by john philip morton b.a. b.a.i. (hons) this thesis …...- i - the dynamic measurement of...

295
- I - THE DYNAMIC MEASUREMENT OF UNDRAINED SHEAR STRENGTH USING AN INSTRUMENTED FREE FALLING SPHERE By John Philip Morton B.A. B.A.I. (Hons) This thesis is submitted for the Degree of Doctor of Philosophy School of Civil, Environmental and Mining Engineering 2015

Upload: others

Post on 19-Feb-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

- I -

THE DYNAMIC MEASUREMENT OF UNDRAINED SHEAR

STRENGTH USING AN INSTRUMENTED FREE FALLING

SPHERE

By

John Philip Morton

B.A. B.A.I. (Hons)

This thesis is submitted for the

Degree of Doctor of Philosophy

School of Civil, Environmental and Mining Engineering

2015

- II -

- III -

ABSTRACT

This thesis introduces a new in situ site investigation tool aimed at providing an

accurate and rapid assessment of the undrained shear strength of near-surface offshore

soil. The research describes a new spherical free-falling penetrometer (FFP) which

represents a new design approach for FFPs which are almost invariably slender, full-

shafted projectiles. The custom-made free-fall sphere is 250 mm in diameter and

consists of two hemispheres that are bolted together with an internal vertically

orientated cylindrical void to accommodate instrumentation and a motion logger. The

free-fall sphere is designed to be released from a hanging position above the seabed and

penetrate the seabed by the kinetic energy obtained through free-fall in water.

This thesis carried out an experimental study involving field tests and centrifuge

experiments. The field tests comprise dynamic embedment data for 87 different tests

undertaken in two soft soil sites: (i) an inland lake, Lower Lough Erne in the Northwest

of Ireland and (ii) an offshore site in the Firth of Clyde which is located off the coast of

Scotland in the Irish Sea. The centrifuge experiments comprise dynamic tests in three

different soils: (i) Laminaria soil, recovered from the Timor Sea, (ii) West Africa clay,

recovered from the Gulf of Angola and (iii) kaolin clay. In both sets of experiments, the

sphere contained instrumentation that accurately measured the motion history in soil.

These data led to the development of a newly proposed theoretical framework for

assessing the dynamic resistance forces acting on the free-fall sphere. The framework

was cast in terms of both fluid mechanics drag resistance and geotechnical shear

resistance, but formulated in terms of a single capacity factor. In each soil a power law

- IV -

function was adopted in order to account for the strain rate dependency. The

appropriateness of the strain rate parameter was demonstrated by varying β within the

typical range reported from variable rate penetrometer tests (β = 0.05 to 0.09). In the

field and centrifuge experiments, the best-fit rate parameter was calculated using β =

0.07.

To improve the strength characterisation of near surface seabeds with a shallowly-

embedded ball penetrometer or free-fall sphere, the thesis describes centrifuge

experiments designed to quantify the shallow penetration effects. The experiments were

carried out with an 11.3 mm diameter ball penetrometer penetrating kaolin clay under

undrained conditions over a range of normalised strength ratios, su/γ'D. The tests

captured the influence of two important mechanisms: (i) the varying soil buoyancy with

penetration depth and (ii) the reduced bearing factor, Nb-shallow, arising from the shallow

failure mechanism. The centrifuge results were combined with reinterpreted data from

LDFE analyses to form a unique relationship between the transition depth and the

normalised strength ratio over the range su/γ'D ≈ 0.07 to 40. This led to the development

of a shallow penetration framework to determine more accurately the undrained shear

strength of near surface soil.

Both frameworks described in the thesis represent significant advances in the

understanding of FFPs and when combined provide accurate estimation of the

undrained shear strength with the IFFS compared to a conventional ball penetrometer.

- V -

DECLARATION

This thesis is submitted as a series of papers within the regulations of the University of

Western Australia. The thesis contains published material and draft submissions that

have been prepared for publication. Unless otherwise stated, the candidate is responsible

for over 90% of the content in this thesis. The contribution of the candidate for the

papers described in Chapter 3–8 are defined as follows.

Chapter 3: Morton, J. P. & O‟Loughlin, C. D., 2012. Dynamic penetration of a

sphere in clay. Proceedings of the 7th International Conference on Offshore

Site Investigation and Geotechnics, London, UK, pp. 223–230.

The field tests and analysis of dynamic embedment in clay were performed

solely by the candidate. Both authors contributed to the publication of this

chapter after a full initial draft was provided by the candidate.

Chapter 4: Morton, J. P., O‟Loughlin, C. D. & White. D. J., 2014. Strength

assessment during shallow penetration of a sphere in clay. Géotechnique Letters

4 (October-December), pp. 262–266.

The candidate designed the centrifuge tests and carried out all physical

modelling experiments and interpreted all results. The paper was written by the

candidate after comments from the co-authors.

Chapter 5: O‟Loughlin, C. D., Gaudin, C., Morton. J. P. & White, D. J., 2014.

MEMS accelerometers for measuring dynamic penetration events in

geotechnical centrifuge tests International Journal of Physical Modelling in

Geotechnics, 14(2), pp. 31–39.

- VI -

The candidate led the data processing and interpretation of the experimental

data and produced a technical report that formed a basis for drafting the paper.

The estimated precent contribution of the author is 30%.

Chapter 6: Blake, A. P., O'Loughlin, C. D., Morton, J. P., O' Beirne, C., Gaudin,

C. & White, D. W., 2015. In-situ measurement of the dynamic penetration of

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

Geotechnical Testing Journal, ASTM. DOI: 10.1520/GTJ20140135.

The field testing was performed in-part by the candidate along with some data

processing. The candidate was also involved in the data interpretation process;

the estimated precent contribution of the author is 20%.

Chapter 7: Morton, J. P., O‟Loughlin, C. D. & White. D. J., 2015. Field testing

an in situ freefalling spherical penetrometer in soft soil. Submitted to

Géotechnique.

The candidate carried out all field tests and analytical modelling in this paper.

The candidate also provided a detailed first draft that was revised after

comments from both co-authors.

Chapter 8: Morton, J. P., O‟Loughlin, C. D. & White. D. J., 2015. Centrifuge

modelling of an instrumented free-fall sphere for measurement of undrained

strength in fine grained soils. Canadian Geotechnical Journal. DOI:

10.1139/cgj-2015-0371

The candidate carried out all field tests and data interpretation in this paper and

produced the final submission draft after comments from the co-authors.

- VII -

The stated contributions have been agreed with the co-authors of each paper and

permission has been granted to include the relevant paper within this thesis.

John Morton, candidate………………………...…………………………………………

Conleth D O‟Loughlin, coordinating supervisor…………………….……………………

- VIII -

Contents

ABSTRACT .................................................................................................................. III

DECLARATION ............................................................................................................ V

LIST OF FIGURES .................................................................................................. XIII

LIST OF TABLES ..................................................................................................... XIX

ACKNOWLEDGEMENTS........................................................................................ XX

NOTATION ..................................................................................................................... 1

CHAPTER 1. INTRODUCTION .................................................................................. 7

1.1. RESEARCH MOTIVATIONS ......................................................................................... 7

1.2. THESIS OBJECTIVES ................................................................................................ 12

1.3. RESEARCH METHODOLOGY .................................................................................... 12

1.4. THESIS ORGANISATION ........................................................................................... 14

CHAPTER 2. REVIEW OF LITERATURE ............................................................. 19

2.1. INTRODUCTION ...................................................................................................... 19

2.2. IN SITU TOOLS ........................................................................................................ 19

2.2.1. In situ vane ..................................................................................................... 20

2.2.2. Cone and piezocone........................................................................................ 20

2.2.3. Full flow penetrometers.................................................................................. 21

2.2.3.1. T-bar and ball penetrometer ............................................................................... 23

2.2.4. Interpretation of shear strength from full flow penetrometers ....................... 24

2.2.5. Bearing capacity factor .................................................................................. 25

2.2.6. Interpretation of shear strength during shallow penetration ......................... 27

2.2.6.1. Transition depth .................................................................................................. 28

2.2.6.2. Wall failure ......................................................................................................... 28

2.2.6.3. Flow failure ........................................................................................................ 29

2.2.6.4. Bearing capacity factor variation with depth ..................................................... 32

2.2.6.5. Soil buoyancy ...................................................................................................... 32

- IX -

2.2.6.6. Operative depth .................................................................................................. 33

2.3. FREE-FALLING PENETROMETERS ............................................................................ 34

2.3.1. Free-falling penetrometer designs for naval mine countermeasure

applications .............................................................................................................. 34

2.3.2. Free-falling penetrometer designs for seabed characterisation .................... 37

2.3.3. FFP measurement systems ............................................................................. 39

2.3.3.1. Acoustic Doppler system .................................................................................... 39

2.3.3.2. Accelerometer system ......................................................................................... 39

2.3.4. Experimental and field studies on free-falling penetrometers ....................... 41

2.4. OCEANIC WASTE CARRIERS .................................................................................... 45

2.5. DYNAMICALLY INSTALLED ANCHORS .................................................................... 46

2.5.1. Experimental and field studies on dynamically installed anchors ................. 48

2.5.2. Centrifuge experiments ................................................................................... 50

2.6. RESISTANCE FORCES ACTING ON A PROJECTILE DURING DYNAMIC EMBEDMENT IN

SOIL .............................................................................................................................. 51

2.6.1. Fluid drag ....................................................................................................... 52

2.6.1.1. Drag in water ..................................................................................................... 53

2.6.1.2. Drag in soil ........................................................................................................ 56

2.6.2. Hydrodynamic mass force .............................................................................. 57

2.6.3. Strain rate effects in clay ................................................................................ 58

2.6.3.1. Strain rate parameter ......................................................................................... 60

2.6.3.2. Dependency of rate parameters on penetrometer geometry .............................. 62

2.6.3.3. Dependency of rate parameters on material properties .................................... 63

2.6.4. Combined fluid mechanics and soil mechanics framework ........................... 64

2.6.5. Unified framework .......................................................................................... 65

2.6.6. Analytical modelling ....................................................................................... 67

2.6.7. Numerical analysis ......................................................................................... 69

2.7. SUMMARY .............................................................................................................. 70

CHAPTER 3. DYNAMIC PENETRATION OF A SPHERE IN CLAY ................. 72

3.1. ABSTRACT ............................................................................................................. 72

3.2. INTRODUCTION ...................................................................................................... 73

3.3. SITE DESCRIPTION AND SOIL PROPERTIES ............................................................... 74

3.3.1. Site location and description .......................................................................... 74

3.3.2. Soil classification ........................................................................................... 74

- X -

3.3.3. Shear strength profiles ................................................................................... 75

3.4. TEST EQUIPMENT AND TESTING PROCEDURES ........................................................ 76

3.4.1. Instrumented free-fall sphere ......................................................................... 76

3.4.2. Motion logger ................................................................................................. 77

3.4.3. Field testing procedure .................................................................................. 78

3.5. TEST RESULTS AND ANALYSIS ................................................................................ 78

3.5.1. Acceleration profile ........................................................................................ 78

3.5.2. Acceleration profile interpretation ................................................................. 81

3.5.3. Velocity and embedment depth profile ........................................................... 82

3.6. EMBEDMENT DEPTH PREDICTION ........................................................................... 82

3.7. CONCLUSIONS ........................................................................................................ 89

CHAPTER 4. STRENGTH ASSESSMENT DURING SHALLOW

PENETRATION OF A SPHERE IN CLAY .............................................................. 90

4.1. ABSTRACT ............................................................................................................. 90

4.2. INTRODUCTION ...................................................................................................... 91

4.3. EXPERIMENTAL DETAILS ........................................................................................ 92

4.4. EXPERIMENTAL PROCEDURE .................................................................................. 93

4.4.1. Preparation of clay specimen ......................................................................... 93

4.5. THEORETICAL BASIS FOR INTERPRETATION OF MEASURED BALL PENETRATION

RESISTANCE .................................................................................................................. 96

4.6. RESULTS AND COMPARISONS ............................................................................... 100

4.6.1. In-flight video camera observations ............................................................. 100

4.6.2. Undrained shear strength profiles ............................................................... 101

4.6.3. Deep mechanism transition depth ................................................................ 102

4.6.4. Shallow bearing capacity factors ................................................................. 103

4.7. CONCLUSIONS ...................................................................................................... 104

CHAPTER 5. MEMS ACCELEROMETERS FOR MEASURING DYNAMIC

PENETRATION EVENTS IN GEOTECHNICAL CENTRIFUGE TESTS ........ 106

5.1. ABSTRACT ........................................................................................................... 106

5.2. INTRODUCTION .................................................................................................... 107

5.3. STATIC CENTRIFUGE TESTS .................................................................................. 111

5.4. EXAMPLE APPLICATION: DYNAMICALLY INSTALLED ANCHORS ............................ 115

5.5. DYNAMIC CENTRIFUGE TESTS .............................................................................. 117

- XI -

5.6. INTERPRETATION OF ACCELEROMETER DATA ....................................................... 122

5.7. CONCLUDING REMARKS ....................................................................................... 130

CHAPTER 6. IN-SITU MEASUREMENT OF THE DYNAMIC PENETRATION

OF FREE FALL PROJECTLES IN SOFT SOILS USING A LOW COST

INERTIAL MEASUREMENT UNIT ....................................................................... 131

6.1. ABSTRACT ........................................................................................................... 131

6.2. INTRODUCTION .................................................................................................... 132

6.3. FREE-FALLING PROJECTILES ................................................................................. 136

6.3.1. Deep penetrating anchors ............................................................................ 136

6.3.2. Dynamically embedded plate anchors ......................................................... 137

6.3.3. Instrumented free-falling sphere .................................................................. 138

6.3.4. Inertial measurement unit ............................................................................ 139

6.4. INTERPRETATION OF IMU MEASUREMENTS ......................................................... 141

6.4.1. Rotation ........................................................................................................ 142

6.4.2. Acceleration .................................................................................................. 144

6.4.3. Velocity and distance .................................................................................... 146

6.4.4. Tilt angles ..................................................................................................... 147

6.5. TEST SITES AND SOIL PROPERTIES ........................................................................ 147

6.6. TEST PROCEDURE ................................................................................................. 150

6.7. RESULTS AND DISCUSSION ................................................................................... 153

6.7.1. Rotation ........................................................................................................ 153

