test (e.g. 1%/hr, 3 × 10-6

s-1

) compared with

approximately 0.18 s-1

associated with for example a piezoball test (Randolph et al.

2007). This is due to the compensating effects of strain softening associated with the

penetrometer strength and lower strain rate dependency at strain rates closer to the

laboratory strain rate (Hossain and Randolph, 2009). Hence, caution is required if

Equation 8.13 is adopted using a laboratory reference strain rate and strength, as the

(intact) laboratory strength needs to be adjusted to reflect the strain softening associated

Page 259: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

225

with a penetrometer or anchor installation and the strain rate dependency needs to be

more moderate over the first few decades increase in strain rate.

The parameter n in Equation 8.13 is introduced to account for the greater rate effects

reported for shaft resistance compared to tip resistance (Dayal et al. 1975, Steiner et al.

2014, Chow et al. 2014). It is quite probable that the higher rate effect for shaft

resistance is due to the higher strain rate at the cylindrical shaft involving curved shear

bands, which can be estimated to be about 20 to 40 times v/d (for β = 0.10 and 0.05

respectively) using a rigorous energy approach maintaining equilibrium (Einav and

Randolph 2006). Therefore, n is taken as 1 for bearing resistance (Zhu and Randolph

2011) and as a function of β (adopted from Einav and Randolph 2006) for estimating

rate effects in frictional resistance according to:

22 1

1 nn

n

Equation 8.14

where nl is taken as 1 for axial loading.

The acceleration data measured during penetration allows for an assessment of the

suitability of the power law at the high strain rate associated with dynamic embedment.

Rf values for bearing and friction may be back-analysed from the acceleration data by

rearranging Equation 8.7, noting that Rf,frict = nβRf,bear to give:

bearfrict

dbs

bearfFFn

FFdt

zdmW

R

2

2

, Equation 8.15

Rf values for bearing and friction were obtained from the acceration data using Cd,a =

0.7 as determined from the free-fall in water phase of the test (see Figure 8.12), but

taking Cd,l = 0 for penetration is soil, as the lines were assumed to become slack on

impact with the relatively strong seabed at Firth of Clyde. Bearing resistance was

Page 260: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

226

calculated using Equation 8.9 using Nc = 12 for the follower tip (O’Loughlin et al.,

2013b) (but not for the padeye as the hole formed by the passage of the anchor was

assumed to remain open during anchor penetration) and Nc = 7.5 for the leading and

trailing edges of the flukes (analogous to a deeply embedded strip footing). The strain

rate parameter was fixed at β = 0.08, which is typical of that measured in variable rate

penetrometer testing (Low et al. 2008, Lehane et al. 2009) and approximates to an 18%

change per log cycle change in strain rate, typical of that measured in laboratory testing

as discussed earlier. The back-analysed values of Rf for friction and bearing (denoted as

‘measured’ in Figure 8.13) are compared for two example tests in Figure 8.13 with Rf

formulated using the power law (Equation 8.13). The best agreement between the

backfigured Rf and that predicted by the power law was obtained using α = 0.22 (Test

12) and 0.29 (Test 2). This is within the range established from cyclic piezoball

remoulding tests, which gave qin/qrem = 0.2 to 0.32. Encouragingly, the power law is

seen to satisfactorily quantify the correct amount and variation in strain rate dependence

over the range of strain rates associated with dynamic penetration. However, Figure

8.13 also plots the same Rf data against depth, revealing that there are other mechanisms

during initial penetration that are not captured by Equation 8.7. Nonetheless, Figure

8.14 shows that the embedment model is seen to predict the experimental velocity

profiles with bounds on α of 0.22 and 0.29, similar to the bounds of qin/qrem from the

cyclic piezoball remoulding tests. More importantly the final embedment depths are

predicted accurately, which in turn increases the reliability of anchor capacity

calculations. This is made clear by Figure 8.15, which shows embedment test data for

all the Firth of Clyde tests together with the corresponding embedment model

predictions using an average α = 0.26, presented in the ‘total energy’ format of Figure

8.14. Also shown on Figure 8.15, are predictions formulated using Equation 8.2, which

are in good agreement with the experimental data and the embedment model

Page 261: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

227

predictions. This presentation format is appealing as predicted anchor embedment can

be presented on a single chart for a variety of model input parameters, making the

model easily adoptable in design as shown below.

Figure 8.13. Evaluation of the power law in quantifying strain rate dependency for

dynamic installation problems: (a) and (c) Test 12, (b) and (d) Test 2

Figure 8.14. Measured and predicted velocity profiles

0 50 100 150 200 2500.0

0.5

1.0

1.5

2.0

2.5

3.0

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0

3.0

2.5

2.0

1.5

1.0

0.5

0.00.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0

4.0

3.5

3.0

2.5

2.0

1.5

1.0

0.5

0.0

0 50 100 150 200 250 300 350 400 4500.0

0.5

1.0

1.5

2.0

2.5

3.0

Friction

Measured Power law

v/d / (v/d)ref

Str

ain

rat

e fa

cto

r, R

f

Bearing

(c)

0 1 2 3 4 5 6 7

Velocity, v (m/s)

Dis

tance

, z

(m)

Friction

Measured Power law

Strain rate factor, Rf

Bearing

(d)

Measured Power law

Strain rate factor, Rf

Dis

tance

, z

(m)

FrictionBearing

(b)

Str

ain

rat

e fa

cto

r, R

f

Measured Power law

v/d / (v/d)ref

Friction

Bearing

(a)0 1 2 3 4 5 6 7

Velocity, v (m/s)

0 1 2 3 4 5 6 7 8 9 10 11 12 13

4.0

3.5

3.0

2.5

2.0

1.5

1.0

0.5

0.0

Measured

Embedment model

Resultant velocity, v (m/s)

An

cho

r tr

avel

dis

tan

ce,

s (m

)

= 0.22 (Test 12)

= 0.27 (Test 13)

= 0.26 (Test 4)

= 0.24 (Test 9)

= 0.29 (Test 2)

Page 262: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

228

Figure 8.15. Measured and predicted DEPLA follower tip embedments

8.7 Design approach

Figure 8.16 summarises embedment predictions made using Equation 8.7 for su = 0 (at

the mudline), undrained strength gradients, k = 1, 2 and 3 kPa/m and interface friction

ratios, α = 0.1, 0.2, 0.3, 0.4 and 0.5 (α may be determined from the inverse of the soil

sensitivity as assessed from piezoball cyclic remoulded tests i.e. α = 1/St). In this

example β = 0.08, which is typical of that derived from variable rate penetrometer

testing in normally consolidated clay (Low et al. 2008, Lehane et al. 2009), although

additional charts may be generated for different strain rate parameters and indeed

different strength gradients. Although total energy is normalised by the strength

gradient, the predicted embedments still demonstrate some dependence on k, which is to

be expected as the normalisation does not account for inertial drag which will vary

relative to the shear resistance as the strength gradient changes. Figure 8.16 can be used

to scale a DEPLA for a given design load as described in the following steps and

summarised schematically in Figure 8.17:

1. Select a DEPLA plate diameter, D, thickness, t, and padeye eccentricity, e.

2. Calculate the embedment depth loss of the plate, Δze,plate using Equation 8.4.

0 1 2 3 4 5 6 7 8 9 10

25

20

15

10

5

0

Measurements

Embedment model = 0.26)

Equation 8.2

Etotal

/kdeff

4

z e/d

eff

× 103

Page 263: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

229

3. Select a capacity factor from Figure 8.10: Nc = 14.9 assuming deep embedment

(ze,p,final ≥ 2.5D).

4. Calculate the required follower tip embedment, ze, required for the chosen D value,

using a modified version of Equation 8.5:

sfplatee

pc

vp

e HLzkAN

Fz , Equation 8.16

5. Select the other anchor dimensions (Lf, Df, Ltip, t, ts, Hs etc.).

6. Calculate ze,p,final using the expression:

plateesfefinalpe zHLzz ,,, 5.0 Equation 8.17

7. If ze,p,final ≥ 2.5D, proceed to step 7, if ze,p,final < 2.5D repeat steps 1 to 6 choosing an

alternative D.

8. Calculate the anchor mass, m, and effective mass, m'.

9. Calculate the anchors impact velocity, vi, based on 80% of terminal velocity, vt (this

is a simplified and conservative approach to account for the drag of the follower

recovery line and mooring lines):

adfw

tiCA

mgvv

,

28.08.0

Equation 8.18

10. Calculate the total energy, Et, using Equation 8.1.

11. Calculate the effective diameter of the anchor, deff: using Equation 8.3.

12. Repeat steps 1 to 11 until the normalised tip embedment, ze/deff, sit on the prediction

line for the required k and α in Figure 8.16.

Page 264: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

230

Figure 8.16. DEPLA design chart for a range of seabed strength gradients and

interface friction ratios

Two example sites are now considered, making use of the design procedure outlined

above, but limiting the anchor impact velocity to 80% of its terminal velocity to keep

the anchor drop height within practical limits. The first site, referred to as site A has a

mudline strength of 0 and a strength gradient of 1 kPa/m and the second site, referred to

as site B also has a mudline strength of 0, but with a strength gradient of 2 kPa/m. Both

sites have a sensitivity, St = 5, such that α = 0.2. The factored design mooring line load

is 5 MN. Site A requires an anchor with a follower length Lf = 10.3 m, a plate diameter

D = 4.1 m and a total anchor mass, mt = 75 tonnes, of which the plate = 33 tonnes. Site

B requires a smaller anchor, with a follower length Lf = 9.5 m, a plate diameter D = 3.8

m and a total anchor mass, mt = 59 tonnes, of which the plate = 26 tonnes.

The above example includes a number of simplifying assumptions to demonstrate the

merit of the design approach, but would be refined for an actual design. These include

the loss of embedment during keying, which has been conservatively calculated

assuming vertical loading, whereas the loss of embedment will be lower for inclined

loading (Gaudin et al. 2009, Song et al. 2009, Yu et al. 2009, Wang et al. 2011) and the

0 2 4 6 8 10 12 14

30

25

20

15

10

5

k = 1 kP/a k = 2 kP/a k = 3 kP/a

Site B:

D = 3.8 m, Lf = 9.5 m

mf = 33 t, m

p = 26 t

Site A:

D = 4.1 m, Lf = 10.3 m

mf = 42 t, m

p = 33 t

Etotal

/kdeff

4

z e/d

eff

× 103

= 0.1, 0.2, 0.3, 0.4, 0.5

Page 265: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

231

adopted capacity factor which may need to be reduced for inclined loading at shallow

embedment (Song et al. 2005, Wang and O’Loughlin 2014).