6.7.2. Acceleration .................................................................................................. 157

6.7.3. Velocity profiles ............................................................................................ 160

6.7.4. Verification of the IMU derived measurements ........................................... 163

6.7.5. Example application of projectile IMU data ................................................ 164

6.8. CONCLUSIONS ...................................................................................................... 168

CHAPTER 7. ESTIMATION OF SOIL STRENGTH BY INSTRUMENTED

FREE-FALL SPHERE TESTS .................................................................................. 170

7.1. ABSTRACT ........................................................................................................... 170

7.2. INTRODUCTION .................................................................................................... 171

7.3. BEARING CAPACITY FACTOR ................................................................................ 174

7.4. SITE DESCRIPTION AND SOIL PROPERTIES ............................................................. 175

7.5. TEST EQUIPMENT AND TESTING PROCEDURES ...................................................... 178

- XII -

7.5.1. Instrumented free-fall sphere ....................................................................... 178

7.5.2. Inertial measurement unit ............................................................................ 178

7.5.3. Field testing procedure ................................................................................ 180

7.6. FORCES ACTING ON A SPHERE DURING FREE-FALL IN WATER ............................... 184

7.7. FORCES ACTING ON A SPHERE DURING DYNAMIC PENETRATION IN SOIL ............... 188

7.8. SOIL STRENGTH ESTIMATION USING FREE-FALL SPHERE DATA ............................. 193

7.9. CONCLUSIONS ...................................................................................................... 196

CHAPTER 8. CENTRIFUGE MODELLING OF AN INSTRUMENTED FREE-

FALL SPHERE FOR MEASUREMENT OF UNDRAINED STRENGTH IN

FINE-GRAINED SOILS ............................................................................................ 197

8.1. ABSTRACT ........................................................................................................... 197

8.2. INTRODUCTION .................................................................................................... 198

8.3. PENETRATION OF A SPHERE IN SOIL ...................................................................... 202

8.4. EXPERIMENTAL DETAILS ...................................................................................... 206

8.4.1. Measurement technique ................................................................................ 206

8.4.2. Instrumented free-fall sphere ....................................................................... 209

8.4.3. Soil preparation technique ........................................................................... 210

8.4.4. Centrifuge test details and procedures ......................................................... 210

8.5. TEST RESULTS AND DISCUSSION ........................................................................... 212

8.5.1. Penetrometer tests and soil properties ......................................................... 212

8.5.2. Free-fall tests ................................................................................................ 216

8.5.3. Interpretation of free-fall acceleration data ................................................ 222

8.6. CONCLUSION ........................................................................................................ 230

CHAPTER 9. CONCLUSIONS ................................................................................. 232

9.1. SUMMARY ............................................................................................................ 232

9.2. MAIN FINDINGS .................................................................................................... 234

9.2.1. Free-fall and dynamic embedment in soil .................................................... 234

9.2.2. MEMS accelerometers in the centrifuge ...................................................... 235

9.2.3. Shallow penetration framework ................................................................... 236

REFERENCES ............................................................................................................ 239

- XIII -

LIST OF FIGURES

Figure 1.1. Ball Penetrometer, T-bar and Cone Penetrometer .......................................... 8

Figure 1.2. Examples of full-shafted penetrometers with conical tips: (a) CPT-Lance

(after Stark et al. 2009b); (b) Nimrod (after Stark et al. 2009b); (c) FFCPT (Furlong

et al. 2006); (d) CPT Stinger (after Young et al. 2011) .............................................. 9

Figure 1.3 Schematic of the Instrumented Free-Fall Sphere........................................... 11

Figure 1.4. Sphere separated to show internally housed data logger housed .................. 11

Figure 2.1. (a) T-bar; (b) Ball penetrometer ................................................................... 23

Figure 2.2. Soil failure mechanism for a ball penetration in uniform clay (Zhou et al.

2013) ......................................................................................................................... 30

Figure 2.3. The effect of su/γ‟D on the transition depth for a ball penetrometer (after

Zhou et al. 2013) ...................................................................................................... 31

Figure 2.4. 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) ..... 36

Figure 2.5. FFP systems developed for applications in seabed characterisation (a)

Marine Impact Penetrometer (after Dayal et al. 1975), (b) MSP-2 (after Colp et al.

1975), (c) Freefall penetrometer (after Denness et al. 1981), (d) XBP (after Stoll &

Akal 1999), (e) FFCPT (www.brooke-ocean.com), (f) LIRmeter (after Stephan et

al. 2012), (g) CPT-Lance (after Stark et al. 2009b), (h) Nimrod (after Stark et al.

2009b), (i) CPT Stinger (after Young et al. 2011) ................................................... 38

Figure 2.6. Typical velocity and penetration depth with time profiles for the FFP

instrumented with an accelerometer (Chow & Airey 2010b) .................................. 41

Figure 2.7. Acceleration profile in soil (after Stephan et al. 2012) ................................. 41

Figure 2.8. Cone tip resistance of the PROBOS compared to the STATPEN, a quasi-

static cone penetrometer (after Stoll et al. 2007) ...................................................... 43

Figure 2.9. (a) Different tip designs for the STING; (b) Tip diameter effect in

interpreted soil strength for the STING (after Abelev et al. 2009b) ........................ 44

- XIV -

Figure 2.10. The European Standard Penetrometer (Freeman & Burdett, 1986) ........... 46

Figure 2.11. Dynamically installed anchors (Medeiros, 2002); (b) Torpedo Anchor

(Brandão et al. 2006) ................................................................................................ 48

Figure 2.12. (a) Forces acting on a full-shafted penetrometer and (b) Thin-shafted

penetrometer during installation ............................................................................... 52

Figure 2.13. Drag coefficient for uniform flow past a sphere R = Re < 2 x 105 (480 data

points) after (Brown & Lawler 2003) ....................................................................... 54

Figure 2.14. Drag coefficient for uniform flow past a sphere Re > 2 x 105 (Achenbach,

1972) ......................................................................................................................... 55

Figure 2.15. Laminar-separated flow and turbulent flow over a sphere - (after

Finnemore & Franzini 2001) .................................................................................... 56

Figure 2.16. Normalised velocity illustration (after Lehane et al. 2009) ........................ 59

Figure 2.17. Variation in back-analysed β and λ strain rate parameters with impact

velocity for reduced scale model DPAs from centrifuge experiments in kaolin clay

(after O‟Loughlin et al. 2013b) ................................................................................ 61

Figure 2.18. Variation of normalised lateral pressure on a pipe with non-Newtonian

Reynolds number ...................................................................................................... 66

Figure 3.1. Site location and bathymetric map of Lower Lough Erne (after Lafferty et al.

2006) ......................................................................................................................... 74

Figure 3.2. Typical undrained shear strength profiles at the test site .............................. 76

Figure 3.3. (a) sphere separated to show internally housed data logger housed, (b)

motion logger and underwater housing, (c) sphere suspended over the water prior to

a drop ........................................................................................................................ 77

Figure 3.4. (a) x, y and z axis acceleration traces from a typical test, (b) x and y axis

rotation traces from the same test ............................................................................. 80

Figure 3.5. z axis acceleration trace for the test shown on Figure 3.4 together with

corresponding velocity and displacement traces ...................................................... 81

Figure 3.6. Velocity profiles in water and soil for release heights of 0.5 m and 1 m ..... 83

Figure 3.7. Measuring the embedment depth using markers on the retrieval rope taken

with the underwater camera ..................................................................................... 84

Figure 3.8. Forces acting on the IFFS during penetration in soil.................................... 84

Figure 3.9. Measured and theoretical velocity profiles of the sphere free-falling in water

.................................................................................................................................. 87

- XV -

Figure 3.10. Predicted and measured velocity profiles of the sphere penetrating the

lakebed ...................................................................................................................... 89

Figure 4.1. (a) and (b) Experimental arrangement in the beam centrifuge ..................... 94

Figure 4.2. A scraped soil sample before a test .............................................................. 95

Figure 4.3. Ball penetrometer and cavity after a penetration test ................................... 95

Figure 4.4. Schematic illustration of soil buoyancy due to (a) the sphere and (b) the

sphere and conical cavity (c) buoyancy function for a typical cavity depth ............ 98

Figure 4.5. Comparison of strength profiles from Equation 4.4 and qnet/Nb-deep ........... 100

Figure 4.6. Effect of strength ratio su/γ'D on transition depth ....................................... 103

Figure 4.7. Measured variation in normalised bearing factor with normalised

embedment depth and equation fit ......................................................................... 104

Figure 5.1. Schematic representation of the operational principle of a MEMS

accelerometer .......................................................................................................... 109

Figure 5.2. MEMS accelerometer: (a) Analog Devices ADXL001 MEMS accelerometer

with an internal view of the chip showing the MEMS sensor surrounded by the

integrated circuitry and (b) Scanning Electron Microscope image of the ADXL001

accelerometer showing the proof mass, plates, springs and anchor points ............ 110

Figure 5.3. Comparison between MEMS acceleration measurements and those derived

from the rotational speed of the centrifuge during spin up to 100 g and down again:

(a) time history, and (b) comparison ...................................................................... 113

Figure 5.4. Comparison between position measurements derived from the MEMS

accelerometer and measured using the motor encoder on the vertical axis of the

actuator ................................................................................................................... 115

Figure 5.5. Dynamically installed anchors: (a) torpedo pile (after Araujo et al. 2004), (b)

OMNI-Max anchor (after Shelton, 2007), (c) deep penetrating anchor (Deep Sea

Anchors, www.deepseaanchors.com/News.html) .................................................. 116

Figure 5.6. Accelerometers installed in a model dynamically installed anchor ............ 117

Figure 5.7. Dynamic anchor experimental arrangement showing the anchor located in

installation guide before release and embedded in the soil sample after release ... 119

Figure 5.8. MEMS and piezoelectric accelerometer data measured before, during and

after a dynamically installed anchor drop in a beam centrifuge: (a) entire trace, (b)

during freefall and embedment ............................................................................... 120

- XVI -

Figure 5.9. Interpretation of the MEMS accelerometer data in the rotating frame of

reference: (a) acceleration, velocity and displacement traces, (b) velocity profile

during freefall and embedment............................................................................... 124

Figure 5.10. Anchor velocity profiles for an anchor installation in kaolin clay with su =

1.1z ......................................................................................................................... 126

Figure 5.11. Effect of anchor tilt during embedment in soil (kaolin clay with su = 1.1z):

(a) 5 degree tilt, (b) 10 degree tilt, (c) 20 degree tilt, (d) 30 degree tilt ................. 129

Figure 6.1. Schematic representation of the operational principle of: (a) MEMS

accelerometers and (b) MEMS gyroscopes ............................................................ 134

Figure 6.2. Deep penetrating anchor ............................................................................. 137

Figure 6.3. Dynamically embedded plate anchor ......................................................... 138

Figure 6.4. Instrumented free-falling sphere ................................................................. 139

Figure 6.5. Inertial Measurement Unit .......................................................................... 140

Figure 6.6. Body frame of reference ............................................................................. 141

Figure 6.7. Resultant tilt angle, μ, defined in the inertial frame ................................... 142

Figure 6.8. Test sites locations ...................................................................................... 149

Figure 6.9. Undrained shear strength profiles: (a) Firth of Clyde and (b) Lough Erne 150

Figure 6.10. (a) RV Aora and (b) Self-propelled barge ................................................ 151

Figure 6.11. DEPLA field test procedure ..................................................................... 152

Figure 6.12. Image capture from ROV camera showing the follower retrieval line at the

seabed ..................................................................................................................... 153

Figure 6.13. Projectile rotations during free-fall through water and soil penetration: (a)

DEPLA, (b) IFFS and (c) DPA .............................................................................. 156

Figure 6.14. Projectile accelerations during free-fall through water and soil penetration:

(a) DEPLA, (b) IFFS and (c) DPA ......................................................................... 159

Figure 6.15. Projectile velocity profiles corresponding to free-fall through water and soil

penetration: (a) DEPLA, (b) IFFS and (c) DPA ..................................................... 162

Figure 6.16. Comparision of IMU derived displacement measurements with those

obtained using a draw wire sensor ......................................................................... 164

Figure 6.17. DEPLA velocity profile derived from the IMU data measured at the Firth

of Clyde test site and corresponding theoretical profile ......................................... 168

Figure 7.1. Undrained shear strength profiles in: (a) Lough Erne and (b) Firth of Clyde

................................................................................................................................ 177

- XVII -

Figure 7.2. Instrumented free-fall sphere: (a) sphere separated with IMU located within

internal void, (b) IMU and (c) assembled sphere prior to a free-fall test in Erne .. 178

Figure 7.3 IMU measurements and their interpretation from a typical free-fall sphere

test in Erne .............................................................................................................. 180

Figure 7.4. Forces acting on a sphere: (a) free-falling in water, (b) during dynamic

penetration in soil ................................................................................................... 186

Figure 7.5. Measured and theoretical evolution of the drag coefficient, CD, during free-

fall in water ............................................................................................................. 188

Figure 7.6. Relationship between N and Renon-Newtonian for: (a) Erne and (b) Clyde ..... 192

Figure 7.7. Comparison of undrained shear strength profiles derived from free-fall

sphere acceleration data and push-in piezoball penetration resistance: (a) Erne and

(b) Clyde ................................................................................................................. 195

Figure 8.1. Examples of free-fall shafted penetrometers with conical tips:(a) CPT-Lance

(courtesy of Dr. Nina Stark); (b) Nimrod (courtesy of Dr. Nina Stark); (c) FFCPT

(Furlong et al. 2006); (d) CPT Stinger (after Young et al. 2011) ........................... 200

Figure 8.2. Free-fall sphere ready for release (Morton and O‟Loughlin 2012) ............ 200

Figure 8.3. Forces acting on the sphere during penetration in soil ............................... 202

Figure 8.4. Model instrumented sphere shown: (a) during fabrication showing the void

in the sphere for the tri-axis MEMS accelerometer (b) after fabrication alongside a

centrifuge scale push-in ball penetrometer ............................................................. 208

Figure 8.5. Experimental arrangement for the push-in ball and instrumented free-fall

sphere tests ............................................................................................................. 211

Figure 8.6. Profiles of: (a) undrained shear strength with depth (from ball penetrometer

tests) and (b) moisture content and effective unit weight with depth established

from post-testing sample cores ............................................................................... 215

Figure 8.7. Accelerometer, velocity and displacement traces in a typical free-fall sphere

centrifuge test in kaolin clay................................................................................... 218

Figure 8.8. Example time histories of: (a)linear acceleration, (b) velocity and (c)

displacement for the three soil types ...................................................................... 220

Figure 8.9. Post-test analysis of the sphere trajectory and measurement of the final

embedment depth ................................................................................................... 222

Figure 8.10. Relationship between N and Renon-Newtonian for a sphere in the three soil

types: (a) Laminaria soil (b) West Africa clay (c) kaolin ....................................... 224

- XVIII -

Figure 8.11. su profiles from free-fall sphere and push-in penetrometer tests in: (a)

Laminaria soil (b) West Africa soil (c) kaolin ....................................................... 229

Figure 8.12. Effect of varying β parameter on free-fall sphere su profile ..................... 230

Figure 9.1. The formulation of the centroidal height for an invert depth < 0.5D ......... 274

Figure 9.2. The formulation of the centroidal height for an invert depth > 0.5D ......... 274

- XIX -

LIST OF TABLES

Table 7.1. Free-fall sphere test data from the Erne tests ............................................... 182

Table 7.2. Free-fall sphere test data from the Clyde tests ............................................. 184

- XX -

ACKNOWLEDGEMENTS

I would like to take this opportunity to sincerely thank my supervisors Dr. Conleth

O‟Loughlin and Professor David White. Thank you for your guidance, patience and

energy throughout this PhD, it has been invaluable.