Figure 8.17. Flow chart for DEPLA design approach

8.8 Conclusions

As part of the development path towards qualification of the DEPLA, a series of

centrifuge and reduced scale field trials, supplemented by analytical and numerical

Step 1: Select DEPLA plate diameter, D, thickness, t, and padeye eccentricity, e

Step 2: Calculate the embedment depth loss of the plate, Δze,plate using Equation 8.4

Step 3: Select a capacity factor from Figure 8.10: Nc = 14.9

assuming deep embedment (ze,p,final ≥ 2.5D)

Step 4: Calculate the required follower tip embedment, ze, required for the chosen D value,

using a modified using Equation 8.16

Step 5: Select the other anchor dimensions (Lf, Df, Ltip, t, ts, Hs etc.).

Step 6: Calculate ze,p,final using the expression Equation 8.17

Step 7: Calculate the anchor mass, m, and effective mass, m'

Step 8: Calculate the anchors impact velocity, vi, based on 80% of terminal velocity, vt using Equation 8.18

Step 9: Calculate the total energy, Et, using Equation 8.1

Step 10: Calculate the effective diameter of the anchor Equation 8.18

Step 11: Is ze/deff, sitting on the prediction line for

the required k and α in Figure 8.16?

Step 7: ze,p,final ≥ 2.5D?

Yes

Select

alternative

D

Yes

Preliminary sizing of DEPLA complete

No

No

Select

alternative

D

Page 266: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

232

studies has been undertaken. This chapter considers the results from the final

experimental campaign, in which field trials on a 1:4.5 reduced scale DEPLA were

conducted in approximately 50 m water depth at a site off the west coast of Scotland.

The data gathered in these trials have been considered in parallel with DEPLA field and

centrifuge data from earlier in the experimental campaign to assess the merit of anchor

embedment and capacity models. The anchor embedment model is cast in terms of

inertial drag resistance and strain rate enhanced shear resistance, the latter described

using a power law with separate dependence for bearing and frictional resistance. Back

analysis of measured acceleration data during dynamic embedment indicated that the

power law is capable of describing the strain rate dependence of undrained shear

strength at the very high strain rates associated with the dynamic embedment process,

although complexities associated with initial embedment (up to one anchor length) are

not captured by the model. However, these shortcomings do not appear to unduly lessen

the capability of the embedment model to predict the measured velocity profiles and

importantly the final anchor embedment depth, which is a prerequisite for reliable

calculation of anchor capacity.

Comparison of the field data from Lough Erne and the Firth of Clyde showed that

similar capacity was achieved at both sites, despite the much shallower embedment

from the Firth of Clyde tests associated with the much higher strength gradient. This is

because the capacity of dynamically installed anchors such as the DEPLA is governed

by the soil strength at the final embedment depth. Embedment will be higher in weaker

soils and lower in stronger soils but with approximately the same strength at the final

plate embedment depth in either case.

Back analysis of anchor capacity from each of the experimental campaigns was

considered in terms of a dimensionless capacity factor and compared with

Page 267: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

233

corresponding numerically derived values. This comparison indicated the potential for

soil tension to be lost at the underside of the anchor plate at shallow embedment,

resulting in a lower capacity factor. At deep embedment, where the plate embedment is

at least 2.5 times the diameter, the capacity factor reaches a limiting value in good

agreement with that determined numerically. A simple design procedure, based on a

design chart presented in terms of the total energy of the anchor at impact with the

mudline for a range of strength gradients and soil sensitivities, was shown to be a useful

approach for scaling a DEPLA for a given design load.

Page 268: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

234

CHAPTER 9 CONCLUSIONS

9.1 Summary

The dynamically embedded plate anchor (DEPLA) has been proposed as a cost effective

and technically efficient anchor for deepwater mooring applications. The DEPLA is

rocket or torpedo shaped anchor that comprises a removable central shaft and a set of

four flukes. The DEPLA penetrates to a target depth in the seabed by the kinetic energy

obtained through free-fall in water. After embedment the central shaft is retrieved

leaving the anchor flukes vertically embedded in the seabed. The flukes constitute the

load bearing element as a plate anchor. The DEPLA combines the installation

advantages of dynamically installed anchors (no external energy source or mechanical

operation required during installation) and the capacity advantages of vertical loaded

anchors (sustain significant horizontal and vertical load components).

Despite these advantages there are no current geotechnical performance data for the

DEPLA, as development of the anchor is in its infancy. This thesis has focused on

assessing the geotechnical performance of DEPLAs through an experimental study

involving a centrifuge testing program and an extensive field testing campaign. The

centrifuge tests were carried out on 1:200 reduced scale model DEPLAs in kaolin clay

and are reported in Chapter 4. The centrifuge tests provided early stage proof of concept

for the DEPLA and motivation for subsequent field testing. The tests examined: anchor

impact velocity, initial anchor embedment depth following dynamic installation,

follower recovery load, embedment depth loss due to the keying process and the

subsequent plate anchor capacity.

The field tests were conducted on 1:12, 1:7.2 and 1:4.5 reduced scale DEPLAs

(geometrically similar to those adopted for the centrifuge tests) at two sites: (i) Lough

Page 269: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

235

Erne, which is an inland lake located in County Fermanagh, Northern Ireland and (ii)

Firth of Clyde (1:4.5 scale anchor only) which is located off the West coast of Scotland.

The reduced scale anchors were instrumented with a custom-design, low cost, six

degree of freedom (6DoF) inertial measurement unit (IMU) for measurement of anchor

motion during installation. This represented the first reported use of a 6DoF IMU for a

geotechnical application. The field tests served to validate the centrifuge findings and

demonstrate DEPLA deployment in an aquatic environment. The field tests provided a

basis for understanding the behaviour of DEPLA during free-fall in water and dynamic

embedment in soil (investigated in Chapters 5, 6 and 8), and quantify the loss in

embedment during keying (examined in Chapter 7) and the subsequent plate anchor

capacity (determined in Chapters 7 and 8). Analytical design tools for the prediction of

DEPLA embedment and capacity were verified, refined and calibrated using the

centrifuge and field data.

9.2 Main findings

9.2.1 Inertial measurement unit

This thesis makes a particular contribution to the body of literature on dynamic

penetration problems by presenting and analysing motion data measured by an IMU

during DEPLA installation. Such data are rarely available, but are important for

assessing the mechanisms associated with dynamic penetration in soil, and represents an

essential step towards anchor qualification.

A comprehensive framework for interpreting the IMU measurements so that they are

coincident with a fixed inertial frame of reference was developed and implemented to

establish anchor rotations, accelerations and velocities during free-fall in water and

embedment in soil. Assessment of this framework showed that embedments calculated

from the body frame acceleration measurements, rather than from accelerations

Page 270: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

236

transformed to an inertial frame of reference, led to derived embedment depths that

were in error by up to 27%. In contrast, embedment depths derived from IMU data

interpreted from within an inertial frame of reference through application of the

interpretation framework were shown to be in excellent agreement with independent

direct measurements. The framework was shown to be a suitable means of establishing

anchor embedment.

The IMU measurements verified the appropriateness of an embedment prediction model

for dynamically installed anchors based on strain rate enhanced shear resistance and

fluid mechanics drag resistance, and facilitated refinement and calibration of the model.

9.2.2 Free-fall and dynamic embedment in soil

Back analysis of the motion data measured by the IMU during anchor installation using

the interpretation framework described in Chapter 5, enabled profiles of net anchor

resistance and velocity during free-fall in water and soil penetration to be established.

The main findings from this analysis are summarised in the follow:

The fluid drag resistance acting on the DEPLA during free-fall in water may be

approximated using a constant mean drag coefficient of Cd,a = 0.7 for Re > 1.0x106

(v > 5.9 m/s), which is in good agreement with the upper range of values typically

reported for dynamically installed anchors and free-fall penetrometers.

DEPLA impact velocities were in the ranges: 4.9 to 6.4 m/s (1:12 scale), 2.7 to 8.2

m/s (1:7.2 scale), 8.5 to 10.5 m/s (1:4.5 scale, at Lough Erne) and 5.6 to 12.9 m/s

(1:4.5 scale, at Firth of Clyde), corresponding to anchor release heights in the of: 3

to 8.2 m (4 to 10.9Lf, 1:12 scale), 0.5 to 6.2 m (0.4 to 5Lf, 1:7.2 scale), 6.1 to 9.1 m

(3.1 to 4.6Lf, 1:4.5 scale, at Lough Erne), 2.3 to 42.5 m (1.2 to 21.3Lf, 1:4.5 scale, at

Firth of Clyde). For the larger release heights the DEPLA reached terminal velocity

but then slowed slightly due to the increasing fluid drag resistance afforded by the

Page 271: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

237

increasing length of mooring and follower recovery lines that mobilised in the

water.

DEPLA tip embedment were in the ranges: 1.6 to 1.8 m (2.1 to 2.4Lf, 1:12 scale),

2.8 to 3.7 m (2.2 to 2.9Lf, 1:7.2 scale), 6.4 to 6.9 m (3.2 to 3.5Lf, 1:4.5 scale, Lough

Erne) and 2.6 to 4 m (1.3 to 2Lf, 1:4.5 scale, Firth of Clyde). These tip embedments

are in good agreement with the centrifuge data (1.6 to 2.8Lf) and reported field data

experience of full scale dynamically installed anchors with a broadly similar

geometry. The much shallower embedments measured at the Firth of Clyde tests

reflect the much higher strength gradient at this site.

The results indicated that anchor embedment increases with increasing impact

velocity, which is sensible as the anchor possesses more kinetic energy when it

impacts the soil.

The appropriateness of an embedment prediction model for DEPLA based on strain

rate enhanced shear resistance and fluid mechanics drag resistance was

demonstrated through comparison with the IMU measurements.

For Lough Erne, the best agreement was achieved between the embedment depth

prediction model and the IMU measurements using a power function with a strain

rate parameter of β = 0.06, which is within the typical range of that reported in

variable rate penetrometer testing (β = 0.06 to 0.08), together with friction ratios of

α = 0.46 (1:12 scale), 0.45 (1:7.2 scale) and 0.32 (1:4.5 scale). These α values are

generally consistent with the range inferred from cyclic piezocone tests. In the case

of Firth of Clyde, the best agreement was achieved using a power law strain

parameter of β = 0.08 and α = 0.26 (average value inferred from cyclic piezocone

tests).

Qualitatively, the embedment depth prediction model captures the dynamic response

adequately for both test locations and is capable of predicting the final embedments

Page 272: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

238

to within 10% of the measurements using parameters that are routinely used

offshore geotechnical design, in conjunction less common parameters (β and Cd,a)

which are both related to the increase in penetration resistance with increasing

anchor velocity. These ‘dynamic’ parameters account for the increase in

geotechnical resistance due to strain-rate effects (β) and the increase in fluid

mechanics resistance due to pressure drag (Cd,a).

Back analysis IMU data indicated that the power function is capable of describing

the strain rate dependence of undrained shear strength at the very high strain rates

associated with the dynamic embedment process, although complexities associated

with initial embedment (up to one anchor length) are not captured by the model.

However, these shortcomings do not appear to unduly lessen the capability of the

embedment model to predict the measured velocity profiles and importantly the

final anchor embedment depth.

ROV video captured at Firth of Clyde showed an open crater at the anchor drop sites

which suggest that the hole caused by the passage of the anchor remains open. This

observation is consistent with numerical analysis and other experimental work,

which suggests that for the strength ratios encountered at this location, the hole will

remain open to an anchor tip embedment of 3.85 m, which is the limit of the

embedments achieved at this test location. Hence, an open hole was assumed in the

embedment depth prediction model.

9.2.3 Follower extraction

DEPLA follower extraction was conducted in the centrifuge and at the Lough Erne test

site. The maximum follower extraction resistance was 2.2 to 3 times the follower dry

mass in the centrifuge tests and 2.9 to 4.1 times the follower dry mass in Lough Erne.

These ranges reflect the range in tip embedment depths achieved and hence the range in

undrained shear strength mobilised during follower extraction. The follower extraction

Page 273: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

239

resistance is slightly higher in Lough Erne than measured in the centrifuge tests due to

the higher tip embedments achieved in the field tests.

9.2.4 Plate anchor keying response and capacity

The keying and pullout of reduced scale model DEPLAs was successfully carried out in

the centrifuge and at both field locations. The embedment depth loss of the plate anchor

during keying was quantified, and anchor capacity for a given plate geometry and

seabed strength profile was determined. The main findings are as summarised in the

follow:

Results for dynamic and jacked DEPLA installations were similar in both the

centrifuge and field indicating that the dynamic installation process does not affect

either the keying response or the plate anchor capacity.

The keying response in the field was qualitatively consistent with the DEPLA

centrifuge data, and data from tests on square anchors installed on the vertical

orientation. However, despite the geometrical similarity between the field and

centrifuge DEPLAs, the loss in embedment, ∆ze,plate = 1D to 1.8D in Lough Erne

(could not be determined in Firth of Clyde) was much higher than that in the

centrifuge tests (∆ze,plate = 0.5D to 0.7D). This was attributed to the soil strength

ratio, su0/γ'D, which was higher in the field tests (su0/γ'D = 4.5 to 7.5 in Lough Erne)

than in the centrifuge tests (su0/γ'D = 0.6 to 1.4) as numerical analyses has shown

that higher local strength results in higher embedment depth loss during keying.

The field tests showed that the plate anchor capacity, Fvp, increased linearly with

depth, which reflects the linear increase in undrained shear strength at both test

locations.

Similar anchor capacity was achieved at both test sites despite the much shallower

embedment from the Firth of Clyde tests associated with the much higher strength

gradient. This is because the capacity of dynamically installed anchors such as the

Page 274: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

240

DEPLA is governed by the soil strength at the final embedment depth. This will be

higher in weaker soils and lower in stronger soils but with approximately the same

strength at the final plate embedment depth in either case.

Field test data indicated that the end of keying coincides with the peak anchor

capacity and the plate anchor bearing capacity factor, Nc, decreases with reducing

load inclination for plate embedment ratios less than four.

Back analysed peak bearing capacity factors from the centrifuge and field tests were

in the range Nc = 14.2 to 15.8 (centrifuge), Nc =14.3 to 14.6 (Lough Erne) and Nc =

4.2 to 12.9 (Firth of Clyde).

At deep embedment, where the plate embedment is at least 2.5D, the bearing

capacity factors derived from the centrifuge and field results reach limiting values

that are in quite good agreement with numerical no-breakaway values for DEPLA.

These are much higher than the exact solution for a deeply embedded vanishingly

thin smooth circular plate. This is due to the combined effect of the fluke thickness

and cruciform arrangement on the DEPLA sleeve, which results in a larger failure

surface (and hence a higher capacity factor) compared with a flat vanishingly thin

circular plate. At shallow embedments where the plate embedment is at less than

2.5D, the experimental capacity factors lie within the bounds of the numerical

immediate breakaway and no-breakaway cases for DEPLA. This is because at

shallow embedment the failure mechanism extends to the soil surface, such that the

gap formed during breakaway may be filled with the free water above the mudline.

A simple design procedure, based on a design chart presented in terms of the total

energy of the anchor at impact with the mudline for a range of strength gradients

and soil sensitivities, was shown to be a useful approach for scaling a DEPLA for a

given design load.

Page 275: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

241

REFERENCES

Abelev, A. and Valent, P. (2009). Strain-rate dependency of strength and viscosity of

soft marine deposits of the Gulf of Mexico. Proceedings of MTS/IEEE Oceans 2009

Conference, Biloxi, Mississippi, USA, 26–29 October 2009, pp. 1–8.

Abelev, A., Simeonov, J. and Valent, P. (2009a). Numerical investigation of dynamic

free-fall penetrometers in soft cohesive marine sediments using a finite element

approach. Proceedings of MTS/IEEE Oceans 2009 Conference, Biloxi, Mississippi,

USA, 26–29 October 2009, pp. 1–10.

Abelev, A., Tubbs, K. and Valent, P. (2009b). Numerical investigation of dynamic free-

fall penetrometers in soft cohesive marine sediments using a finite difference

approach. Proceedings of MTS/IEEE Oceans 2009 Conference, Biloxi, Mississippi,

USA, 26–29 October 2009, pp. 1–9.

Adams, J. I. and Hayes, D. C. (1967). The uplift capacity of shallow foundations.

Ontario Hydro Research Quarterly, Vol. 19, No. 1, pp. 1–13.

Akal, T. and Stoll, R. D. (1995). An expendable penetrometer for rapid assessment of

seafloor parameters. Proceedings of MTS/IEEE Oceans 1995 Conference, San

Diego, California, USA, 9–12 October 1995, pp. 1822–1826.

Ali, J. I. (1968). Pullout resistance of anchor plates and anchored piles in soft bentonite

clay. MSc Thesis, Duke University, North Carolina, USA.

Allmond, J. D., Hakhamaneshi, M., Wilson, D. W. and Kutter, B. L. (2014) Advances

in measuring rotation with MEMS accelerometers. Proceedings of the 8th

International Conference on Physical Modelling in Geotechnics, Perth, Australia,

14–17 January 2014, Vol. 1, pp. 353–359.

Page 276: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

242

Andersen, K. H., Murff, J. D., Randolph, M. F., Cluckey, E. C., Erbrich, C. T., Jostad,

H. P., Hansen, B., Aubeny, C., Sharma, P. and Supachawarote, C. (2005). Suction

anchors for deepwater applications. Proceedings of the 1st International Symposium

on Frontiers in Geotechnics, Perth, Ausralia, 19–21 September 2005, pp. 3–30.

Araujo, J. B., Machado, R. D. and Medeiros, C. J. (2004). High holding power torpedo

pile – results from the first long term application. Proceeding of the 23rd

International Conference on Offshore Mechanics and Arctic Engineering,

Vancouver, British Columbia, Canada, 20–25 June 2004, Vol. 1, pp. 417–421.

Aubeny, C. P. and Shi, H. (2006). Interpretation of impact penetration measurements in

soft clays. Journal of Geotechnical and Geoenvironmental Engineering, Vol. 132,

No. 6, pp. 770–777.

Aubeny, C. P., Murff, J. D. and Roesset, J. M. (2001). Geotechnical issues in deep and

ultra deep waters. International Journal of Geomechanics, Vol. 1, No. 2, pp. 225–

247.

Audibert, J. M. E., Movant, M. N., Won, J. Y, and Gilbert, R. B. (2006). Toprepdo

piles: laboratory and field research. Proceedings of the 16th International Offshore

and Polar Engineering Conference, San Francisco, California, USA, 28 May–2 June

2006. pp. 135–142.

Beard, R. M. (1977). Expendable doppler penetrometer: a performance evaluation,

Technical Report R855, Civil Engineering Laboratory, Port Hueneme, California,

USA.

Beard, R. M. (1981). A penetrometer for deep ocean seafloor exploration. Proceedings

of MTS/IEEE Oceans 1981 Conference, Boston, Massachusetts, USA, 16–18

September 1981, Vol. 13, pp. 668–673.

Page 277: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

243

Beard, R. M. (1985). Expendable doppler penetrometer for deep ocean sediment

measurements. ASTM Special Technical Publication 883, pp. 101–124.

Bennett V., Abdoun T., Shantz T., Jang D. and Thevanayagam, S. (2009) Design and

characterization of a compact array of MEMS accelerometers for geotechnical

instrumentation. Smart Structures and Systems, Vol. 5, No. 6, pp. 663–679.

Biscontin, G. and Pestana, J. M. (2001). Influence of peripheral velocity on vane shear

strength of an artificial clay. Geotechnical Testing Journal, Vol. 24, No. 4, pp. 423–

429.

Bjerrum, L. (1973). Problems of soil mechanics and construction on soft clays and

structurally unstable soils (collapsible, expansive and others). Proceedings of the 8th

International Conference on Soil Mechanics and Foundation Engineering, Moscow,

Russia, 2–3 August 1973, Vol. 3, pp. 111–159.

Blake, A. P. and O’Loughlin, C. D. (2012). Field testing of a reduced scale dynamically

embedded plate anchor. Proceedings of the 7th International Conference in Offshore

Site Investigation and Geotechnics, London, United Kingdom, 12–14 September

2012. pp. 621–628.

Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate

anchors as assessed through field tests. Canadian Geotechnical Journal, Vol. 52, No.

9, pp. 1270–1282.

Blake, A. P., O’Loughlin, C. D. and C. Gaudin (2014). Capacity of dynamically

embedded plate anchors as assessed through field tests. Canadian Geotechnical

Journal, Vol. 52, No. 1, pp. 87–95.

Blake, A. P., O’Loughlin, C. D. and Gaudin, C. (2010). Setup following keying of plate

anchors assessed through centrifuge tests in kaolin clay. Proceedings of the 2nd

Page 278: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

244

International Symposium on Frontiers in Offshore Geotechnics, Perth, Australia, 8–

10 November 2010. pp. 705–710.

Blake, A. P., O’Loughlin, C. D., Morton, J. M., O’Beirne, C., Gaudin, C. and White D.