To my colleagues in the Centre for Offshore Foundation Systems (COFS) UWA, thank

for your assistance and technical expertise and insights with regard to the experimental

work presented in this thesis. A special thanks to the soils and centrifuge laboratory and

workshop staff, in particular John Breen, Manuel Palacios, Kelvin Leong, Khin Seint,

Dave Jones and Shane De Catania. Your technical support and team attitude has made

the testing component one of the most enjoyable aspects of my research.

The financial support provided an Australian Postgraduate Award, Scholarship for

International Research Fees and COFS Top Up Scholarship is gratefully acknowledged.

Thank you to my colleagues and friends Janice Brogan, Cathal Colreavy, Anthony

Blake, Colm O‟Beirne, Michael Cocjin, Chao Han, Emma Leitner, Raffaele Ragni,

Henning Mohr, Simon Leckie, Fillippo Gaone, Lisa Melvin, Shiaohuey Chow, Cristina

Vulpe and Joe Tom JR. I am thankful for your consideration and the friendship we have

developed over the past four years. Thank you for the kindness, honesty and sense of

humour that you have all provided.

Finally, and most importantly, I would like to thank my family - Joan and Gary, Scott,

Gary JR, Claire and Grandparents John Flood and Iris Morton. Thank you for you open

minded attitude towards my research and thank you for you relentless love and warmth.

Without your continual support and encouragement I would not be the positive person I

am today.

- 1 -

NOTATION

Roman

a resultant linear acceleration

abz linear acceleration coincident with the body frame z-axis

aix linear acceleration coincident with the inertial frame x-axis

aiy linear acceleration coincident with the inertial frame y-axis

aiy linear acceleration coincident with the inertial frame x-axis

A acceleration measurement

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

ADC analogue to digital converter

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

Ap frontal/projected area

Ar resultant linear acceleration

As surface area

ch horizontal coefficient of consolidation

cv vertical coefficient of consolidation

Cm added mass coefficient

CD drag coefficient

D diameter

dt change in time

FAM added mass force

Fbuoy soil buoyancy force

Fb bearing resistance force

Fd fluid drag resistance force

Ffrict frictional resistance force

Fresist combined resistance force

- 2 -

FSS submerged weight in soil

g gravitational acceleration

hiz,drop drop height derived from the inertial frame acceleration and rotation measurements

Hc height of the conical cavity

Hw cavity depth

k undrained shear strength gradient

m mass

m' added mass

N bearing capacity factor

Nball ball penetrometer/ piezoball bearing capacity factor

Nb-shallow reduced bearing factor

Nc bearing capacity factor

Nc,deep maximum bearing capacity factor

Nkt piezocone bearing capacity factor

NT-bar T-bar bearing capacity factor

n, nl strain rate parameters

qm measured penetration resistance

Rbi direction cosine matrix (body frame to inertial frame)

Re Reynolds number for a Newtonian fluid

Renon-Newtonian Reynolds number for a non-Newtonian fluid

Re effective radius

Rf strain rate function

Rf,frict strain rate factor for shaft resistance

Rf,bear strain rate factor for tip resistance

Rx roll matrix

Ry pitch matrix

Rz yaw matrix

s distance travelled in the direction of motion

s0 initial distance travelled in the direction of motion

sbz vertical distance travelled coincident with the body-frame

sz vertical distance travelled coincident with the inertial-frame

sz0 initial vertical distance travelled coincident with the inertial-frame

su undrained shear strength

su,op operative undrained shear strength

St soil sensitivity

- 3 -

t time

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

v velocity

vbz velocity coincident with the body frame z-axis

vi impact velocity

vix velocity coincident with the inertial frame x-axis

viy velocity coincident with the inertial frame y-axis

viz velocity coincident with the inertial frame z-axis

vix,0 initial velocity coincident with the inertial frame x-axis

viy,0 initial velocity coincident with the inertial frame y-axis

viz,0 initial velocity coincident with the inertial frame z-axis

V non-dimensional velocity

VS full scale voltage output range

VADC full scale input voltage range of analogue to digital converter

Vdisp volume of the displaced soil

Ws submerged weight

w ball invert depth

ŵop normalised operative depth

ŵdeep-op normalised transition depth

xb body frame x-axis

xi inertial frame x-axis

yb body frame y-axis

yi inertial frame y-axis

z embedment depth (for keying and capacity) or travel distance (free-fall in water and dynamic embedment in soil)

zb body frame z-axis

zi inertial frame z-axis

ziz vertical distance travelled (coincident with the inertial frame)

ziz,e embedment depth (coincident with the inertial frame)

zs separation depth

Greek α adhesion factor

β strain rate parameter (power law)

γ unit weight

γ' effective unit weight of soil

- 4 -

γ strain rate

γref reference strain rate

δrem ratio of fully remoulded strength to intact soil strength

η soil viscosity parameter

θ pitch angle

θb pitch angle coincident with the body frame

θi pitch angle coincident with the inertial frame

θi,acc pitch angle coincident with the inertial frame (derived from accelerometer measurement)

∆N minimum resolvable acceleration resolution

∆R minimum displacement resolution

Δze difference between measured and predicted embedment depth

λ strain rate parameter (semi-logarithmic law)

ν dynamic viscosity

μ strain rate parameter

μ' resultant tilt angle (relative to Earth‟s gravity)

ξ95 accumulated absolute plastic shear strain rate for the 95% remoulding

π ratio of a circle‟s circumference to its diameter

ρ fluid/soil density

σ'v vertical effective stress

τ shear stress

τy shear yield stress

ψb yaw angle coincident with the body frame

ψi yaw angle coincident with the inertial 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

ωix rotation rate about the inertial frame x-axis

ωiy rotation rate about the inertial frame y-axis

ωiz rotation rate about the inertial frame z-axis

ϕ roll angle

ϕb roll angle coincident with the body frame

ϕb0 initial roll angle coincident with the body frame

ϕi roll angle coincident with the inertial frame

ϕi,acc roll angle coincident with the inertial frame (derived from accelerometer measurement)

- 5 -

Abbreviations

ALE Arbitrary Lagrangian-Eulerian

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

CRP Continous Rate Penetration

DCM Directional Cosine Matrix

DEPLA Dynamically Embedded PLate Anchor

DoF Degrees of Freedom

DPA Deep Penetrating Anchor

ESP European Standard Penetrometer

FEP FAU Experimental Penetrometer

FFCPT Free-Falling Cone Penetrometer Test

FFP Free-Falling Penetrometer

FPSO Floating, Production, Storage and Offloading

GB Giga Byte

GME Great Meteor East

GOM Gulf of Mexico

IFFS Instrumented Free-Falling Sphere

IMU Inertial Free-Falling Sphere

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

PROBOS Proboscis

- 6 -

RITSS Remeshing and Interpolation Technique with Small Strain

SEM Scanning Electron Microscope

STING Seabed Terminal Impact Naval Gauge

UWA University of Western Australia

XBP eXpendable Bottom Penetrometer

XDP eXpendale Doppler Penetrometer

- 7 -

CHAPTER 1. INTRODUCTION

1.1. Research motivations

The progression of the offshore energy industry has led to operation in very deep water,

approaching 3000 metres. This has driven the transition from fixed platforms to floating

production facilities which has led to an increased use of subsea facilities such as

flowlines, pipelines and mudmats. The geotechnical design of these infrastructures

requires the precise assessment of the undrained shear strength (su) in the near-surface

soil which is often limited to the upper 1–2 metres and in the case of pipelines, even the

upper 0.5 m (Randolph & White 2008). Offshore soil is often fine-grained and

normally-consolidated with very low strength (< 10 kPa) at the surface with an

increasing strength with depth (1–2 kPa/m). Characterisation of these soils is

challenging, and usually incorporates in situ testing due to the difficulty of recovering

high quality samples in deep water. In situ testing requires the deployment of a large

submersible penetration rig that is lowered to the seabed where a cone penetrometer, T-

bar or ball penetrometer (Figure 1.1) is driven into the soil at a constant rate of 20 mm/s.

The costs associated with these in situ, quasi-static penetration tests are significant and

increase with water depth.

- 8 -

Figure 1.1. Ball Penetrometer, T-bar and Cone Penetrometer

A less expensive and quicker alternative to traditional penetrometer testing is to utilise a

free falling penetrometer (FFP). FFPs are designed so that after release no mechanical

energy is required as the penetrometer uses kinetic energy gained during free-fall in the

water column to penetrate the seabed. Most modern FFP designs are slender full-shafted

projectiles with a 60° conical tip (i.e. similar to a cone penetrometer shown in Figure

1.1). As shown in Figure 1.2, FFP sizes vary depending on the application, with shaft

diameters varying between 0.04–0.11 m (e.g. Stegmann et al. 2006, Mosher et al. 2007,

Stark et al. 2009, Young et al. 2011 and Stephan et al. 2012). In some cases these

devices have flukes near the rear of the penetrometer for stability during free-fall (e.g.

see Figure 1.2c).

- 9 -

Figure 1.2. Examples of full-shafted penetrometers with conical tips: (a) CPT-

Lance (after Stark et al. 2009b); (b) Nimrod (after Stark et al. 2009b); (c) FFCPT

(Furlong et al. 2006); (d) CPT Stinger (after Young et al. 2011)

FFPs usually measure the dynamic penetration resistance in soil in one of two ways: (i)

using a load cell (similar to a CPT) or (ii) using an accelerometer. In the latter case, a

single-axis accelerometer is embedded into the projectile and the forces acting on the

FFP are calculated by considering Newton‟s second law of motion. The undrained shear

strength is then calculated by considering the FFP to be a single particle with a

penetrating force linked to the submerged weight in soil and an opposing force which

- 10 -

comprises the reaction force (i.e. the mass multiplied by linear acceleration) and various

other resistance forces linked in part to the undrained shear strength. One of the most

attractive properties of a FFP is the ease and speed of installation compared to

traditional penetrometers. For example, the penetrometer shown in Figure 1.2c can be

deployed from a constantly moving vessel. This allows a survey a FFP to cover a large

seabed area in a relatively short period of time. This can reduce the cost of seabed

surveys and improve the quantification of spatial variability.

Despite a number of successful FFP field trials undertaken since the 1960s, their

application is not widespread. This is partly due to uncertainties regarding the strain rate

effects in soil and also due to the interpretation of the undrained shear strength which

requires the definition of a bearing capacity factor, Nc. This is traditionally difficult to

ascertain for a conical tipped penetrometer and many numerical and empirical

correlations vary widely (Aubeny & Shi 2006, Nazem et al. 2012, Lunne et al. 1997).

Further uncertainties exist in relation to the frictional resistance generated along the

shaft and the different strain rate dependency of the penetrometer shaft compared to the

tip (Dayal et al. 1975, Steiner et al. 2014).

With these concerns in mind, this thesis introduces a new FFP known as the

Instrumented Free-Fall Sphere (IFFS) which represents a step change compared to the

traditional full-shafted, slender FFPs. The IFFS used in this study is a custom-made,

250 mm diameter mild steel sphere that consists of two hemispheres that are bolted

together with an internal vertically-orientated cylindrical void to accommodate

instrumentation that measures the motion history in water and soil (Figure 1.3 and

Figure 1.4). The sphere and instrumentation weighs 620N in air and has a submerged

weight of 549N in water. Attached to the padeye of the sphere is a retrieval line that is

used for deploying and recovering the IFFS after penetration in soil.

- 11 -

Figure 1.3 Schematic of the Instrumented Free-Fall Sphere

Figure 1.4. Sphere separated to show internally housed data logger housed

In the same way that a ball penetrometer is often preferable in soft soils over a cone

penetrometer, the IFFS could be advantageous over a conical-tipped, slender FFP. This

- 12 -

is because more closely-bracketed plasticity solutions are available for deducing shear

strength from the net penetration resistance of a ball penetrometer compared to a cone

(Randolph & Houlsby 1984, Randolph et al. 2000, Einav & Randolph. 2005). Therefore,

this thesis, for the first time, considers the merit of the IFFS through a set of centrifuge

and field tests and compares the performance to push-in ball penetrometer tests in

numerous soil types.

1.2. Thesis objectives

The main objective of this thesis is to examine the potential for the IFFS to be used as a

tool for measuring the undrained shear strength of near-surface, fine-grained soils.

Encompassed within this main objective are the following specific objectives:

To investigate the performance of the IFFS in a range of soil types through a

series of offshore field tests and physical modelling experiments in a

geotechnical centrifuge.

To develop frameworks that improve the interpretation of the undrained shear

strength from the IFFS measured penetration resistance.

To demonstrate the merit of the tool and the associated interpretative

frameworks by comparing the shear strength profiles derived from the IFFS

with corresponding data obtained from well-calibrated push-in ball

penetrometer tests.

1.3. Research methodology

This thesis comprises two major components – a dynamic penetration component

studying the complex processes associated with the embedment of an Instrumented

Free-Fall Sphere (IFFS) and also a continuous rate penetration (CRP) component

studying shallow penetration effects of a shallowly-embedded ball penetrometer in

- 13 -

kaolin clay. The experimental studies involve field tests and physical modelling

experiments using a geotechnical centrifuge. Eighty-seven separate free-fall field tests

have been undertaken in two separate soil sites, an offshore site in the Firth of Clyde in

Scotland and a lake site in Lough Erne, County Fermanagh in Northern Ireland. Both

soil sites were chosen due to their soft soil conditions, surrounding resources and ease

of access. In the centrifuge, IFFS tests were carried out on three different soil types that

were available within the catalogue of soils at UWA, kaolin clay and two natural soils,

West Africa clay recovered from the Gulf of Guinea and Laminaria clay recovered from

the Timor Sea. These soils are typically well-characterised and are documented in the

literature.

An important aspect of the research was the instrumentation system used to indirectly

measure the undrained shear strength (su). The centrifuge experiments used micro

electromechanical system (MEMS) accelerometers to deduce the net penetration

resistance and hence the shear strength. The field experiments used an inertial

measurement unit (IMU), which measures an object‟s six degree of freedom motion in

three-dimensional space using a combination of gyroscope and accelerometer sensors.

Due to the geometry of the IFFS, it can undergo high levels of rotation during free-fall

and soil penetration. Therefore, the instrumentation adopted in these experiments was

able to interpret the acceleration in the direction of motion. Special attention was given

to the novel instrumentation system and validation of the technique using direct

measurements of the sphere embedment was required in both the field tests and

centrifuge experiments. A comprehensive framework for interpreting the

instrumentation data was developed in order to measure the resistance force acting on

the IFFS during freefall in water and penetration in soil. This led to a new framework

used to describe the dynamic penetration force acting on the IFFS in soil. The

- 14 -

framework accounts for both geotechnical shearing resistance and fluid mechanics drag

resistance, but cast in terms of a single capacity factor that can be expressed in terms of

the non-Newtonian Reynolds number.

The second component of this thesis involved centrifuge tests carried out at 100 g with

an 11.3 mm diameter ball penetrometer. These experiments focused on the penetration

resistance and the degree of hole-closure during continuous rate of penetration (CRP)

tests in kaolin clay sample with a progressively higher overconsolidation ratio. Video

footage observed the progressive hole-closure during each test, and provided a means of

determining the depth at which the cavity, formed by the passage of the ball, closed

over. These data were used in the development of a shallow penetration interpretation

framework which offers a more rigorous and reliable means of assessing soil strength in

the upper few metres of the seabed with a ball penetrometer or an IFFS.

1.4. Thesis organisation

The thesis consists of 9 chapters. A brief outline of each chapter is given below:

1. The first chapter is an introduction to the thesis. The research motivations,

research methodology and thesis organisation is outlined.

2. The second chapter provides a review of the current literature on full-flow

penetrometers and free-falling penetrometers (FFPs) and also discusses the

interpretation of the complex soil-structure interaction that occurs during

dynamic penetration events.

3. Chapter 3, which is published as a conference paper, introduces the

Instrumented Free-Fall Sphere (IFFS) and describes a number of dynamic tests

undertaken in a lakebed. In each test, the IFFS acceleration was measured and

- 15 -

the velocity and displacement profiles were calculated. The chapter also

introduces a theoretical model that describes the motion response of the sphere

during embedment in soil by considering drag resistance and the enhanced shear

strength due to strain rate effects. This model is assessed by comparing

theoretical profiles with the measured profiles using parameters that describe the

strain rate and drag resistance.

Morton, J. P. & O‟Loughlin, C. D., 2012. Dynamic penetration of a sphere in

clay. Proceedings of the 7th International Conference on Offshore Site

Investigation and Geotechnics, London, UK, pp. 223–230.

4. The fourth chapter, which is published as a journal paper considers centrifuge

tests that focus on the evolution of the measured penetration resistance of a ball

penetrometer as it penetrates near-surface soil. The response was observed in

tests on kaolin clay under undrained conditions over a range of undrained

strength ratios. Of particular interest was the transition depth or the depth at

which the open cavity, formed by the passage of the ball, closed over. The

results from the centrifuge experiments, combined with reinterpreted LDFE data

led to the proposal of a theoretical framework that offers a more rigorous and

reliable means of assessing soil strength in the upper few metres of the seabed.

Morton, J. P., O‟Loughlin, C. D. & White. D. J., 2014. Strength assessment

during shallow penetration of a sphere in clay. Géotechnique Letters 4 (October-

December), pp. 262–266.

5. Chapter 5, which is published as a journal paper explores the potential of using a

MEMS accelerometer to measure the motion response of a free-falling anchor in

the centrifuge. The paper describes tests in which a single-axis MEMS and

- 16 -

piezoelectric accelerometer are housed within a dynamically installed model

anchor. The paper concludes that the MEMS accelerometer is capable of

measuring accelerations during both the free-fall phase and the soil embedment

phase, whereas the piezoelectric accelerometer is only able to measure the

changing accelerations that occur during the soil embedment phase. The

measurement technique is verified by comparing the velocity and displacement

profiles derived from numerical integration of the MEMS accelerations to

independent velocity and displacement measurements.

O‟Loughlin, C. D., Gaudin, C., Morton. J. P. & White, D. J., 2014. MEMS

accelerometers for measuring dynamic penetration events in geotechnical

centrifuge tests International Journal of Physical Modelling in Geotechnics,

14(2), pp. 31–39.

6. Chapter 6, which is published as a journal paper, describes using an inertial

measurement unit (IMU), consisting of a tri-axis accelerometer and a three-

component gyroscope to measure the motion response of free-falling projectiles

in the field. The six degree-of-freedom motion data from a number of projectiles

(including the IFFS) was assessed during free-fall in water and penetration in

soil. A comprehensive framework for interpreting the measured data is described

and the merit of this framework is validated by comparing the displacement

derived from the IMU measurements with direct displacement measurements.

Blake, A. P., O'Loughlin, C. D., Morton, J. P., O' Beirne, C., Gaudin, C. & White, D.

W., 2014. In-situ measurement of the dynamic penetration of freefall projectiles

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

Journal. ASTM. DOI: 10.1520/GTJ20140135.

- 17 -

7. Chapter 7, which is a journal paper under review, presents a study of the

dynamic penetration of the IFFS from 87 separate field tests carried out in two

soft soil sites. Instrumentation housed within the sphere measured accelerations

in three orthogonal axes and rates of rotation about the same three axes. These

data were used to calculate the resistant forces acting on the IFFS during free-

fall in water and embedment in soil. This chapter presents a novel approach that

unifies the soil mechanics and the fluid mechanics frameworks to analyse the net

penetration resistance. This allows for an assessment of the undrained shear

strength, su, throughout the depths of penetration. The potential of using the

IFFS as a site investigation tool is also assessed where the undrained shear

strength profiles measured with the IFFS compare well with push-in piezoball

profiles.

Morton, J. P., O‟Loughlin, C. D. & White. D. J., 2015. Field testing an in situ

freefalling spherical penetrometer in soft soil. Submitted to Géotechnique.

8. Chapter 8, which is a journal paper under review, describes centrifuge tests in

which a 20 mm diameter (0.25 m in prototype scale) model IFFS was allowed to

free-fall in water and dynamically penetrate soft soil. Three different natural soil

types were used in the centrifuge – a calcareous soil from offshore Australia, a

high-plasticity West Africa clay and kaolin clay. The paper highlights that an

orthogonal axis measurement system is required to properly capture the linear

acceleration and indirectly measure the forces acting on the IFFS as it penetrates

the soil. The paper presents shear strength profiles derived from the IFFS

measurements for each soil type with equivalent measurements made using a

- 18 -

„pushed-in‟ ball penetrometer.

Chapter 8: Morton, J. P., O‟Loughlin, C. D. & White. D. J., 2015. Centrifuge

modelling of an instrumented free-fall sphere for measurement of undrained

strength in fine-grained soils. Canadian Geotechnical Journal. DOI:

10.1139/cgj-2015-0371.

9. The concluding remarks from the thesis are presented in Chapter 9. This closing

chapter provides a summary discussion of the thesis and highlights the key

conclusions.

- 19 -

CHAPTER 2. REVIEW OF LITERATURE

2.1. Introduction

This chapter is categorised into three main sections. The first section explores the use of

in situ tools and the acquisition of the undrained shear strength (su). The second section

looks at free-falling penetrometers (FFPs). This includes an up-to-date review of the

current literature relevant to FFPs, including projectile geometries, instrumentation and

measurement systems and reported experimental analysis. The third section focuses on

the interpretation of the dynamic penetration forces and the complex processes that take

place at elevated velocities, including inertial effects and strain rate effects. The review

will attempt to demonstrate the merits and uncertainties in applying the available

formulae to account for these phenomena, as well as setting the scene for the

introduction of the IFFS as a new type of FFP.

2.2. In situ tools

In the past two decades the importance of high quality in situ survey tools has

dramatically increased. Traditional offshore site investigation methods remove an intact

soil sample with a tube corer for subsequent laboratory experiments. However,

removing soil from its original position can cause the soil to undergo mechanical

changes due to the relieving of stress and also physical changes due to disturbance

caused by the tube penetration, retrieval and transportation. In addition, chemical

changes, such as the exsolution of gas can take place once the sample is removed from

its original state (Lunne et al. 2006). Therefore, laboratory tests are limited in the sense

that they can be prone to misleading results and significant time is required to obtain the

- 20 -

undrained shear strength. Consequently, modern offshore site investigation chiefly uses

laboratory tests as a calibration tool at discrete intervals for in situ penetrometer testing

that is the principal tool for assessing design strength profiles.

2.2.1. In situ vane

In situ vane testing has been used worldwide since the 1940s and used widely offshore

since the 1970s. Depending on the soil strength, three different sizes of vane are used

ranging in height between 80 mm and 130 mm, all with a height to width fin-ratio of 2.

The vane shear test apparatus consists of a four-blade stainless steel vane attached to a

steel rod that is vertically trust into the ground ahead of the drill bit by a minimum

distance of 0.5 m. The vane is then rotated at a rate of 0.1°/s to 0.2°/s. Typically, a test

at a single depth takes more than 30 minutes for the vane to rotate 0.5–1 revolution.

Additional tests can be undertaken by pushing the vane deeper into the soil, ensuring a

minimum separation distance of 0.5 m. High quality data obtained from offshore tests

carried out in normally consolidated clays in the Gulf of Mexico have been reported

(Quiros & Young 1988) that is of similar quality to onshore laboratory experiments.

However, due to the low testing speed, the test is relatively time consuming and is often

used to supplement continuous profile data at discrete intervals.

2.2.2. Cone and piezocone

The most common in situ tool used offshore is the cone penetrometer test (CPT) which

was first introduced in 1932. The cone consists of a 60° conical tip and measures

penetration resistance, sleeve friction and pore pressure to measure the coefficient of

consolidation (Baligh et al. 1981). Cone sizes vary depending on the objective of the

test. The standard cone has a projected area of 1000 mm2, but smaller cones down to

100 mm2 are employed to penetrate to greater depths or to operate from lighter frames

- 21 -

that offer less reaction force. The cone is penetrated at a standard rate of 20 mm/s and

the resistance is recorded during penetration.

2.2.3. Full flow penetrometers

Full flow penetrometers, such as the T-bar and ball penetrometer (Figure 2.1), have

evolved from the CPT and since their introduction have become commonplace for

characterising offshore soft soils (Randolph et al. 1998). The conical tip is replaced with

a large cylindrical bar or a sphere that is typically 10 times larger than the shaft. The

rationale for replacing the cone is partly due to the higher theoretical rigour of the

appropriate bearing capacity factor (Nc) compared to a cone (Chung & Randolph, 2004)

and also the ability to perform cycles to identify the remoulded strength. Unlike the

CPT, the soil is allowed to flow fully around the device as it penetrates the soil. There

are a number of advantages of a full flow penetrometer compared to the CPT, these

include:

1. The measured penetration resistance requires minimal correction to provide net

penetration resistance.

2. Improved resolution is obtained in soft soils due to the larger projected area of

the penetrometer head which provides a higher penetration resistance.

3. Closely bracketed plasticity solutions are available for deducing undrained shear

strength (su) from penetration resistance. The range of solutions is somewhat

narrower compared to cone factors.

4. The remolded shear strength may be assessed from cyclic penetration and

extraction of full flow penetrometer.

Due to the large projected area of the penetrometer head, the load cell measures a

differential force (or net pressure) so the adjustment for the overburden stress and

ambient pore water pressure is minimal. The „unequal area‟ effect correction, caused by

- 22 -

the presence of the shaft, is typically ten times smaller than the correction that is

required for a CPT. Nevertheless, a correction is required for all shafted penetrometers

due to the effective stress which acts uniformly around the penetrometer head except in

the position where it is attached to the shaft. In order to ascertain the net penetration

resistance, qnet, from the measured penetration resistance, qm the following equation is

considered (Chung & Randolph 2004):

p

sovmnet A

Aα1uσqq 2.1

where σv is the overburden pressure, u0 is the hydrostatic pore pressure, As is the shaft

area, Ap is the area of the penetrometer head, α is the „unequal area‟ ratio. This is

required due to the ingress of water into the inner part of the load cell chamber that

reduces qm (Baligh et al. 1981, Lunne et al. 2007). The „unequal area‟ ratio is the net

area ratio of the load cell core to the shaft, typical values are in the range of 0.7 to 0.9

(Lunne et al. 1997) which can be verified in a pressure chamber.

In addition to a reduced overburden correction factor, the penetration resistance of a full

flow penetrometer is less affected by secondary soil characteristics such as the rigidity

index

and the in-situ stress ratio,

(Teh & Houlsby 1991). This is

not the case for the CPT due to the insertion of additional volume from the cone shaft

into the ground.

Typically the T-bar and ball are penetrated at the same rate as the cone (20 mm/s).

However, unlike the CPT, the resistance during extraction is also recorded throughout

the profile and at least one cyclic penetration and extraction test is usually performed in

order to provide the remoulded strength as well as an indication of load cell offset. In

order to measure the remoulded shear strength, a minimum of 12 cycles should be

- 23 -

undertaken which is usually sufficient to achieve a steady remoulded soil resistance

(DeJong et al. 2010).