J. (2016). In -situ measurement of the dynamic penetration of free-fall projectiles in

soft soils using a low cost inertial measurement unit. Geotechnical Testing Journal,

forthcoming.

Boukpeti, N., White, D. J. and Randolph, M. F. (2012). Analytical modelling of the

steady flow of a submarine slide and consequent loading on a pipeline.

Géotechnique, Vol. 62, No. 2, pp. 137–146.

Bowman, L., March, R., Orenberg, P., True, D. and Herrmann, H. (1995). Evaluation of

dropped versus static cone penetrometers at a calcareous cohesive site. Proceedings

of MTS/IEEE Oceans 1995 Conference, San Diego, California, USA, 9–12 October

1995, Vol. 3, pp. 1846–1858.

Boylan, N., Long, M., Ward, D., Barwise, A. and Georgious, B. (2007). Full flow

penetrometer testing in Bothkennar clay. Proceedings of the 6th International

Offshore and Site Investigation and Geotechnics Conference, London, United

Kingdom, 11–13 September 2007, pp. 177–186.

Brandão, F. E. N., Henriques, C. C. D., Araújo, J. B., Ferreira, O. C. G. and Amaral, C.

D. S. (2006). Albacora Leste field development - FPSO P-50 mooring system

concept and installation. Proceedings of the Offshore Technology Conference,

Houston, Texas, USA, 1–4 May 2006, OTC 18243.

Brown, R. P., Wong, P. C. and Audibert, J. M. (2010). SEPLA keying prediction

method based on full-scale offshore tests. Proceedings of the 2nd International

Page 279: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

245

Symposium on Frontiers in Offshore Geotechnics, Perth, Australia, 8–10 November

2010, pp. 717–722.

Carter, J. P. and Nazem, M. (2013). Analysis of dynamic penetration of soils.

Proceedings of the 22nd Australian Conference on the Mechanics of Structures and

Materials, Sydney, Australia, 11–14 December 2013, pp. 3–14.

Carter, J. P., Nazem, M., Airey, D. W. and Chow, S. H. (2010). Dynamic analysis of

free-falling penetrometers in soil deposits. Proceedings of GeoFlorida 2010

Conference, West Palm Beach, Florida, USA, 20–24 February 2010, pp. 53–68.

Casagrande, A. and Wilson, S. D. (1951). Effect of rate of loading on the strength of

clays and shales at constant water content. Géotechnique, Vol. 2, No. 3, pp. 251–263.

Cassidy, M. J., Gaudin, C., Randolph, M. F., Wong, P. C., Wang, D. and Tian, Y.

(2012). A plasticity model to assess the keying of plate anchors. Géotechnique, Vol.

62, No. 9, pp. 825–836.

Cenac Π, W. A. (2011). Vertically loaded anchor: drag coefficient, fall velocity and

penetration depth using laboratory measurements. Master’s thesis, Texas A&M

University, College Station, USA.

Chari, T. R., Smith, W. G. and Zielinski, A. (1978). Use of free fall penetrometer in sea

floor engineering. Proceedings of MTS/IEEE Oceans 1978 Conference, Washington,

District of Columbia, USA, 6–8 Septmebr 1978, Vol. 3, pp. 686–691.

Chen, Z., Tho, K. K., Leung, C. F. and Chow, Y. K. (2013). Influence of overburden

pressure and soil rigidity on uplift behaviour of square plate anchor in uniform clay.

Computer and Geotechnics, Vol. 52, pp. 71–81.

Page 280: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

246

Chow, S. H. (2013). Free-falling penetrometer tests in clay. PhD thesis, The University

of Sydney, Sydney, Australia.

Chow, S. H. and Airey, D. H. (2010). Laboratory free falling penetrometer test into

clay. Proceeding of the 2nd International Symposium on Frontiers in Offshore

Geotechnics, Perth, Australia, 8–10 November 2010, pp. 265–270.

Chow, S. H. and Airey, D. W. (2013). Soil strength characterisation using free-falling

penetrometers. Géotechnique, Vol. 63, No. 13, pp. 1131–1143.

Chow, S. H. and Airey, D. W. (2014). Free-falling penetrometers: a laboratory

investigation. Journal of Geotechnical and Geoenvironmental Engineering, Vol.

140, No. 1, pp. 201–214.

Chow, S. H., O’Loughlin, C. D. and Randolph, M. F. (2014). Soil strength estimation

and pore pressure dissipation for free-fall piezocones in soft clay. Géotechnique, Vol.

64, No. 10, pp. 817–827.

Chung, S. F. and Randolph, M. F. (2004). Penetration resistance in soft clay for

different shaped penetrometers. Proceedings of the 2nd International Conference on

Site Characterisation, Porto, Portugal, Vol. 1, pp. 671–677.

Chung, S. F., Randolph, M. F. and Schneider, J. A. (2006). Effect of penetration rate on

penetrometer resistance in clay. Journal of Geotechnical and Geoenvironmental

Engineering, Vol. 132, No. 9, pp. 1188–1196.

Cilingir, U. and Madabhushi, S. P. G. (2011). A model study on the effects of input

motion on the seismic behaviour of tunnels. Soil Mechanics and Earthquake

Engineering, Vol. 31, No. 3, pp. 452–462.

Page 281: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

247

Colp, J. L., Caudle, W. N. and Schuster, C. L. (1975). Penetrometer system for

measuring in situ properties of marine sediment. Proceedings of MTS/IEEE Oceans

1975 Conference, San Diego, California, USA, 22–25 September 1975, pp. 405–411.

Colreavy, C., O’Loughlin, C. D. and Ward, D. (2012). Piezoball tests in a soft lake

sediments. Proceedings of the 4th International Conference on Geotechnical and

Geophysical Site Characterisation, Porto de Galinhas, Brazil, 18–21 September

2012, pp. 597–602.

Das, B. M. (1978). Model tests for uplift capacity of foundations in clay. Soils and

Foundations, Vol. 18, No. 2, pp. 17–24.

Das, B. M. (1980). A procedure for estimation of ultimate capacity of foundations in

clay. Soils and Foundations, Vol. 20, No. 10, pp. 77–82.

Das, B. M. and Puri, V. K. (1989). Holding capacity of inclined square plate anchors in

clay. Soils and Foundations, Vol. 29, No. 3, pp. 138–144.

Das, B. M. and Singh, G. (1994). Uplift capacity of plate anchor in clay. Proceedings of

the 4th International Offshore and Polar Engineering Conference, Osaka, Japan, 10–

15 April 1994. Vol. 1, pp. 436–442.

Das, B. M., Shin, E. C. and Dallo, K. F. (1985b). Ultimate pullout capacity of shallow

vertical anchors in clay. Soils and Foundations, Vol. 25, No. 2, pp. 148–152.

Das, B. M., Shin, E. C., Dass, R. N. and Omar, M. T. (1994). Suction force below plate

anchors in soft clay. Marine Georesources and Geotechnology, Vol. 12, No. 1, pp.

71–81.

Page 282: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

248

Das, B. M., Tarquin, A. J. and Moreno, R. (1985a). Model tests for pullout resistance of

vertical anchors clay. Civil Engineering for Practicing Engineers, Pergamon, New

York, USA, Vol. 4, No. 1, pp. 191–209.

Davie, J. M. and Sutherland, H. B. (1977). Uplift resistance of cohesive soils. Journal of

the Soil Mechanics and Foundations Division, Vol. 103, No. 9, pp. 935–952.

Dayal, U. (1974). Instrumented impact cone penetrometer. PhD thesis, Memorial

University of Newfoundland, St. John’s, Newfoundland, Canada.

Dayal, U. and Allen, J. H. (1973). Instrumented impact cone penetrometer. Canadian

Geotechnical Journal, Vol. 10, No. 3, pp. 397–409.

Dayal, U., Allen, J. H. and Jones, J. (1975). Use of an impact penetrometer for the

evaluation of the in-situ strength of marine sediments. Marine Georesources and

Geotechnology, Vol. 1, No. 2, pp. 73–89.

Deglo de Besses, B. D., Magnin, A. and Jay, P. (2003).Viscoplastic flow around a

cylinder in an infinite medium. Journal of Non-Newtonian Fluid Mechanics, Vol.

115, pp. 27–49.

Denness, B., Berry, A., Darwell, J. and Nakamura, T. (1981). Dynamic seabed

penetration. Proceeding of MTS/IEEE Oceans 1981 Conference, Boston,

Massachusetts, USA, 16–18 September 1981, pp. 662–667.

Department of Industry and Resources (2008). Western Australian Oil and Gas Review

2008, Department of Industry and Resources, Government of Western Australia,

Perth, Australia. Accessed 9 June 2014 at http://www.dmp.wa.gov.au/

Page 283: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

249

Douglas, B. L. and Wapner, M. P. (1996). Test results of the doppler penetrometer

seafloor sediment profiling system. Proceeding of MTS/IEEE Oceans 1996

Conference, Fort Lauderdale, Florida, USA, 23–26 September, Vol. 2, pp. 722–727.

Dove, P., Treu, H. and Wilde, B. (1998). Suction embedded plate anchor (SEPLA): A

new Anchoring Solution for ultra-deep water mooring. Proceeding of the 10th Deep

Offshore Techology Conference, New Orleans, Louisiana, USA, 17–19 November

1998.

Ehlers, C. J., Young, A. G. and Chen, J. (2004). Technology assessment of deepwater

anchors. Proceedings of the Offshore Technology Conference, Houston, Texas, USA,

3–6 May 2004, OTC 16840.

Einav, I. and Randolph, M. F. (2006). Effect of strain rate on mobilised strength and

thickness of curved shear bands. Géotechnique, 56(7): 501–504.

Einav, I., Klar, A., O'Loughlin, C. D. and Randolph, M. F. (2004). Numerical modelling

of Deep Penetrating Anchors. Proceedings of the 9th Australian and New Zealand

Conference on Geomechanics, Auckland, New Zealand, 8–11 February 2004, Vol. 2,

pp. 818–824.

Eltaher, A., Rajapaksa, Y. and Chang, K. T. (2003). Industry trends for design of

anchoring systems for deepwater offshore structures. Proceeding of the Offshore

Technology Conference, Houston, Texas, USA, 5–8 May 2003. OTC 15265.

Fernandes, A. C., Araujo, J. B., Lima de Almeida, J. C., Machado, R. D. and Matos, V.

(2006). Torpedo anchor installation hydrodynamics. Journal of Offshore Mechanics

and Artic Engineering, Vol. 128, No. 11, pp. 286–293.