Figure 2.1. (a) T-bar; (b) Ball penetrometer

2.2.3.1. T-bar and ball penetrometer

The T-bar (Figure 2.1a) was first developed as a tool to improve the accuracy of

strength profiling in centrifuge tests in the University of Western Australia (Stewart &

Randolph 1991). The model T-bar comprises of a 5 mm diameter by 20 mm long

cylinder attached at right angles to the end of a vertical shaft. The penetrometer is

typically installed at a rate of 1 mm/s and the penetration resistance is measured with a

load cell that is housed within the shaft. The success of the tests led to field trials in

Australian waters described by Randolph et al. (1998). T-bars used in the field are

- 24 -

typically 250 mm in length with a 40 mm diameter. This gives a projected area of 10

times that of the standard cone shaft.

The ball penetrometer was first suggested as an alternative to the T-bar in order to

reduce the potential of axial bending of the shaft (Watson et al. 1998). Centrifuge

experiments on a 12 mm diameter model ball penetrometer were first described by

Watson et al. 1998 and the first field tests were carried out in 2003 as described in

Peuchen et al. (2005). The standard ball penetrometer used in field tests is 113 mm

diameter (area of 100 cm2) with a lightly sand blasted surface.

A recent development for the ball penetrometer has involved fitting pore pressure

sensors to obtain the coefficient of consolidation (Peuchen et al. 2005, Kelleher &

Randolph 2005). The „piezoball‟ ball penetrometer (Figure 1.1 and Figure 2.1b) shows

particular potential as being able to provide data on consolidation characteristics by

means of pore pressure dissipation tests (Low et al. 2007).

2.2.4. Interpretation of shear strength from full flow penetrometers

The main aim of this thesis is to analyse the undrained shear strength from static and

dynamic penetration in soil. The su profile can be determined from the net penetration

resistance (qT-bar or qball) of a deeply embedded T-bar or ball penetrometer using the

following expressions:

barT

barTu N

qs

2.2

ball

ballu N

qs 2.3

where NT-bar and Nball are the bearing capacity factors for a T-bar and ball penetrometer.

- 25 -

2.2.5. Bearing capacity factor

Randolph & Houlsby (1984) first proposed a theoretical solution for the bearing

capacity factor for a T-bar based on the solution for a laterally loaded pipe. The soil was

simplified to a plain strain problem of plasticity theory. An exact solution (i.e. the upper

bound and lower bound were equal) was proposed for a deeply embedded T-bar with

NT-bar values ranging from 9.14 for a fully smooth interface to 11.94 for a fully rough

interface. The upper-bound solution was subsequently corrected by Martin & Randolph

(2006) which raised the upper bound solution for all cases except for a fully rough pipe.

The largest discrepancy between the lower bound and upper bound was found to be

9.1% for the case of a smooth pile. Since the discrepancy between the upper and lower

bounds for NT-bar for all values of the T-bar interface friction ratio (αs) is small, the

original closed-form expression derived by Randolph & Houlsby (1984) may still be

used to describe the variation of NT-bar with αs for ideal non-softening, rate independent

and perfectly plastic isotropic soil:

2sin

sin2

sincos4sincos2sin2

1111 ss

ssbarTN

2.4

The theoretical solutions were applied by Einav & Randolph (2005) using a novel soil

model that incorporated softening and rate effects thus providing a basis for deriving

bearing capacity factors considering these features of soil behaviour. The method was

based on the documented plasticity solutions (Randolph & Houlsby 1984, Randolph et

al. 2000, Martin et al. 2006) and accounted for the effects of strength degradation due to

remoulding and strength enhancement due to strain rate.

The theoretical Nball was derived based on the upper and lower bound approaches used

to derive NT-bar. The flow mechanism was modified to account for the axisymmetric

sphere geometry (Randolph et al. 2000, Einav & Randolph 2005). A Von Mises failure

- 26 -

criterion was utilised to calculate the upper bound N and a Tresca model to calculate the

lower bound theoretical solution. The results range from 10.98 to 15.10 for a fully

smooth and fully rough ball respectively. These values are 20 to 28% (Tresca) or 18 to

21% (Von Mises) higher than the corresponding NT-bar. This method of analysis was

later improved by Zhou & Randolph (2009a) using large deformation finite element

(LDFE) analysis. The effects of strength degradation due to remoulding and strain rate

on NT-bar and Nball are expressed as:

idealbarTremrembarT NeNbarT

95

5.1

18.41

2.5

idealballremremball NeNball

95

5.1

18.41

2.6

where μ is the rate coefficient for a semi-logarithmic strain rate law, δrem is the ratio of

fully remoulded strength to intact soil strength and ξ95 is the accumulated absolute

plastic shear strain rate for the soil to undergo 95% remoulding. ξT-bar is taken as 3.7, ξball

is taken as 3.3. NT-bar(ideal) and Nball(ideal) are the NT-bar and the Nball for non-softening and

rate independent soil.

Equations 2.5 and 2.6 imply that the T-bar and sphere have the same average rate of

strain between the two flow mechanisms. A rate factor of 4.8 in both equations implies

that the average rate of strain experienced by soil flowing around either penetrometer is

some 104.8%/hr, i.e. five orders of magnitude greater than the reference strain of

1%/hour.

When equations 2.5 and 2.6 are considered with typical values of μ and ξ95, (0.05–0.2

and 10–25 respectively), the range of Nball (10–18) is approximately 20% higher than

- 27 -

the range of NT-bar (8–16). Similar results were found from analysing rate-independent

and non-softening soil (Lu et al. 2000) using LDFE analysis. However, this finding has

not been verified experimentally. In the centrifuge and field tests, the ratio has been

found to be almost identical (Chung & Randolph 2004, Boylan et al. 2007, Low et al.

2011, Colreavy et al. 2012). In fact, there is evidence suggesting lower rate dependency

for Nball compared to NT-bar. This may explain the disparity between theoretical and

experimental Nball/NT-bar which is a likely result of subtle soil characteristics such as soil

sensitivity and anisotropy that are not accounted for in the plasticity solution.

Another important finding from equations 2.5 and 2.6 is that NT-bar and Nball depend on

the sensitivity St (or the inverse of sensitivity, δrem, the remoulded shear strength ratio)

and the cumulative shear strain, δ95. This has been verified numerically and

experimentally by Yafrate & Dejong (2006) who found Nball and NT-bar values as low as

7 for high sensitivity soils.

In the field, a simpler approach is often adopted where a value of Nball and NT-deep = 10.5

is usually accepted (Stewart & Randolph 1991). This value has been established through

experimental calibration using a range of material types, stress histories and stress levels

(House et al. 2001). The solution takes into account that penetrometers used in the

centrifuge and in the field are neither fully smooth nor fully rough.

2.2.6. Interpretation of shear strength during shallow penetration

The su in the upper few meters of the seabed is a key parameter for the design of almost

all shallow offshore infrastructure such as, as-laid pipeline embedment (Westgate et al.

2012) and mobile seabed foundations (Feng et al. 2013). Assessing the shear strength

within the first 0.5 m soil depth (less than five diameters for a conventional 113 mm

diameter ball penetrometer) can be difficult with a T-bar or ball penetrometer. In order

- 28 -

to measure near surface strength (~1 kPa), shallow penetration effects must be

considered. Shallow penetration effects alter the usual relationship between the

penetration resistance (which is the measured quantity) and the soil strength (which is

the calculated quantity). The shallow penetration effects occur until the penetrometer

has become sufficiently embedded to reach the transition depth, which can take up to

several diameters embedment (see Section 2.2.6.1). This arises when a full-flow

mechanism around the penetrometer is established after the closure of soil above the top

of the penetrometer. In order to account for the changes, a number of factors must be

considered. These include an enhancement of the soil buoyancy (see Section 2.2.6.5), a

reduction of the bearing capacity factor (see Section 2.2.6.4) and a change to the

operative depth (see Section 2.2.6.6).

2.2.6.1. Transition depth

The transition depth or the limiting cavity depth has been estimated as the depth where

the soil flows over the top of the penetrometer, i.e. the crown of the T-bar or ball

penetrometer. Modern experiments have been carried out for a range of penetrometers

and subsea infrastructure such as T-bars and ball penetrometers (White et al. 2010,

Zhou et al. 2013) and also spudcans (Hossain et al. 2005).

2.2.6.2. Wall failure

Wall failure refers to cavity collapse; this is indicated by inward and downward soil

movements extending up to the soil surface. Skempton (1951) proposed a dimensionless

ratio of kd/sum (where sum is the mudline strength, k is the shear strength gradient and d

penetrometer diameter) in order to assess the stability of a circular footing. The stability

of a circular cavity has been investigated extensively by Meyerhof (1972) using

Rankine‟s pressure theory, and by Britto & Kusakabe (1982, 1983) using upper-bound

- 29 -

plasticity analysis. The cavity depth (Hw) at which wall failure is initiated is expressed

as:

su

wNsH 2.7

where γ' is the effective unit weight, su is the uniform undrained shear strength and Ns is

the stability number. Britto & Kusakabe (1983) formulated an empirical approach for Ns

for a circular footing. This approach, described in (SNAME, 1997), predicts the wall

failure or collapse of a cavity wall. However, with the availability of centrifuge

experiments, a different type of soil failure mechanism has been observed known as

flow failure.

2.2.6.3. Flow failure

Flow failure describes the progressive infilling of soil as it flows from beneath an object

to cover the top. At shallow penetration depths, flow failure is associated with upwards

and outwards soil flow resulting in heave at the soil surface. After reaching a certain

penetration depth, the soil flows back onto the exposed crown of the penetrometer.

During penetration, this gradual infilling continues until at some depth the penetrometer

reaches the transition depth where it becomes fully embedded. The result is a conical

crater that tapers towards the transition depth as shown in Figure 2.2 for a ball

penetrometer in uniform soil. This has been verified by both centrifuge observations

(Hossain et al. 2004b. Zhou et al. 2013) and field test subsea inspections (Kee & Ims

1984).

Centrifuge experiments have shown that the cavity depth due to flow failure is much

shallower than the criterion for wall failure (Hossain et al. 2005). A potential reason for

this is that circular cavity tested in the early studies was much smaller than the

- 30 -

penetrometers that display flow failure. In addition the ratio of strength to self-weight

was higher in the early studies. It is therefore pertinent that the flow failure approach

has been shown to better estimate the transition depth of a penetrating spudcan (Hossain

et al. 2005).

Figure 2.2. Soil failure mechanism for a ball penetration in uniform clay (Zhou et

al. 2013)

In order to predict the transition depth for full flow failure, the soil is grouped into a

non-dimensional parameter which reflects the strength ratio - the ratio of undrained

shear strength (su) to self-weight (γ') and penetrometer diameter (D). Recently, su/γ'D

has been confirmed as the relevant dimensionless parameter to predict the transition

- 31 -

depth for spudcans and the T-bar (Hossain et al. 2005, White et al. 2010). LDFE

analyses using AFENA finite element software using the RITSS method (remeshing and

interpolation technique with small strain: Hu & Randolph 1998) have been conducted

for a T-bar (White et al. 2010, Tho et al. 2012), spudcan (Hossain et al. 2005) and ball

penetrometer (Zhou et al. 2013, Stanier & White 2014). The results have shown that

higher strength ratios are associated with a delay in the transition depth. In the case of a

ball penetrometer, Zhou et al (2013) identified the transition depth over a range of su/γ'D

between 2.95 and 44 (see Figure 2.3). The correlations appear very robust over the

range of su/γ'D. However, the range does not necessarily cover the range of practical

interest of strength ratios and penetrometer diameters that could be encountered in

offshore problems. The author has identified this as a key area that can be expanded to

facilitate a wider range of strength ratios than previously examined.

Figure 2.3. The effect of su/γ’D on the transition depth for a ball penetrometer

(after Zhou et al. 2013)

- 32 -

2.2.6.4. Bearing capacity factor variation with depth

Significant progress has been made in the theoretical studies of bearing capacity factors

for deeply embedded penetrometers. However, at shallow penetration depths a full-flow

mechanism does not develop around a penetrometer and su should be interpreted from

the measured penetration resistance using a shallow bearing capacity factor, Nb-shallow.

This is an important aspect of accurately quantifying the su over the upper 1 to 2 metres

of the seabed, if the shallow bearing capacity is ignored, the su may be underestimated

in the near surface soil (White et al. 2010).

Much of the available information on Nb-shallow comes from the study of shallow circular

foundations in non-homogeneous clay (e.g. Skempton 1951, Salencon & Matar 1982,

Houlsby & Wroth 1983, Kusakabe et al. 1986, Tani & Craig 1995, Martin & Randolph

2001). The Nb-shallow variation was first shown for a T-bar by White & Randolph (2007)

and result from LDFE and centrifuge analysis led to a shallow penetration framework

for a T-bar (White et al. 2010) and a ball penetrometer (Zhou et al. 2013). Both

frameworks derived a simple formula to quantify the variation of Nb-shallow with depth

which was conveniently defined as a proportion of the deep bearing factor, Nb-deep which

varies with the dimensionless strength ratio su/γ'D.

2.2.6.5. Soil buoyancy

Whilst a penetrometer is submerged in soil, the soil buoyancy increases the penetration

resistance. If soil heave at the mudline is ignored, Archimedes‟ principle can be used to

calculate the net upward force equal to the volume of the displaced soil multiplied by

the effective unit weight (γ') of the soil (Fbuoy = Vdispγ'). However, if there is a delay in

reaching the transition depth (which is the case for soils with a higher su/γ'D), there is an

extra component of soil buoyancy. This is caused by the cavity, which is formed by the

passage of the penetrometer. The soil that would have filled the cavity is instead

- 33 -

accommodated by heave at the soil surface. This increases the buoyancy contribution

because different degrees of soil heave, and thus work done against soil self-weight are

required at different depths. The heave effect has been captured for a T-bar and ball

penetrometer using a multiplier applied to the buoyancy force (Merifield et al. 2009,

White et al. 2010, Chatterjee et al. 2012, Zhou et al. 2013). The multiplier gradually

reduces to 1 when the penetrometer has reached the transition depth. For pipelines,

Merifield et al. (2009) suggested that the multiplier factor should be = 1.5. Chatterjee et

al. (2012) used LDFE analyses to show that it varies around ~ 1.5, depending on the

dimensionless shear strength gradient, kD/su,avg,whereas Zhou et al. (2013) estimated the

buoyancy factor as 3 for a ball penetrometer.