FloaTEC (2012). Deepwater Solutions and Records for Concept Selection, FloaTEC,

Houston, Texas, USA. Accessed 9 June 2014 at http://www.offshore-mag.com

Page 284: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

250

Fossen, T. I. (2011). Handbook of marine craft hydrodynamics and motion control.

John Wiley & Sons, Ltd., Chichester, United Kingdom.

Freeman, T. J. and Schuttenhelm, R. T. E. (1990). A geotechnical investigation of a

deep ocean site. Geotechnical Engineering of Ocean Waste Disposal, ASTM STP

1087, pp. 255–275.

Freeman, T. J., Murray, C. N. and Schuttenhelm, R. T. E. (1988). The Tyro 86

penetrator experiments at Great Meteor East. Proceedings of Advances in

Underwater Technology, Ocean Science and Offshore Engineering, Vol. 16, pp.

217–226.

Freeman, T. J., Murray, C. N., Francis, T. J. G., McPhail, S. D. and Schultheiss, P. J.

(1984). Modelling radioactive waste disposal by penetrator experiments in the

abyssal Atlantic Ocean. Letters to Nature, Vol. 310, pp. 130–133.

Freeman, T.J. and Burdett, J. R. F. (1986). Deep ocean model penetrator experiments.

Commission of the European Communities Nuclear Science and Technology Report

EUR 10502, Commission of the European Communities, Brussels, Luxembourg.

Furlong, A., Osler, J., Christian, H., Cunningham, D. and Pecknold, S. (2006). The

Moving Vessel Profiler (MVP) - a rapid environmental assessment tool for the

collection of water column profiles and sediment classification. Proceedings of the

Undersea Defence Technology Pacific Conference, San Diego, California, USA.

Gaudin, C., O’Loughlin, C. D., Hossain, M. S and Zimmerman, E. H. (2013). The

performance of dynamically embedded anchors in calcareous silt. Proceeding of the

32nd International Conference on Offshore Mechanics and Arctic Engineering,

Nantes, France, 9–14 June 2013, Vol. 6, OMAE2013-10115.

Page 285: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

251

Gaudin, C., O'loughlin, C. D., Randolph, M. F. and Lowmass, A. C. (2006). Influence

of the installation process on the performance of suction embedded plate anchors.

Géotechnique, Vol. 56, No. 6, pp. 381–391.

Gaudin, C., Simkin, M., White, D. J. and O’Loughlin, C. D. (2010). Experimental

investigation into the influence of a keying flap on the keying behaviour of plate

anchor. Proceedings of the 20th International Offshore and Polar Engineering

Conference, Beijing, China, 20–26 June 2010. Vol. 2, pp. 533–540.

Gaudin, C., Tham, K. H. and Ouahsine, S. (2009). Keying of plate anchors in NC clay

under inclined loading. International Journal of Offshore and Polar Engineering,

Vol. 19, No. 2, pp. 135–142.

Gibson, G. C. and Coyle, H. M. (1968). Soil damping constants related to common soil

properties in sands and clays. Technical Report 125-1, Texas Transportation

Institute, College Station, Texas, USA.

Gilbert, R. B., Morvant, M. N. and Audibert, J. M. E. (2008). Torpedo piles joint

industry project – model torpedo pile tests in kaolinite test beds. Final Project

Report, Report No. 02/08B187, Offshore Technology Research Center, The

University of Texas at Austin, Austin, Texas, USA.

Graham, J., Crooks, J. H. A. and Bell, A. L. (1983). Time effects on the stress–strain

behaviour of natural soft clays. Géotechnique, Vol. 33, No. 3, pp. 327–340.

Hasanloo, D., Pang, H. and Yu, G. (2012). On the estimation of the falling velocity and

drag coefficient of torpedo anchor during acceleration. Ocean Engineering, Vol. 42,

pp. 135–146.

Hoerner, S. F. (1965). Fluid-dynamic drag, practical information on aerodynamic drag

and hydrodynamic resistance. Hoerner Fluid Dynamics, Bakersfield, California, USA.

Page 286: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

252

Hossain, M. S, Kim, Y. and Wang, D. (2013). Physical and numerical modelling of

installation and pull-out of dynamically penetrating anchors in clay and silt.

Proceeding of the 32nd International Conference on Offshore Mechanics and Arctic

Engineering, Nantes, France, 9–14 June 2013, Vol. 6, OMAE2013-10322.

Hossain, M. S. and Randolph, M.F. (2009). Effect of strain rate and strain softening on

the penetration resistance of spudcan foundations on clay. International Journal of

Geomechanics, Vol. 9, No. 3, pp. 122–132.

Hossain, M. S., Kim, Y. and Gaudin, C. (2014). Experimental investigation of

installation and pullout of dynamically penetrating anchors in clay and silt. Journal

of Geotechnical and Geoenvironmental Engineering, Vol. 140, No. 7, pp. 04014026.

House, A. R., Oliveira, J. R. M. S. and Randolph, M. F. (2001). Evaluating the

coefficient of consolidation using penetration tests. International Journal of Physical

Modelling in Geotechnics, Vol. 1, No. 3, pp. 17–25.

International Energy Agency (2011). Medium-Term Oil and Gas Markets Report 2011,

International Energy Agency, Paris, France. Accessed 9 June 2014 at

http://www.iea.org

International Energy Agency (2012). World Energy Outlook 2012, International Energy

Agency, Paris, France. Accessed 9 June 2014 at http://www.iea.org

International Energy Agency (2013). Medium-Term Oil Market Report 2013,

International Energy Agency, Paris, France. Accessed 9 June 2014 at

http://www.iea.org

Jeanjean, P. (2006). Setup characteristics of suction anchors for soft Gulf of Mexico

clays: experience from field installation and retrieval. Proceedings of the Offshore

Technology Conference, Houston, Texas, USA, 1–4 May 2006, OTC 18005.

Page 287: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

253

Jeanjean, P., Spikula, D. and Young, A. (2012). Technical vetting of free-fall cone

penetrometer. Proceedings of the 7th International Conference in Offshore Site

Investigation and Geotechnics, London, United Kingdom, 12–14 September 2012,

pp. 179–186.

Jeong, S. W, Leroueil, S. and Locat, J. (2009). Applicability of power law for

describing the rheology of soils of different origins and characteristics. Canadian

Geotechnical Journal, Vol. 46, No. 9, pp. 1011–1023.

Jonkman, J. M. (2007). Dynamics modelling and loads analysis of an offshore floating

wind turbine. Technical Report NREL/TP-500-41958, National Renewable Energy

Laboratory, Golden, Colorado, USA.

Kestin, J., Sokolov, M, and Wakeham, A. 1978. Viscosity of liquid water in the range -

8°C to 150°C. Journal of Physical Chemistry, Vol. 7, No. 3, 941–948.

King, K., Yoon, S. W., Perkins, N. C. and Najafi, K. (2008). Wireless MEMS inertial

sensor system for golf swing dynamics. Sensors and Actuators, Vol. 141, No. 2, pp.

619–630.

Kunitaki, D., de Lima, B., Evsukoff, A. G. and 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, pp. 512343.

Kupferman, M. (1965). The vertical holding capacity of marine anchors in clay

subjected to static and cyclic loading. Master’s thesis, University of Massachusetts,

Massachusetts, USA.

Lefebvre, G. and LeBoeuf, D. (1987). Rate effects and cyclic loading of sensitive clays.

Journal of Geotechnical Engineering, Vol. 113, No. 5, pp. 476–489.

Page 288: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

254

Lehane, B. M., O’Loughlin, C. D., Gaudin, C. and Randolph, M. F. (2009). Rate effects

on penetrometer resistance in kaolin. Géotechnique, Vol. 59, No. 1, pp. 41–52.

Levacher, D. (1985). Penetrometre a chute libre: impact et penetration dans des argiles

reconstituees. Canadian Geotechnical Journal, Vol. 22, No. 1, pp. 129–135.

Levy, F. M. and Richards, D. J. (2012) Rapid soil displacements from MEMS

accelerometers, Proceedings of the 30th IMAC Conference on Structural Dynamics,

Jacksonville, Florida, USA, 30 January–2 February 2012, Vol. 1, pp. 197–208.

Lieng, J. T., Hove, F. and 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, pp. 613–619.

Lieng, J. T., Kavli, A., Hove, F. and 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,

USA, 28 May–2 June 2000, pp. 410–416.

Lieng, J. T., Tjelta, T . I. and Skaugset, K. (2010). Installation of two prototype deep

penetrating anchors at the Gjøa field in the North Sea. Proceedings of the Offshore

Technology Conference, Houston, USA, 3–6 May 2010, OTC 20758.

Lott, D. F. and Poeckert, R. H. (1996). Extending Cooperative Research: Canada, New

Zealand, USA in joint effort in British Columbia to evaluate penetrometers for

groundtruthing acoustic classifiers for mine countermeasures. Sea Technology, Vol.

37, No. 9, pp. 56–61.

Page 289: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

255

Low, H. E., Lunne, T., Andersen, K. H., Sjursen, M. A., Li, X and Randolph, M. F.

(2010). Time effects on the stress–strain behaviour of natural soft clays.

Géotechnique, Vol. 60, No. 11, pp. 843–859.

Low, H. E., Randolph M. F., De Jong, J. T., and Yafrate, N. J. (2008). Variable rate full

flow penetration tests in intact and remoulded soil. Proceedings of the 3rd

International Conference on Geotechical and Geophysical Site Characterization,

Taipei, Taiwan, 1–4 April 2008, pp. 1087–1092.

Lunne, T. and Andersen, K. H. (2007). Soft clay shear strength parameters for

deepwater geotechnical design. Proceedings of the 6th International Offshore and

Site Investigation and Geotechnics Conference, London, United Kingdom, 11–13

September 2007, pp. 151–176.

Martin, C. M. and Randolph, M. F. (2001). Applications of the lower and upper bound

theorems of plasticity to collapse of circular foundations. Proceedings of the 10th

International Conference of the International Computer Methods and Advances in

Geomechanics, Tucson, Arizona, USA, 7–12 January 2001, Vol. 2, pp. 1417–1428.

Massey, S. J. (2000). Feasibility study of a new concept for the anchorage of offshore

floating production platforms. Honours thesis, Curtin University of Technology,

Perth, Australia.

Medeiros, C. J. (2001). Torpedo anchor for deep water. Proceedings of the Deepwater

Offshore Technology Conference, Rio de Janeiro, Brazil, October 2001.

Medeiros, C. J. (2002). Low cost anchor system for flexible risers in deep waters.