2.2.6.6. Operative depth

An important correction to the shallow penetration of a full flow penetrometer is to

account for the operative depth, (i.e. the depth at which the interpreted soil strength

applies). For a deeply embedded penetrometer with a symmetrical failure mechanism

above and below the penetrometer, qnet, and hence the inferred su, corresponds to the

mid-height of the penetrometer. This is highlighted in Figure 2.2 for a ball penetrometer

at the transition depth. For a shallowly embedded full-flow penetrometer, the

normalised operative depth, ŵop, is assumed to vary linearly from the invert to the mid-

height position which occurs at the transition depth (Hc) (White et al. 2010):

0.5DDH

0.5D

wc

op

www 2.8

where w is the invert depth and D is the penetrometer diameter.

- 34 -

2.3. Free-falling penetrometers

Free-falling penetrometers are expendable or retrievable projectiles designed to be

released from a vessel and dynamically penetrate the seabed after free-fall in water.

Typically, FFPs are slender in geometry and can be categorised into two designs (i)

thin-shafted and (ii) full-shafted, a number of penetrometer designs incorporating a

number of flukes for hydrodynamic stability. FFPs have been considered for a range of

offshore applications including radioactive waste disposal, naval mine countermeasure

research, deep water anchoring systems and paleolimonology applications, as well as

tools to measure the seabed strength. A review of the FFP experiments is provided here,

with particular emphasis on FFPs that estimate the undrained shear strength. The review

focuses on the FFP design and also the instrumentation system used to measure the

dynamic penetration resistance.

2.3.1. Free-falling penetrometer designs for naval mine countermeasure applications

A total of seven FFPs have been designed for naval mine countermeasure applications,

shown in Figure 2.4. These include the Seabed Terminal Impact Naval Gauge, STING

(Mulhearn et al. 1998); Electronic Strength Profiler, ESP (Mulhearn et al. 1998);

Burying Mock Mine Body, BMMB (Mulhearn et al. 1999); Australian Underwater

Sediment Strength Instrument, AUSSI (Mulhearn et al. 1999); FAU Experimental

Penetrometer, FEP (Stoll et al. 2007) and Proboscis, PROBOS (Stoll et al. 2007). These

FFPs are designed to estimate the shear strength in the upper few metres of soil and are

not necessarily designed to penetrate as deep as possible in the seabed. Therefore, a

number of these FFPs are light enough to be installed by hand, such as the XBP (see

Section Free-falling penetrometer designs for seabed characterisation2.3.2), Nimrod

(Figure 1.2b), ESP and FEP.

- 35 -

In order to control the penetration resistance in different soil types, a number of FFPs

shown in Figure 2.4 are designed with interchangeable tip shapes and sizes (e.g. XDP,

AUSSI, FEP, PROBOS and STING) with most exhibiting flat or blunt ends (STING,

ESP, BBMB, AUSSI, FEP). Most FFP designs are full-shafted projectiles except the

STING (Figure 2.4b) which is a thin-shafted FFP with a 19 mm diameter shaft and a

tip-diameter in the range 25 - 70 mm (Abelev et al. 2009b). The merits of using a thin-

shafted FFP will be discussed in the following sections.

- 36 -

Figure 2.4. 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)

- 37 -

2.3.2. Free-falling penetrometer designs for seabed characterisation

A number of other FFPs have been developed for seabed characterisation (summarised

in Figure 2.5) including the Marine Impact Penetrometer (Dayal 1974); Marine

Sediment Penetrometer, MSP-1 and MSP-2 (Colp et al. 1975); free-fall penetrometer

(Denness et al. 1981); Expendable Bottom Probe, XBP (Akal & Stoll 1995); Free Fall

Cone Penetrometer, FFCPT (Mulukutla 2009); CPT lance (Stegmann 2007), Nimrod

(Stark et al. 2009a) and the CPT Stinger (Young et al. 2011, Jeanjean et al. 2012). These

FFPs are designed to penetrate as deeply as possible in the seabed. Therefore, they are

slender in geometry and often allow the inclusion of additional weights to increase the

penetration force (e.g. Marine Impact Penetrometer and LIRmeter). Most of the FFPs

shown in Figure 2.5 are full-shafted, conical-tipped penetrometers with shaft diameters

ranging from 36 to 70 mm and lengths ranging from 0.02 to 8.14 m.

- 38 -

Figure 2.5. FFP systems developed for applications in seabed characterisation (a)

Marine Impact Penetrometer (after Dayal et al. 1975), (b) MSP-2 (after Colp et al.

1975), (c) Freefall penetrometer (after Denness et al. 1981), (d) XBP (after Stoll &

Akal 1999), (e) FFCPT (www.brooke-ocean.com), (f) LIRmeter (after Stephan et

al. 2012), (g) CPT-Lance (after Stark et al. 2009b), (h) Nimrod (after Stark et al.

2009b), (i) CPT Stinger (after Young et al. 2011)

- 39 -

2.3.3. FFP measurement systems

FFPs measure the resistance force during penetration in soil in one of two ways, either

directly, using a load cell (or combination of load cells) (e.g. Steiner et al. 2012, Chow

et al. 2014) or indirectly, using an acoustic Doppler shift system or an accelerometer.

The indirect measurement technique calculates the resistance force by considering

Newton‟s second law of motion where the FFP is considered to be a single particle.

2.3.3.1. Acoustic Doppler system

The acoustic Doppler instrumentation system comprises a sound source mounted onto

the FFP and a receiving surface hydrophone on a support vessel. The frequency of the

signal received on the hydrophone is considered alongside the speed of sound in water

and soil to calculate the penetrometer velocity. The velocity data can be integrated to

provide an indirect calculation of penetration depth, thus allowing the vertically-

orientated motion response to be captured.

The Doppler measurement system is still in use today in the XDP which has been used

in over 270 individual deployments (e.g. Beard 1985, CYR 1990, Bowman et al. 1995,

Thompson et al. 2002, Ortman 2008) and results indicate that in the upper 2 m of soil,

the XDP provides reasonable assessment of undrained shear strength and penetration

depth (Beard 1981).

2.3.3.2. Accelerometer system

The majority of modern FFP designs (e.g. FFCPT, XBP, STING and AUSSI) are

instrumented with an internally-housed single-axis accelerometer that measures the

vertically-orientated acceleration during free-fall in water and dynamic penetration in

the seabed. The acceleration profile is numerically integrated in order to produce the

velocity profile which in turn is numerically integrated to produce the displacement

profile. The first accelerometers to be trialled in FFPs (e.g. Dayal 1974, Colp et al.

- 40 -

1975) and deep sea corers (Scott, 1970) were piezoelectric accelerometers. These

accelerometers rely on the use of a piezoelectric material as the sensing component

where the electrical output-signal is proportional to the stress applied to the

accelerometer. Piezoelectric accelerometers exhibit wide linear frequency ranges and

large amplitude ranges but struggle to accurately measure low frequency events. Early

FFP experiments using piezoelectric accelerometers reported some anomalous results

(e.g. Colp et al. 1975, Dayal 1974, Chari et al. 1978). These results are likely to occur

because during free-fall, a piezoelectric accelerometer loses signal as the piezoelectric

material discharges electrostatic charge (Stringer et al. 2010).

In a number a recent FFP experiments (Lavieri et al. 2011, Blake & O‟Loughlin 2015)

the piezoelectric accelerometer has been replaced with a MEMS accelerometer. MEMS

accelerometers have a number of advantages over piezoelectric accelerometers. For

example MEMS accelerometers can measure low frequency events and have

comparable performance during high frequency events (Stringer et al. 2010). Typical

acceleration profiles for a FFP test, measured with a MEMS accelerometer are shown in

Figure 2.6 and Figure 2.7. Figure 2.6 shows the measured acceleration alongside the

velocity and displacement profiles of a projectile during free-fall and dynamic

penetration in soil and Figure 2.7 highlights the parabolically-shaped acceleration

profile after the FFP has impacted the seabed.

- 41 -

Figure 2.6. Typical velocity and penetration depth with time profiles for the FFP

instrumented with an accelerometer (Chow & Airey 2010b)

Figure 2.7. Acceleration profile in soil (after Stephan et al. 2012)

2.3.4. Experimental and field studies on free-falling penetrometers

As shown in Figure 2.4 and Figure 2.5 a relatively large number of FFP designs have

been tested at full-scale. Some of the FFPs including the XDP, XBP, STING, FFCPT,

CPT-Lance and Nimrod have been widely investigated in multiple locations. For

example, the XDP, (shown in Figure 2.4a) has been extensively tested in over 270 field

tests (Beard 1977, 1981, 1985, Bowman et al. 1995, Douglas & Wapner 1996,

Thompson et al. 2002, Ortman 2008) and the XBP (Figure 2.5d) has been tested in over

- 42 -

500 field tests in various locations such as New York Harbour, Italy, Spain, Germany,

Turkey and in the Black Sea and the Gulf of Mexico (Stoll & Akal 1999).

Results from full-scale field trials suggest that the final embedment depth of a FFP is

found to depend on its mass, soil type, penetration velocity and geometry. For example

the XDP reported by Bowman et al (1995) weighed approximately 0.26 kN and

achieved tip embedments between 0.5–3 m. Whereas the XDP used by Beard (1977,

1981, 1985) was longer and weighed approximately 1.31 kN and reported larger tip

embedments in the range 5.8–10 m.

In addition to mass, the final embedment depth of a FFP is heavily dependent on soil

type with reduced embedment depths being observed in sandy soil compared to soft

clay. For example, very low tip-embedment of less than 0.75 m has been reported from

experiments carried out on the ESP (Figure 2.10) which comprised 32 tests in carbonate

sands and muds. Similar results have been reported using the AUSSI (Mulhearn et al.

1999), BBMB (Mulhearn et al. 1999), ESP (Mulhearn et al. 1998) and STING

(Mulhearn et al. 1999) where very low tip-embedment in the range of 0.22 to 1 m were

achieved in hard/shelly sand in Sydney harbour. Stark et al (2009b, 2012) also reported

very low tip-embedment in the range 0.04 to 0.34 m in North Sea quartz sand and also

carbonate sand in Hawaii. At the same test sites, the CPT-Lance (Figure 2.5g) achieved

very low tip-embedment between 0.08 to 0.12 m (Stark et al. 2009b) whereas much

higher tip-embedment of approximately 5 to 6 m was reported in Osthafen silty mud

(Stegmann 2007) and Holocene silty clay (Steiner et al. 2012). Laboratory experiments

carried out on the Marine Impact Penetometer (Dayal 1974, Dayal et al. 1975), XBP

(Akal & Stoll 1995) and PROBOS (Stoll et al. 2007) in sand revealed similar results

where the low tip-embedment was due to the effects of dilative behaviour. This infers

that the application of FFPs may be restricted to soft soils such as clays and silts.

- 43 -

The effect of penetration rate has been shown to play an important role in the

penetration resistance of FFPs. This is highlighted in Figure 2.8 which shows the

penetration resistance acting on a dynamically installed PROBOS (Figure 2.4g)

compared to a quasi-static cone penetrometer (Stoll et al. 2007). These laboratory test

results show that the penetration resistance is increased by a factor of 2 to 3 due to the

strain rate effect. Quantifying the rate effect - which may be due to drainage or viscous

effects - remains to be one of the largest uncertainties in the acquisition of strength

properties using FFPs as described in section 2.6.3.

Figure 2.8. Cone tip resistance of the PROBOS compared to the STATPEN, a

quasi-static cone penetrometer (after Stoll et al. 2007)

Results from field trials suggest that the FFP geometry influences the accuracy of the

measured undrained shear strength. Field trials carried out with the STING in the Gulf

- 44 -

of Mexico (Abelev et al. 2009b) and Sydney harbour (Mulhearn 2003) investigated the

shaft diameter effect by using different circular plate diameters ranging from 23 mm to

70 mm. The results indicated that the 70 mm diameter plate appeared to give the best

agreement with a vane shear test as shown in Figure 2.9b. This is likely to be due to the

reduced contribution from the dynamic shaft resistance.

Figure 2.9. (a) Different tip designs for the STING; (b) Tip diameter effect in

interpreted soil strength for the STING (after Abelev et al. 2009b)

Despite a number of successful full-scale FFP field trials (e.g. Stegmann et al. 2006,

Mulukutla 2009, Stark et al. 2009a, Abelev et al. 2009b, Young et al. 2011 and Stephan

et al. 2012) their application is not widespread. The limited application of FFPs is partly

due to uncertainties in the interpretation of the su which is challenging for a full-shafted,

conical-tipped FFP (e.g. XDP, XBP, ESP, CPT-Lance, MSP and Marine Impact

Penetrometer). This is due to the wide range of potential normalised bearing capacity

factors reported for a conical penetrometer (Nazem et al. 2012, Lunne et al. 1997). This

- 45 -

problem is somewhat reduced when a thin-shafted FFP (e.g. STING) is considered.

Although not as well investigated as a full-shafted FFP, the thin-shafted FFP seem to

provide a better assessment of the undrained shear strength due to the negligible

dynamic shaft resistance (Mulhearn 2003). This reduces the complexity of the data

interpretation, and avoids the difficulties caused by the shaft resistance which appears to

attract greater rate effects than the tip resistance (Dayal et al. 1975, Steiner et al. 2014).

In addition, a thin-shafted geometry is similar to the full-flow penetrometers (described

in Section 2.2.3), where a narrow band of bearing capacity factors have been derived

from plasticity theory (see Section 2.2.5).