Proceedings of the Offshore Technology Conference, Houston, Texas, USA, 6–9

May 2002, OTC 14151.

Page 290: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

256

Merifield, R. S., Lyamin, A. V. and Sloan, S. W. (2005). Stability of inclined strip

anchors in purely cohesive soil. Journal of Geotechnical and Geoenvironmental

Engineering, Vol. 131, No. 6, pp. 792–798.

Merifield, R. S., Lyamin, A., Sloan, S. W. and Yu, H. S. (2003). Three-dimensional

lower bound solutions for stability of plate anchors in clay. Journal of Geotechnical

and Geoenvironmental Engineering, Vol. 129, No. 3, pp. 243–253.

Merifield, R. S., Sloan, S. W. and Yu, H. S. (2001). Stability of plate anchors in

undrained clay. Géotechnique, Vol. 51, No. 2, pp. 141–153.

Meyerhof, G. G. (1973). Uplift resistance of inclined anchors and piles. Proceedings of

the 8th International Conference on Soil Mechanics and Foundation Engineering,

August 1973, Moscow, Vol. 2, pp. 167–172.

Meyerhof, G. G. and Adams, J. I. (1968). The ultimate uplift capacity of foundations.

Canadian Geotechnical Journal, Vol. 5, No. 4, pp. 225–244.

Morison, J. R., O'Brien, M. P., Johnson, J. W. and Schaaf, S. A. (1950). The force

exerted by surface waves on piles. Petroleum Transactions, Vol. 189, pp. 149–154.

Morton, J. P. and O’Loughlin, C. D. (2012). Dynamic penetration of a sphere in clay.

Proceedings of the 7th International Conference in Offshore Site Investigation and

Geotechnics, London, United Kingdom, 12–14 September 2012, pp. 223–230.

Morton, J. P., O’Loughlin, C. D. and White, D. J. (2014). Strength assessment during

shallow penetration of a sphere in clay. Géotechnique Letters, Vol. 4, No. 4, pp. 262–

266.forthcoming.

Morton, J. P., O’Loughlin, C. D. and White, D. J. (2015). Estimation of soil strength by

instrumented free-fall sphere tests. Géotechnique, forthcoming.

Page 291: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

257

Mulhearn, P. J. (2003). Influences of penetrometer tip geometry on bearing strength

estimates. International Journal of Offshore and Polar Engineering, Vol. 13, No. 1,

pp. 73–78.

Mulhearn, P. J., Morgan, S., Poeckert, R. H. and Holtkamp, R. H. (1999). Penetrometer

comparison trials, Sydney Harbour, March 1998. TTCP Report, DSTO, Australia.

Mulhearn, P. J., Poeckert, R. H. and Holtkamp, R. H. (1998). Report on penetrometer

comparisons from TTCP MAR-TP-13’s MCM Australian sea trials 1997 (MAST97)

off Cairns, Great Barrier Reef, Australia. TTCP Report, DSTO, Australia.

Mulukutla, G. K. (2009). Determination of geotechnical properties of seafloor

sediment. PhD thesis, University of New Hampshire, New Hampshire, USA.

Mulukutla, G. K., Huff, L. C., Melton, J. S., Baldwin, K. C. and Mayer, L. A. (2011).

Sediment identification using free fall penetrometer acceleration-time histories.

Marine Geophysical Research, Vol. 32, pp. 397–411.

Murff, J. D., Randolph, M. F., Elkhatib, S., Kolk, H. J., Ruinen, R. M., Strom, P. J. and

Thorne, C. P. (2005). Vertically loaded plate anchors for deepwater applications.

Proceedings of the 1st International Symposium on Frontiers in Offshore

Geotechnics, Perth, Australia, 19–21 September 2005, pp. 313–48.

Murray, C. N. (1988). The disposal of heat-generating nuclear waste in deep ocean

geological formation: A feasible option. Proceedings of the Advances in Underwater

Technology, Ocean Science and Offshore Engineering Conference, Vol. 16, pp. 227–

233.

Nakase, A. and Kamei, T. (1986). Influence of strain rate on undrained shear

characteristics of Ko-consolidated cohesive soils. Soils and Foundations, Vol. 26,

No. 1, pp. 85–95.

Page 292: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

258

Nazem, M., Carter, J. P., Airey, D. W. and Chow, S. H. (2001), Dynamic Analysis of a

Smooth Penetrometer Free-Falling into Uniform Clay. Géotechnique, Vol. 62, No.

10, pp. 893–905.

Nazem, M., Carter, J. P., Airey, D. W. and Chow, S. H. (2012). Dynamic analysis of a

smooth penetrometer free falling into uniform clay. Géotechnique, Vol. 62, No. 10,

pp. 893–905.

NCEL. (1985). Handbook for Marine Geotechnical Engineering. US Naval Civil

Engineering Laboratory, Port Hueneme, California, USA.

Nebot, E. and Durrant-Whyte, H. (1999). Initial calibration and alignment of low cost

inertial navigation units for land vehicle applications. Journal of Robotics Systems,

Vol. 16, No. 2, pp. 81–92.

Nixon, L. D., Shepard, N. K., Bohannon, C. M., Montgomery, T. M., Kazanis, E. G.

and Gravois, M. P. (2009). Deepwater Gulf of Mexico 2009: Interim Report of 2008

Highlights. Report No. MMS 2009-016, USA Department of the Interior, Minerals

Management Service, Gulf of Mexico OCS Region, New Orleans, Louisiana, USA.

Noureldin, A., Karamat, T. B. and Georgy, J. (2012). Fundamentals of inertial

navigation, satellite-based positioning and their integration. Springer-Verlag, Berlin,

Germany.

O’Loughlin, C. D., Blake, A. P. and Gaudin, C. (2016). A simpleTowards a simple

design approach for dynamically embedded plate anchorsDEPLA. Géotechnique,

Unpublished manuscriptUnder review.

O’Loughlin, C. D., Blake, A. P., Richardson, M. D., Randolph, M. F., and Gaudin, C.

(2014b). Installation and capacity of dynamically installed plate anchors as assessed

through centrifuge tests. Ocean Engineering, Vol. 88, pp. 204–213.

Page 293: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

259

O’Loughlin, C. D., Blake, A. P., Wang, D., Gaudin, C. and Randolph, M. F. (2013a).

The dynamically embedded plate anchor: results from an experimental and numerical

study. Proceedings of the 32nd International Conference in Ocean, Offshore and

Artic Engineering, Nantes, France, 9–14 June 2013. OMAE2013-11571.

O’Loughlin, C. D., Gaudin, C., Morton, J. P. and White D. J. (2014a). MEMS

accelerometer for measuring dynamic penetration events in geotechnical centrifuge

tests. Physical Modelling in Geotechnics, Vol. 50, No. 5, pp. 31–39.

O’Loughlin, C. D., Lowmass, A. C., Gaudin, C. and Randolph, M. F. (2006). Physical

modelling to assess keying characteristics of plate anchors. Proceedings of the 6th

International Conference on Physical Modelling in Geotechnics, Hong Kong, China,

4–6 August 2006, Vol. 1, pp. 659–666.

O’Loughlin, C. D., Richardson, M. D., and Randolph, M. F. (2009). Centrifuge tests on

dynamically installed anchors. Proceeding of the 28th International Conference on

Offshore Mechanics and Arctic Engineering, Honolulu, Hawaii, USA, 31 May–5

June 2009, Vol. 7, OMAE2009-80238.

O’Loughlin, C. D., Richardson, M. D., Randolph, M. F. and Gaudin, C. (2013b).

Penetration of dynamically installed anchors in clay. Géotechnique, Vol. 63, No. 11,

pp. 909–919.

Offshore Magazine (2005). Constitution sets foundation for new GoM developments.

Offshore Magazine, June 2005, Vol. 65, No. 6, p20.

Offshore Magazine (2008). Brazil reserves soar on new find. Offshore Magazine,

January 2008, Vol. 68, No. 1, p66.

Offshore Magazine (2010). Perdido advances deepwater GoM production possibilities.

Offshore Magazine, December 2010, Vol. 670, No. 12, p30.

Page 294: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

260

Offshore Magazine (2012). Shell opts for offshore FLNG for Australia’s stranded

Prelude field. Offshore Magazine, May 2012, Vol. 72, No. 5, p100.

Offshore Magazine (2013). Exxon proposes FLNG for deepwater offshore Western

Australia. Accessed 25th March 2014 from Offshore Magazine at

http://www.offshore-mag.com/articles/2013/04/exxon-proposes-flng-for-deepwater-

offshore-western-australia.html

O'Loughlin, C. D., Randolph, M. F. and Einav, I. (2004a). Physical modelling of deep

penetrating anchors. Proceedings of the 9th Australian and New Zealand Conference

on Geomechanics, Auckland, New Zealand, 8–11 February 2004, Vol. 2, pp. 710–

716.

O'Loughlin, C. D., Randolph, M. F. and Richardson. M. (2004b). Experimental and

theoretical studies of deep penetrating anchors. Proceedings of the Offshore

Technology Conference, Houston, Texas, USA, 3–6 May 2004, OTC 16841.

Ortman, C. M. 2008. The effect of diameter on dynamic seabed penetration. Trident

Scholar Project Report No. 373, USA Naval Academy, Annapolis, Maryland, USA.

Ove Arup and Partners. (1982). An initial study into penetrator shape and size for

emplacement in deep ocean sediments. Report DOE/RW/82.055, Department of the

Environment, London, United Kingdom.

Øye, I. (2000). Simulation of trajectories for a deep penetrating anchor. CFD Norway

Report 250:2000, CFD Norway AS, Trondheim, Norway.

Peuchen, J. and Mayne, P. (2007). Rate effects in vane shear testing. Proceedings of the

6th International Conference on Offshore Site Investigation and Geotechnics,

London, United Kingdom, 11–13 September 2007, pp. 187–194.

Page 295: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

261

PFC Energy (2011). White Paper: Importance of the Deepwater Gulf of Mexico. PFC

Energy, Washington DC, USA.

Pipeline and Gas Journal (2012). BW Pioneer: FPSO begins producing oil and gas in

Gulf. Pipeline and Gas Journal, April 2012, Vol. 239, No. 4.

Poorooshasb, F. and James, R. G. (1989). Centrifuge modelling of heat-generating

waste disposal. Canadian Geotechnical Journal, Vol. 26, No. 4, pp. 640–652.

Raie, M. S. and Tassoulas, J. L. (2009). Installation of torpedo anchors: numerical

modelling. Journal of Geotechnical and Geoenvironmental Engineering, Vol. 135,

No. 12, pp. 1805–1813.

Randolph M. F., Gaudin, C., Gourvenec, S. M., White, D. J., Boylan, N. and Cassidy,

M. J. (2011). Recent advances in offshore geotechnics for deep water oil and gas

developments. Ocean Engineering, Vol. 38, pp. 818–834.

Randolph, M. F. (2013). Analytical contributions to offshore geotechnical engineering.

Proceedings of the 18th International Conference on Soil Mechanics and

Geotechnical Engineering, Paris, France, 2–6 September 2013, pp. 85–105.

Randolph, M. F. and Hope, S. (2004). Effect of cone velocity on cone resistance and

excess pore pressures. Proceedings of the International Symposium on Engineering

Practice and Performance of Soft Deposits, Osaka, Japan, 2–4 June 2004, pp. 147–

152.

Randolph, M. F. and White, D. J. (2012). Interaction forces between pipelines and

submarine slides – a geotechnical viewpoint. Ocean Engineering, Vol. 48, pp. 32–37.

Randolph, M. F., Cassidy, M. J., Gourvenec, S. and Erbrich, C. T. (2005). Challenges of

offshore geotechnical engineering. Proceedings of the 16th International Conference

Page 296: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

262

on Soil Mechanics and Geotechnical Engineering, Osaka, Japan, 12–16 September

2005, Vol. 1, pp. 123–176.

Randolph, M. F., Jewell, R. J., Stone, K. J. L. and Brown, T. A. (1991). Establishing a

new centrifuge facility. Proceedings of the International Conference on Centrifuge

Modelling, Boulder, Colorado, USA, 13–14 June 1991, pp. 659–666.

Randolph, M. F., Low, H. E. and Zhou, H. (2007). Keynote lecture: In situ testing for

design of pipeline and anchoring systems. Proceedings of the 6th International

Conference on Offshore Site Investigation and Geotechnics, London, United

Kingdom, 11–13 September 2007, pp. 251–262.

Ranjan, G. and Arora, V. B. (1980). Model studies on anchors under horizontal pull in

clay. Proceedings of the 3rd Australian and New Zealand Conference on

Geomechanics, Wellington, New Zealand, 12–16 May 1980, Vol. 1, pp. 65–70.

Richardson, M. D, O’Loughlin, C. D., Randolph, M. F and Gaudin, C. (2009). Setup

following installation of dynamic anchors in normally consolidated kaolin clay.

Journal of Geotechnical and Geoenvironmental Engineering, Vol. 135, No. 4, pp.

487–496.

Richardson, M. D. (2008). Dynamically installed anchors for floating offshore

structures. PhD thesis, The University of Western Australia, Perth, Australia.

Richardson, M., O’Loughlin, C. D. and Randolph, M. F. (2005). The geotechnical

performance of deep penetrating anchors in calcareous sand. Proceedings of the

International Symposium on Frontiers in Geotechnics, Perth, Australia, 19–21

September 2005, pp. 357–363.

Richardson, M., O’Loughlin, C. D. and Randolph, M. F. (2006). Drum centrifuge

modelling of dynamically penetrating anchors. Proceedings of the 6th International

Page 297: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

263

Conference on Physical Modelling in Geotechnics, Hong Kong, China, 4–6 August

2006, Vol. 1, pp. 673–678.

Rowe, R. K. and Davis, E. H. (1982). The behaviour of anchor plates in clay.

Géotechnique, Vol. 32, No. 1, pp. 9–23.

Sabetamal, H. (2014). Finite element algorithms for dynamic anlaysis of geotechnical

problems. Forthcoming PhD thesis, The University of Newcastle, Australia.

Sahdi, F., Gaudin, C., White, D. J., Boylan, N. and Randolph, M. F. (2014). Centrifuge

modelling of active slide – pipeline loading in soft clay. Géotechnique, Vol. 64, No.

1, pp. 16–27.

Schlue, B. F., Moerz, T. and Kreiter, S. (2010). Influence of shear rate on undrained

vane shear strength of organic harbor mud. Journal of Geotechnical and

Geoenvironmental Engineering, Vol. 136, No. 10, pp. 1437–1447.

Schofield, A. N. (1980). Cambridge geotechnical centrifuge operations. Géotechnique,

Vol. 30, No. 3, pp. 227–268.

Sharma, A. (2007). CMOS systems and circuits for sub-degree per hour MEMS

gyroscopes. PhD thesis, Georgia Institute of Technology, Atlanta, Georgia, USA.

Sheahan, T. C., Ladd, C. C. and Germaine, J. T. (1996). Rate-dependent undrained

shear behaviour of saturated clay. Journal of Geotechnical Engineering, Vol. 122,

No. 2, pp. 99–108.

Shelton, J. T., Nie, C. and Shuler, D. (2011). Installation penetration of gravity installed

plate anchors – laboratory study results and field history data. Proceedings of the

Offshore Technology Conference, Houston, Texas, USA, 2–5 May 2011, OTC

22502.

Page 298: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

264

Shin, E. C., Dass, R. N., Das, B. M. and Cook, E. E. (1994). Mud suction force in the

uplift of plate anchors in clay. Proceedings of the 4th International Offshore and

Polar Engineering Conference, Osaka, Japan, 10–15 April 1994, Vol. 1, pp. 462–

466.

Skempton, A. W. (1951). The bearing capacity of clays. Proceedings of the building

research congress, London, United Kingdom, September 1951, Vol. 1, pp. 180–189.

Song, Z. and Hu, Y. (2005). Vertical pullout behaviour of plate anchors in uniform clay.

Proceedings of the 1st International Symposium on Frontiers in Offshore

Geotechnics, Perth, Australia, 19–21 September 2005, pp. 205–211.

Song, Z., Hu, Y. and Gaudin, C. (2007). The influence of disturbed zone on capacity of

suction embedded plate anchors. Proceedings of the 17th International Offshore and

Polar Engineering Conference, Lisbon, Portugal, 1–6 July 2007, Vol. 2, pp. 1340–

1346.

Song, Z., Hu, Y. and Muelle, E. (2008a). Bearing capacity and keying of anchor in

normally consolidated clay. Proceedings of the 18th International Offshore and

Polar Engineering Conference, Vancouver, Canada, 6–11 July 2008. Vol. 2, pp.

737–742.

Song, Z., Hu, Y. and Randolph, M. F. (2008b). Numerical simulation of vertical pullout

of plate anchors in clay. Journal of Geotechnical and Geoenvironmental

Engineering, Vol. 134, No. 6, pp. 866–875.

Song, Z., Hu, Y., and Randolph, M. F. (2005). Pullout Behaviour of Inclined Plate

Anchors in Uniform Clay. Proceedings of the 11th International Conference of the

International Computer Methods and Advances in Geomechanics, Turin, Italy, 19–

24 June 2005, pp. 715–722.

Page 299: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

265

Song, Z., Hu, Y., O’Loughlin, C. D. and Randolph, M. F. (2009). Loss in anchor

embedment during plate anchor keying in clay. Journal of Geotechnical and

Geoenvironmental Engineering. Vol. 135, No. 10, pp. 1475–1485.

Song, Z., Hu, Y., Wang, D. and O’Loughlin, C. D. (2006). Pullout capacity and

rotational behaviour of square anchors in kaolin clay and transparent soil.

Proceeding of the 6th International Conference on Physical Modelling in

Geotechnics, Hong Kong, China, 4–6 August 2006, Vol. 2, pp. 1325–1331.

Spooner, I. S., Williams, P. and Martin, K. (2004). Construction and use of an

inexpensive, lightweight free-fall penetrometer: applications to paleolimnological

research. Journal of Paleolimnology, Vol. 32, No. 3, pp. 305–310.

Stark, N. and Wever, T. F. (2009). Unraveling subtle details of expendable bottom

penetrometer (XBP) deceleration profiles. Geo-Marine Letters, Vol. 29, No. 1, pp.

39–45.

Stark, N., Hanff, H. and Kopf, A. (2009a). Nimrod: A tool for rapid geotechnical

characterization of surface sediments. Sea Technology, Vol. 50, No. 4, pp. 10–14.

Stark, N., Kopf, A., Hanff, H., Stegmann, S. and Wilkens, R. (2009b). Geotechnica

linvestigations of sandy seafloors using dynamic penetrometers. Proceeding of

MTS/IEEE Oceans 2009 Conference, Biloxi, Mississippi, USA, 26–29 October 2009.

Stegmann, S. (2007). Design of a free-fall penetrometer for geotechnical

characterisation of saturated sediments and its geological application. PhD thesis,

Bremen University, Bremen, Germany.

Stegmann, S., Morz, T. and Kopf, A. (2006). Initial results of a new Free Fall-Cone

Penetrometer (FF-CPT) for geotechnical in situ characterisation of soft marine

sediments. Norwegian Journal of Geology, Vol. 86, No. 3, pp. 199–208.

Page 300: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

266

Steiner, A., Kopf, A. J., L’Heureux, J. S., Kreieter, S., Stegmann, S., Haflidason, H. and

Moerz, T. (2014). In situ dynamic piezocone penetrometer tests in natural clayey

soils – a reappraisal of strain-rate corrections. Canadian Geotechnical Journal, Vol.

51, No. 3, pp. 272–288.

Steiner, A., L'Heureux, J., Kopf, A., Vanneste, M., Longwa, O., Lange, M. and Haflidi,

H. (2012). An in-situ free-fall piezocone penetrometer for characterizing soft and

sensitive clays at Finneidfjord (Northern Norway). In Y. Yamada et al. (eds.).

Submarine Mass Movements and Their Consequences, Springer, Doerdrecht,

Netherlands, pp. 99–109.

Stephan, S., Kaul N. and Villinger H. (2012). The lance insertion retardation

(LIRmeter): an instrument for in-situ determination of sea floor properties – technical

description and performance evaluation. Marine Geophysical Research, Vol. 33, No.

3, pp. 209–221.

Stewart, D. P. and Randolph, M. F. (1991). A new site investigation tool for the

centrifuge. Proceedings of the International Conference on Centrifuge Modelling,

Boulder, Colorado, USA, 13–14 June 1991, pp. 531–538.

Stewart, D.P. (1992). Lateral loading of piled bridge abutments due to embankment

construction. PhD thesis, The University of Western Australia, Perth, Australia.

Stoll, R. D. (2004). Measuring sea bed properties using static and dynamic

penetrometers. Proceedings of Civil Engineering in the Oceans VI, Baltimore,

Maryland, USA, 20–22 October 2004, pp. 386–395.

Stoll, R. D. and Akal, T. (1999). XBP-Tool for rapid assessment of seabed sediment

properties. Sea Technology, Vol. 40, No. 2, pp. 47–51.