2.4. Oceanic waste carriers

Research on the dynamic penetration of streamlined projectiles into the seabed

commenced in the 1960s in order to assess the feasibility of using FFPs as oceanic

waste carriers for the disposal of radioactive material. Field experiments trialled a

number of projectiles at two sites in the Atlantic Ocean, the Great Meteor East (GME)

(Freeman et al. 1984, Freeman et al 1988) and the Nares Abyssal Plain (NAP) (Freeman

& Burdett 1986). The first tests was carried out using four similar penetrometer designs

including the European Standard Penetrometer, shown in Figure 2.10, which measured

3.25 m long, 0.325 m diameter and weighed 1.8 tons in air (Freeman et al. 1984). The

European Standard Penetrometer was subsequently adopted as the standard for

comparison with later experiments at the NAP site. Results indicated that impact

velocities of 30 to 35 m/s were achieved, resulting in embedment of up to 46 to 51 m at

the GME site. At the NAP site, impact velocities between 45 to 56 m/s were achieved,

resulting in penetration depths of 21 to 35 m. Separate field tests relating to the disposal

of nuclear waste were also carried out in Antibes in the Mediterranean Sea. A total of

nine tests were conducted with five different penetrometer designs resulting in

- 46 -

penetration depths in the range of 9 to 15 m (Audibert et al. 2006). The reduced

embedment depth was reported to be due to much stiffer soil in the Mediterranean site

compared to the GME and NAP sites.

Figure 2.10. The European Standard Penetrometer (Freeman & Burdett, 1986)

2.5. Dynamically installed anchors

Four types of dynamically installed anchor (Figure 2.11) have been developed: the Deep

Penetrating Anchor (DPA) (Lieng et al. 1999), the Torpedo Anchor (Medeiros. 2002),

the OMNI-Max anchor (Zimmerman & Spikula 2005) and the Deep Embedded Plate

Anchor (DEPLA) (Blake et al. 2012). These anchors have mainly been developed in an

effort to provide floating production platforms with a cost effective anchoring solution.

The DPA and Torpedo anchor are similar in function insofar as they are designed to

deeply-penetrate the seabed, typically up to 3 times the anchor length (O'Loughlin et al.

- 47 -

2004b) and come to rest in a vertically orientated position. The geometry of these

anchors is relatively simple; they are arrow shaped and have a trailing chain from the

end of the flukes. They comprise a cylindrical steel pipe with a streamlined tip and a

padeye to connect to the anchor chain. The anchor is filled with scrap metal and

concrete to increase the weight and maintain the centre of gravity towards the anchor tip

(Medeiros et al. 1997). Since their introduction, both torpedo anchors and DPAs have

increased in size and mass. The maximum reported length is 17 m for the Torpedo

anchor (Brandão et al. 2006) and 22 m for the DPA (Lieng et al. 1999) with the mass of

each anchor approaching 1000 kN.

The DEPLA and OMNI-Max anchors are installed in much the same way as the torpedo

anchor and the DPA. However, after free-fall and dynamic embedment, the anchors or

part of the anchor is designed to rotate or “key” when a pull-out load is applied, thus

increasing their overall pull out resistance. The DEPLA (Figure 2.11b) is designed so

that after seabed penetration, the central shaft of the DEPLA is retrieved, leaving the

anchor flukes (which constitute the load bearing element as a plate anchor) in the

seabed.

- 48 -

Figure 2.11. Dynamically installed anchors (Medeiros, 2002); (b) Torpedo Anchor

(Brandão et al. 2006)

2.5.1. Experimental and field studies on dynamically installed anchors

The Torpedo anchor has seen widespread commercial use in Brazilian waters since the

late 1990s. Medeiros (2002) reported the first commercial application of torpedo

anchors for anchoring flexible risers in the Campos Basin, Brazil in water depths of up

to 1000m. Since then, over 1000 torpedo anchors have been installed in offshore Brazil

for the mooring of flexible risers, MODUs and FPSOs (Wilde 2009).

- 49 -

The DPA (Figure 2.11a) was first introduced by (Lieng et al. 1999) as an alternative to

the Torpedo anchor. Although the anchor is not in wide use today, a number of reduced-

scale and full-scale tests have been carried out (Strum et al. 2011, Lieng et al 2010). The

installation of a DPA is achieved by lowering the anchor from a stationary vessel to a

target height (typically 40–50 m above the seabed). In spring 2003, twelve 1–3 scale, 4

m long DPA models were tested in both the Trondheim Fjord and at the Troll Field in

the North Sea. In these experiments, at least one anchor was instrumented with a depth

sensor, accelerometer, inclinometer, and pore pressure sensors. However, the results of

these tests have not been published.

In 2008, a total of twelve 1:3 reduced-scale model DPAs were tested in the Troll Field

in the North Sea. An important finding from the 1:3 scale tests was that the

hydrodynamic stability was verified. During free-fall, very little lateral drift was

recorded and on average a 2° tilt was observed for tests released from 15 m to 75 m

above the mudline which embedded between 7 and 9 m (Strum et al. 2011). In 2009,

two full-scale DPAs were installed on StatoilHydro's Gjøa field in 360 meters of water.

The anchors were 13 m in length and had a mass of 736 kN. The DPAs were dropped

from heights of 50 m and 75 m and achieved impact velocities of 24.5 and 27 m/s,

corresponding to tip-embedment of 24 m and 31 m respectively (Lieng et al. 2010).

Full scale tests of the OMNI-Max anchor have been conducted in over 1600 m of water

in the Gulf of Mexico (Shelton, 2007) in soft to medium clay using a 9 m long anchor

with a mass of 333 kN (Zimmerman, 2007). The tests demonstrated the stability of the

anchor during free-fall. In addition, the same anchor design has been used for the

mooring of MODUs in the Gulf of Mexico in 427 m water depth (Shelton, 2011).

- 50 -

A number of field tests using reduced-scale DEPLAs have been reported by (Blake &

O‟Loughlin 2012, 2015, Blake et al. 2014). The tests demonstrated that the tip

embedment was similar to other dynamically installed anchors and was in the range 2–

3.7 times the anchor length. However, the studies reported far higher pull-out capacity

compared to the Torpedo anchor and DPA due to the large projected area of the rotated

flukes in the seabed.

2.5.2. Centrifuge experiments

Centrifuge experiments are often used in the analysis of dynamically installed anchors

partly because high prototype velocities up to 25 m/s cannot be achieved with reduced

scale models at 1 g. Centrifuge experiments have been carried out on reduced-scale

1:200 model DPAs (Figure 2.11a, Richardson 2008, O‟Loughlin et al. 2009), Torpedo

anchors (Figure 2.11d, Hossain et al. 2014), OMNI-Max (Figure 2.11c, Gaudin et al

2013) and DEPLA (Figure 2.11b, O‟Loughlin et al. 2014).

The results are similar to the FFP field test results where the final embedment depth is

dependent on soil type. The experiments carried out in normally consolidated kaolin

clay typically embedded between 2.0–2.9 times the anchor length (O‟Loughlin et al.

2009, (O‟Loughlin et al. 2014) and in slightly overconsolidated kaolin clay the anchors

typically embedded between 1.2–2.1 (Hossain et al. 2014) and 1.8–2 (Gaudin et al

2013). However, the experiments carried out in stiffer calcareous silt revealed far lower

embedment depths in the range 1.14–1.46 (Gaudin et al. 2013) and approximately 1.17–

1.4 (Hossain et al. 2014).

- 51 -

2.6. Resistance forces acting on a projectile during dynamic

embedment in soil

The resistance forces acting on a deeply-embedded full-shafted FFP (with a 60° conical

tip, similar in geometry to most FFPs (e.g. Stegmann et al. 2006, Mosher et al. 2007,

Stark et al. 2009, Young et al. 2011, Stephan et al. 2012) and a thin-shafted FFP (similar

in geometry to the STING with a cylindrical tip, see Figure 2.4b) are shown in Figure

2.12a and Figure 2.12b. The net resistance force of each FFP comprises the drag force

(Section 2.6.1), the bearing force (Section 2.6.6) and for the case of the full-shafted

FFP, the side friction force (Section 2.6.6). Additional resistance forces (not shown on

Figure 2.12a and Figure 2.12b) that have been considered in the analysis of FFPs

include the added mass (Section 2.6.2) and the strain rate-enhanced soil resistance

(Section 2.6.3.1).

- 52 -

Figure 2.12. (a) Forces acting on a full-shafted penetrometer and (b) Thin-shafted

penetrometer during installation

2.6.1. Fluid drag

The fluid drag or drag refers to the resistance force that a moving object is subjected to

in a fluid. The total drag is the sum of two components: (i) form drag, also commonly

known as pressure drag, caused by an adverse pressure gradient between the front and

rear of the object which creates an opposing force and (ii) friction drag, also referred to

as viscous drag. The friction drag is dominant during very low flow rates, however, for

most practicable cases, the contribution of the friction drag to the total drag is usually

- 53 -

low ~ 2–3% for bluff bodies such as a cylinder or a sphere (Achenbach 1971). The drag

force is expressed as (Morison et al. 1950):

2

21 AvCF Dd 2.9

where ρ is the fluid density, A is the area and v is the velocity. CD is the drag

coefficient, dictated largely by object geometry, i.e. the ratio of body length over

diameter, surface roughness and Reynolds Number (Pazwash & Robertson 1975).

2.6.1.1. Drag in water

In classical fluid mechanics studies of Newtonian fluids (e.g. air and water), the shear

stress is linearly proportional to the shear strain rate and the Reynolds Number, Re is

defined as the ratio of the inertia forces to the viscous forces, expressed as:

vDRe 2.10

Where D, is the projectile diameter, and is the temperature and salinity-dependent

kinematic viscosity of the fluid,

where, ɳ is the absolute viscosity.

For FFPs, it is commonly assumed that the CD is the same in both water and soil

regardless of any difference on Re with reported values ranging widely from 0.2–1.2

(True 1976, Freeman et al. 1984, Freeman & Burdet 1986, Bowman et al. 1995, Øye

2000, Cenac II 2011, Hansaloo et al. 2012).

However, for a simplified geometry such as a sphere, the experimental correlation for

CD versus Re has been investigated extensively for Re < 2 x 105. A non-exhaustive list

of the experimental data includes: (Allen 1900, Shakespear 1914, Wieselsberger, 1923,

- 54 -

Bacon & Reid 1924, Liebster 1927, Lunnon 1928, Schmiedel 1928, Millikan & Klein

1933, Moller 1938, Davies 1945, Pettyjohn & Christiansen 1948, Gunn & Kinzer 1949,

Maxworthy 1965, 1969, Goin & Lawrence 1968, Pruppacher & Steinberger 1968,

Dennis & Walker 1971, Roos & Willmarth 1971, Achenbach 1972, Hartman et al.

1994, Beard & Pruppacher 1969, Rimon & Cheng 1969). The correlation for CD and Re

agree well with theoretically derived parameters for various Re ranges. For example, for

Re < 0.5 the correlation agree with theoretical values derived by Stokes (1880) and for

all ranges if Re, the correlation agree with theoretical values derived using a

computational fluid dynamics (CFD) (e.g. Richardson. 2008, Jones & Clarke 2008).

Figure 2.13. Drag coefficient for uniform flow past a sphere R = Re < 2 x 105 (480

data points) after (Brown & Lawler 2003)

For a sphere in the laminar flow regime, Re << 1 no separation between the fluid and

sphere exists and the friction drag dominates, resulting in very high CD. With increasing

- 55 -

Re, the flow regime goes through a number of transitions leading to the drag crisis,

highlighted in Figure 2.14. Firstly, small double vortices are formed as separation

occurs at the azimuthal angle of approximately 80°, highlighted in Figure 2.15a. The

flow is separated but steady and periodic with a very wide wake that generates a large

amount of drag. Vortex shedding has been identified in this region; this refers to an

unsteady oscillating flow that takes place when a fluid flows past a bluff body

(Govardhan & Williamson 2005). If the sphere mass is below a critical mass, the vortex

shedding causes vortex-induced vibration.

Figure 2.14. Drag coefficient for uniform flow past a sphere Re > 2 x 105

(Achenbach, 1972)

At Re ~ ≥ 2 x 105, the CD takes a sudden dip as shown in Figure 2.14. This phenomenon

is referred to as the drag crisis and is due to the reduction of the size of the turbulent

wake (see Figure 2.15b). During the drag crisis, the separation point moves further

downstream increasing the azimuthal angle from ~80 to ~110° (Fage 1936,

Suryanarayana & Prabhu 2000, Bakić 2004).

- 56 -

Figure 2.15. Laminar-separated flow and turbulent flow over a sphere - (after

Finnemore & Franzini 2001)

2.6.1.2. Drag in soil

Compared to the large amount of research on Newtonian fluids, very little is known

about the drag forces during flow through soil. However, most geotechnical studies on

FFPs have considered the drag force in soil in view of the very soft viscous soil that is

often encountered at the surface of most seabed deposits (e.g. True 1976, Bowman et al.

1995, Mulhearn et al. 1998, Fernandes et al 2006, Abelev et al. 2009b, O‟Loughlin et al.

2009). In these studies it is commonly assumed that the drag coefficient is the same in

both soil and water. A small number of fluid mechanics experimental studies have been

carried out in non-Newtonian fluid such as debris flow material (Coussot et al. 1998)

and clay-water mixtures, similar to the upper layer of an offshore seabed (e.g. Houwink,

1952, Pazwash & Robertson 1969, Robertson & Pazwash 1971). The results indicate

that these fluids exhibit non-Newtonian behaviour, formulated in terms of Herschel-

Bulkley model which accounts for the effect of strain rate using an additive term:

- 57 -

ny 2.11

where τ is the shear stress, η is a measure of viscosity and, is the yield stress or the

minimum shear stress required to initiate flow at a particular strain rate, and n is the

shear-thinning (pseudoplastic) index.

Within the fluid mechanics framework, the drag force is estimated using Equation 2.9,

and the CD has been proposed to be a function of the non-Newtonian Reynolds number,

Renon-Newtonian where,

2

Re vNewtoniannon

2.12

where ρ is the density, v is velocity and τ is the mobilised shear stress within the flow

material (Equation 2.11). Research on the rheology of debris flow materials, such as an

underwater landslide on a submerged pipeline has been carried out and an empirical

correlation has been proposed (Zakeri et al. 2008, Zakeri et al. 2009):

25.1Re5.174.1

NewtoniannonDC

2.13

2.6.2. Hydrodynamic mass force

In the case of non-stable flows where the object is accelerating or decelerating, there is

an additional force opposing motion called the hydrodynamic mass force. This force is

required because some volume of fluid (V) surrounding an object is accelerated from

rest to a velocity capable of displacing the fluid out of its path as the object and fluid

cannot occupy the same physical space simultaneously (Sumer & Fredsoe 1997). The

- 58 -

force is manifested as inertia or added mass, m' that is added to the mass of the system

where m' can be calculated in the traditional way (Lamb 1932):

VCm m 2.14

Where Cm is the geometry dependent hydrodynamic-mass coefficient of a sphere that

has been shown both experimentally and theoretically to be equal to 0.5, ρ is the density

of the fluid and V is the enclosed volume of the accelerated body.