Page 301: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

267

Stoll, R. D., Sun, Y. F. and Bitte, I. (2007). Seafloor properties from penetrometer tests.

Oceanic Engineering, Vol. 32, No. 1, pp. 57–63.

Stovall, S. H. (1997). Basic inertial navigation. Navigation and Data Link Section

Systems Integration Branch, Naval Air Warfare Center Weapons Division Report

NAWCWPNS TM 8128, China Lake, California, USA.

Stringer, M. E., Heron, C. M. and Madabhushi, S. P. G. (2010) Experience using

MEMS-based accelerometers in dynamic testing. Proceedings of the 7th

International Conference on Physical Modelling in Geotechnics, 28 June–1 July

2010, Zurich, Switzerland, Vol. 1, pp. 389–394.

Strum, H. and Andersen, L. (2010). Large deformation analysis of the installation of

dynamic anchor. Proceedings of the 7th European Conference on Numerical

Methods in Geotechnical Engineering, Trondheim, Norway, 3–4 June 2010, pp. 255–

260.

Strum, H., Lieng, J. T. and Saygili, G. (2011). Effect of soil variability on the

penetration depth of dynamically installed drop anchors. Proceedings of the Offshore

Technology Conference, Rio de Janeiro, Brazil, 4–6 October 2011, OTC 22396.

Tanaka, H. and Locat, J. (1999). A microstructural investigation of Osaka Bay Clay: the

impact of microfossils on its mechanical behaviour. Canadian Geotechnical Journal,

Vol. 36, No. 3, pp. 493–508.

Taylor, R. N. (1995). Centrifuges in modelling: Principles and scale effects. Blackie

Academic and Professional, Glasgow, United Kingdom.

Tho, K. K., Chen, Z., Leung, C. F. and Chow, Y. K. (2013). Pullout behaviour of plate

anchor in clay with linearly increasing strength. Canadian Geotechnical Journal,

Vol. 51, No. 1, pp. 92-102.

Page 302: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

268

Thompson, D., March, R. and Herrmann, H. (2002). Groundtruth results for dynamic

penetrometers in cohesive soils. Proceeding of MTS/IEEE Oceans 2002 Conference,

Boston, Massachusetts, USA, 16–18 September 1981, Biloxi, Mississippi, USA, Vol.

4, pp. 2117–2123.

Thorne, C. P., Wang, C. X. and Carter, J. P. (2004). Uplift capacity of rapidly loaded

strip anchors in uniform strength clay. Géotechnique, Vol. 54, No. 8, pp. 507–517.

Tian, Y., Gaudin, C., Cassidy, M. J. and Randolph, M.F. (2013). Considerations on the

design of keying flap of plate anchors. Journal of Geotechnical and

Geoenvironmental Engineering, Vol. 139, No. 7, pp. 1156–1164.

True, D. G. (1976). Undrained vertical penetration into ocean bottom soils. PhD thesis,

University of California, Berkeley, California, USA.

Vaid, Y. P. and Campanella, R. G. (1977). Time-dependent behaviour of undisturbed

clay. Journal of the Geotechnical Engineering Division, Vol. 103, No. 7, pp. 693–

709.

Vesic, A. S. (1971). Breakout resistance of objects embedded in ocean bottom. Journal

of the Soil Mechanics and Foundations Division, Vol. 97, No. 9, pp. 1183–1205.

Vryhof Anchors (2010). Anchor Manual 2010: The Guide to Anchoring, Vryhof

Anchors, Rotterdam, The Netherlands. Accessed 9 June 2014 at

http://www.vryhof.com

Wang, D. and O’Loughlin, C. D. (2014). Numerical study of pull-out capacity of

dynamically embedded plate anchors. Canadian Geotechnical Journal , Vol. 51, No.

12, pp. 1–10.

Page 303: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

269

Wang, D., Gaudin, C. and Randolph, M. F. (2013a). Large deformation finite element

analysis investigating the performance of anchor keying flap. Ocean Engineering,

Vol. 59, pp. 107–116.

Wang, D., Han, C. and Gaudin, C. (2013b). Effect of roughness of plate anchors.

Proceedings of the International Conference on Installation Effects in Geotechnical

Engineering, Rotterdam, The Netherlands, 24–27 March 2013, pp. 264–269.

Wang, D., Hu, Y. and Randolph M. F. (2008a). Three-dimensional large deformation

analyses of plate anchor keying in clay. Proceedings of the ASME 27th International

Conference on Offshore Mechanics and Arctic Engineering, 15–20 June 2008,

Estoril, Portugal, pp. 131–137.

Wang, D., Hu, Y. and Randolph, M. F. (2008b). Effect of loading rate on the uplift

capacity of plate anchors. Proceedings of the 18th International Offshore and Polar

Engineering Conference, Vancouver, Canada, 6–11 July 2008, Vol. 2, pp. 727–731.

Wang, D., Hu, Y. and Randolph, M. F. (2010). Three-dimensional large deformation

finite-element analysis of plate anchors in uniform clay. Journal of Geotechnical and

Geoenvironmental Engineering, Vol. 136, No. 2, pp. 355–365.

Wang, D., Hu, Y. and Randolph, M. F. (2011). Keying of rectangular plate anchors in

normally consolidated clays. Journal of Geotechnical and Geoenvironmental

Engineering, Vol. 137, No. 12, pp. 1244–1253.

Wilde, B. (2009). Torpedo pile anchors enter the GoM. Accessed 23 May 2014 from E

& P Magazine at http://www.epmag.com/Production/Torpedo-pile-anchors-enter-

GoM_45716

Page 304: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

270

Wilde, B., Treu, H. and Fulton, T. (2001). Field testing of suction embedded plate

anchors. Proceedings of the 11th International Offshore and Polar Engineering

Conference, Stavanger, Norway, 17–22 June 2001, Vol. 2, pp. 544–551.

Wong, P., Gaudin, C., Randolph M. F., Cassidy, M. J. and Tian, Y. (2012). Performance

of suction embedded plate anchors in permanent mooring applications. Proceedings

of the 22nd International Offshore and Polar Engineering Conference, Rhodes,

Greece, 17–23 June 2012, pp. 640–645.

Yafrate, N. J. and Dejong, J. T. (2007). Influence of penetration rate on measured

resistance with full flow penetrometers in soft clay. Proceedings of Sessions of Geo-

Denver 2007, Denver, Colorado, USA, 18–21 February 2007.

Yafrate, N., DeJong, J. T., DeGroot, D. and Randolph, M. F. (2009). Evaluation of

remoulded shear strength and sensitivity of soft clay using full flow pentrometers.

Journal of Geotechnical and Geoenvironmental Engineering, Vol. 122, No. 2, pp.

99–108.

Yang, M., Aubeny, C. and Murff, J. (2012). Behaviour of suction embedded plate

anchors during keying process. Journal of Geotechnical and Geoenvironmental

Engineering, Vol. 138, No. 2, pp. 174–183.

Young, A. G., Bernard, B. B., Remmes, B. D. Babb, L. V. and Brooks, J. M. (2011).

CPT Stinger - An innovative method to obtain CPT data for integrated geoscience

studies. Proceedings of the Offshore Technology Conference, Houston, Texas, USA,

2–5 May 2011, OTC 21569.

Yu, H. S. (2000). Cavity expansion methods in geomechanics. Kluwer, Dordrecht,

Netherlands.

Page 305: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

271

Yu, L., Liu, J., Kong, X. and Hu, Y. (2009). Three-dimensional numerical analysis of

the keying of the vertically installed plate anchors in clay. Computers and

Geotechnics, Vol. 36, No. 4, pp. 558–567.

Yu, L., Liu, J., Kong, X. and Hu, Y. (2011). Numerical study on plate anchor stability in

clay. Géotechnique, Vol. 61, No. 3, pp. 235–246.

Zhang, Y., Bienen, B., Cassidy, M. J. and Gourvenec, S. (2012). Undrained bearing

capacity of deeply buried flat circular footings under general loading. Journal of

Geotechnical and Geoenvironmental Engineering, Vol. 138, No. 3, pp. 385–397.

Zhou, H. and Randolph, M. F. (2009). Numerical investigations into cycling of full-

flow penetrometers in soft clay. Géotechnique, Vol. 59, No. 10, pp. 801–812.

Zhu, H. and Randolph, M. F. (2011). Numerical analysis of a cylinder moving through

rate-dependent undrained soil. Ocean Engineering, Vol. 38, pp. 943–953.

Zhu, J. G. and Yin, J. H. (2000). Strain-rate-dependent stress-strain behavior of

overconsolidated Hong Kong marine clay. Canadian Geotechnical Journal, Vol. 37,

No. 6, pp. 1272–1282.

Zimmerman, E. H. (2007). New gravity installed anchor requires less soil penetration.

Offshore, Vol. 67, No. 1, pp. 70–71.

Zimmerman, E. H. and Spikula, D. (2005). A new direction for subsea anchoring and

foundations. Accessed 19th August 2014 from Society of Naval and Marine

Engineering at http://www.sname.org/sections/texas/Meetings/Presentations.

Zimmerman, E. H., Smith, M. W. and Shelton, J. T. (2009). Efficient gravity installed

anchor for deepwater mooring. Proceedings of the Offshore Technology Conference,

Houston, Texas, USA, 2–5 May 2011, OTC 20117.

Page 306: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

272

APPENDIX I – SETUP FOLLOWING KEYING OF PLATE

ANCHORS ASSESSED THROUGH CENTRIFUGE TESTS

IN KAOLIN CLAY

Page 307: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

273

Page 308: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

274

Page 309: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

275

Page 310: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

276

Page 311: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

277

Page 312: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

278

Page 313: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

279

APPENDIX II – FIELD TESTING OF A REDUCED SCALE

DYNAMICALLY EMBEDDED PLATE ANCHOR

Page 314: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

280

Page 315: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

281

Page 316: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

282

Page 317: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

283

Page 318: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

284

Page 319: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

285

Page 320: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

286

Page 321: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

287

Page 322: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

288

APPENDIX III – THE DYNAMICALLY EMBEDDED

PLATE ANCHOR: RESULTS FROM AN

EXPERIMENTAL AND NUMERICAL STUDY

Page 323: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

289

Page 324: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

290

Page 325: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

291

Page 326: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

292

Page 327: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

293

Page 328: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

294

Page 329: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

295

Page 330: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

296

Page 331: AN EXPERIMENTAL STUDY TO ASSESS THE GEOTECHNICAL ...€¦ · Chapter 6: Blake, A. P. and O’Loughlin, C. D. (2015). Installation of dynamically embedded plate anchors as assessed

297