2.6.3. Strain rate effects in clay

The geotechnical resistance forces acting on the FFPs shown in Figure 2.12 (i.e. the

bearing and shaft resistance) are linked to the undrained shear strength which is

considered to be a function of strain rate. This has been backed up with a large database

of plates, cones and full-flow penetrometer tests in clays (e.g. Murff & Coyle 1973,

Lehane et al. 2009). Figure 2.16 plots the normalised penetration resistance and the

normalised velocity (V = vD/ch) where v is the penetration velocity, D in the diameter

and ch is the horizontal coefficient of consolidation. Below a V of ~ 15, consolidation is

dominant and viscous effects are almost negligible (Lehane et al. 2009).

- 59 -

Figure 2.16. Normalised velocity illustration (after Lehane et al. 2009)

Above a V ~ 15, Figure 2.16 highlights that the viscous effects dominate and the shear

strength increases with increasing strain rate (Casagrande & Wilson 1951, Sheehan et

al. 1996, Graham et al. 1983, Biscontin & Pestana 2001). The mobilised or operational

undrained shear strength, su,op can be accounted for by a single multiplicative parameter,

Rf:

refufopu sRs ,. 2.15

where su,ref is the undrained shear strength at the reference strain rate ref which is

usually taken as the strain rate where the viscous effects cease to exist (Hyde et al. 2000,

Lehane et al. 2009, Randolph & Hope 2004). The rate function is usually expressed

using a semi-logarithmic function (e.g. Graham et al. 1983, Chung et al. 2006) or a

power law (Biscontin & Pestana 2001, Lehane et al. 2009):

- 60 -

reffR

log1 2.16

reffR

2.17

where and and β are the strain rate parameters representing the increase in shear

strength.

2.6.3.1. Strain rate parameter

For variable rate penetrometers a representative FFP‟s the strain rate is usually taken as

the normalised penetration rate, v/D (where v is the velocity and D is the diameter,

(Lehane et al. 2009, Stark et al. 2012, Steiner et al. 2012, O‟Loughlin et al. 2013). The

reported strain rate parameters have been found to be dependent on the strain rate range

(Sheehan et al 1996, Jeong et al 2009, O‟Loughlin et al 2013b). For example, the semi-

logarithmic rate formula (Equation 2.16) has been found to increase from 7% to 11.5%

for a corresponding increase in strain rate of 0.05%/hr to 50%/hr in triaxial rate studies

(Sheahan et al. 1996). Similarly, back analysed values of λ and β from centrifuge tests

on dynamically installed anchors in normally consolidated kaolin clay have been shown

to increase with increasing impact velocity (e.g. Richardson 2008, O‟Loughlin et al

2013b) as shown in Figure 2.17.

- 61 -

Figure 2.17. Variation in back-analysed β and λ strain rate parameters with

impact velocity for reduced scale model DPAs from centrifuge experiments in

kaolin clay (after O’Loughlin et al. 2013b)

A large range of strain rate parameters have been reported in the literature. Results from

variable rate penetrometer tests (e.g. Randolph & Hope 2004, Chung et al. 2006, Boylan

et al. 2007, Lehane et al 2009, Young et al 2011 Steiner et al 2013) indicate that λ and β

generally lie in the range λ = 0.1–0.2 and β = 0.06–0.08 These values denote a 10% to

20% increase in soil strength per log cycle increase in strain rate and lie within the

commonly reported range (e.g. Vaid & Campenella 1977, Graham et al. 1983, Lefebvre

& Leboeuf 1987). Back calculated values of the strain rate parameters from dynamic

installed anchor experiments in the centrifuge (e.g. Richardson et al. 2006, Richardson,

- 62 -

2008, O‟Loughlin et al. 2009) report higher values for λ and β that lie in the range λ =

0.19–1 and β = 0.06–0.17. The wide range of λ and β values may be (at least in part)

attributed to the order of magnitude difference between the strain rate, and the

reference strain rate, ref . The maximum (which is proportional to v/D) associated

with the free-falling experiments in the centrifuge can be as high as 4250 s-1 (see Figure

2.17) or five to seven orders of magnitude greater than ref for a standard penetration

rate of 0.2 s-1 and up to 10 orders of magnitude greater than the nominal strain rate

associated with laboratory triaxial compression tests and shear vane tests which lie in

the range 3 x 10-6 to 2 x 10-3 s-1 (Einav & Randolph 2006). Therefore it is difficult to

deduce strain rate parameters from laboratory analysis to be adopted in free-fall

experiments in the centrifuge or the field. Similarly, caution is advised when comparing

the strain rates in the centrifuge to the field tests. The strain rates (proportional to v/D)

in the centrifuge are higher because the impact velocities are similar but the diameter is

reduced due to the scaling factor.

2.6.3.2. Dependency of rate parameters on penetrometer geometry

The effects of strain rate dependency on full-flow penetrometer geometry have been

investigated numerically and experimentally. Higher rate parameters have been obtained

for a ball penetrometer than a T-bar (Einav & Randolph 2006, Zhou & Randolph 2007,

Lehane et al. 2009). However, these results were not verified experimentally by Chung

& Randolph (2004). Similarly, for the thin-shafted STING (Figure 2.4b) negligible tip

shape effects were found on the rate dependence (Hurst & Murdoch 1991, Mulhearn

2003).

For full-shafted FFP tests, higher rate dependency has been reported for the shaft

resistance compared to tip resistance (Dayal et al. 1975, Steiner et al. 2014, Chow et al

- 63 -

2014). The higher dependency is believed to be due to the pattern of shearing which is

concentrated in shear bands, and has been accounted for by enhancing the power strain

rate law (Equation 2.17) using a multiplicative term (Zhu & Randolph 2011, Einav &

Randolph 2006). The rate dependency of the shaft is approximately 19–23% higher than

the tip (Chow et al 2014).

2.6.3.3. Dependency of rate parameters on material properties

A large number of soil properties including plasticity, moisture content, stress history,

anisotropy and soil sensitivity has been found to influence strain rate dependency. Soil

plasticity has been recognised to influence the strain rate effect since the 1960s. Higher

rate effects have been observed with higher soil plasticity from triaxial tests (e.g.

Gibson & Coyle 1968, Nakase & Kamei 1986, Diaz-Rodriguez & Martinez-Vasquez

2005) as well as vane studies (e.g. Bjerrum 1973, Peuchen & Mayne 2007, Schlue et al.

2010). Mixed results have been reported on the soil moisture content with most

experiments showing that a reduced moisture content results in lower strain rate effects

(e.g. Gibson & Coyle 1968, Dayal 1974, Brown & Hyde 2008b, Abelev & Valent 2009,

Schlue et al. 2010). However, contradictory results were observed in soils with high

moisture contents, above the liquid limit (LL) from submarine landslide studies (Jeong

et al. 2009) and from vane, viscometer, T-bar and ball penetrometer studies in kaolin

clay (Boukpeti et al. 2012). Therefore, the rate dependency on moisture content or soil

strength may be restricted to soils with moisture contents lower than the liquid limit.

Stress history has been investigated by Lehane et al. (2009) through penetration tests

into soil with an over consolidation ratio (OCR) of 1, 2 and 5. The results indicated that

rate dependence increased with higher OCR. However, reduced rate effects with

increasing OCR have also been observed by Balderas Meca (2004) for relatively high

strain rate tests (up to 60%/min) and negligible findings of OCR on rate dependency

- 64 -

have also been reported in the literature (e.g. Graham et al. 1983, Lunne & Andersen

2007). The apparent conflicting trends surrounding the effect of OCR is likely to be a

consequence of the lack of experimental data at high OCR and also the possible

variation of the influence of OCR at different stain rate ranges. Laboratory tests have

shown that soil anisotropy plays a role in strain rate dependency where higher rate

dependency has been reported in triaxial extension tests than in triaxial compression

tests (Nakase & Kamei 1986, Zhu & Yin 2000). Higher rate effects have been observed

in remoulded clay compared to intact clay in field T-bar and ball penetrometer tests in

soft clay (Yafrate & Dejong 2007).

2.6.4. Combined fluid mechanics and soil mechanics framework

The soil mechanics power strain rate law (Equation 2.17) and the fluid mechanics

Herschel Bulkley model (Equation 2.11) both capture the non-linear variation of

mobilised shear strength with strain rate. However, the two types of formulation differ

by the fact that they link the mobilised shear strength to two separate material

parameters. In order to assess the dynamic resistant force (Fresist) that occurs during a

dynamic event, the current geotechnical standard is to combine the two components

comprising the fluid dynamics component - the drag force (Equation 2.9) and the

geotechnical component – the strain rate dependent bearing force (i.e. Fresist = Fd + Fb).

This approach has been widely adopted in experiments involving assessment of the

dynamic impact on pipelines (e.g. Boukpeti et al. 2012, Randolph & White 2012, Sahdi

et al. 2014):

AsNAvCF opucDresist ,2

21

2.18

where A is the projected area and Nc is the normalised bearing capacity factor.

- 65 -

Sahdi et al. (2014) highlighted the importance of considering both terms in Equation

2.18 in the analysis a submarine slide on an offshore pipeline. The analysis

demonstrated that when either the fluid dynamics component or the geotechnical

component is considered separately, they both fail to capture the flow behaviour over

the large range of flow regimes. If the geotechnical component (the second term on the

right-hand side of Equation 2.18) is neglected then the resistant force is linked to the

material density via a CD. This masks the influence of strength which is the main

resistant component in the creeping flow region (low Re) where CD is more

appropriately a function of the soil strength and not density alone (Deglo De Besses et

al. 2003, Zhu & Randolph 2011). If the drag force (the first term on the right-hand side

of Equation 2.18) is ignored then the forces due to inertia effects, which are often the

dominant resistant forces, particularly for bluff bodies during the first portion of

embedment (O‟Loughlin et al. 2013) are neglected.

2.6.5. Unified framework

A hybrid approach has been proposed, unifying the soil mechanics and the fluid

mechanics frameworks. The approach proposes that both the drag force and the

geotechnical bearing force is a function of the non-Newtonian Reynolds number - Renon-

Newtonian. In order to unite the fluid mechanics and geotechnical frameworks, it is

assumed that the mobilised shear stress, τ (Equation 2.11), is equal to the geotechnical

rate-enhanced shear strength, su,op (see Equation 2.15). The hybrid relationship

superposes separate drag and bearing components and relates both resistant components

shown in Equation 2.18 to a single bearing capacity factor, N (Sahdi et al. 2014):

ANsF opuresist , 2.19

- 66 -

where, Fresist is the dynamic resistant force acting on the pipeline, shown in Equation

2.18, and A is the area. The formula has been calibrated for horizontal slide tests in a

flume (Zakeri et al. 2008, Zakeri et al. 2011) or through horizontal translation tests of a

pipe in a stationary soil in a geotechnical centrifuge (Sahdi et al. 2014). Both sets of

experiments are analogous to a horizontally moving underwater landslide impacting a

suspended pipeline. The results shown in Figure 2.18, plot the normalised bearing

capacity factor (N) against Renon-Newtonian.

Figure 2.18. Variation of normalised lateral pressure on a pipe with non-

Newtonian Reynolds number

- 67 -

Figure 2.18 reveals a number of key findings that may have implications for a deeply-

embedded FFP: N is constant for static and dynamic penetration up to a threshold Renon-

Newtonian, above which, drag forces dominate the value of N increases linearly. This

assumes that an identical failure mechanism for static and dynamic resistance is

effective for Renon-Newtonian < ~ 3. For Renon-Newtonian > ~ 3 the drag term increases linearly

with Renon-Newtonian. In the pipeline analysis this could be accounted for by using a

constant drag coefficient CD. This supports the approach used in other studies (e.g.

Boukpeti et al. 2012, Randolph & White 2012) where a constant CD, independent of

viscous effects has been assumed.

2.6.6. Analytical modelling

A relatively large amount of analytical and semi-empirical models have been produced

in order to calculate the final embedment depth of a FFP. In comparison, a reduced

amount of analytical models have been produced in order to interpret the su profile using

a full-shafted FFP (e.g. Dayal 1974, Elsworth & Lee 2005 and Mulukutla 2009) and a

thin-shafted FFP (e.g. Chow 2012). These models are similar to the original analytical

model proposed by Schmid (1969) and Migliore & Lee (1971) and are based on

Newton‟s second law of motion and the forces acting on the projectile. Typically, the

equation of motion for a full-shafted FFP (Figure 2.12a) falling solely due to its self-

weight is expressed as:

dfrictbs FFFWdt

zdm 2

2

2.20

where m is the projectile mass, z is the projectile tip embedment, t is the time, Ws is the

submerged weight of the FFP (in water or soil) and Fb is the bearing resistance, Fd is the

drag force (Section 2.6.1) and Ffrict is the side frictional resistance. For a thin-shafted

- 68 -

FFP the equation of motion (Equation 2.20) is simplified by removing Ffrict (Chow

2013). A number of small changes have been proposed to Equation 2.20 including the

inclusion of an added mass force (Fernedes et al. 2006, Kunitaki 2008) and also a soil

buoyancy force (Aubeny & Shi 2006, O‟Loughlin et al. 2013b).

The bearing force is expressed as:

AsNF tipucb , 2.21

where Nc is the normalised bearing capacity factor for the projectile tip, su,tip is the

undrained shear strength at the projectile tip and A is the projected area.

For a full-shafted FFP, the side friction force accounts for the friction between the FFP

shaft and the soil, which is usually fully remoulded and has a su close to the fully-

remoulded shear strength, su,rem. Therefore, the friction force is usually reduced by the

interface friction ratio parameter, αs (see Equation 2.23, Anderson et al. 2005):

susfrict AsF 2.22

Where su is the undrained shear strength averaged over the contact area, As and α is the

friction ratio parameter is expressed as a function of the soil sensitivity, St:

u

remu

ts s

sS

,1 2.23

The assessment of Nc for full-shafted FFPs commonly utilises the theoretical solutions

for a CPT with most studies adopting an Nc in the range 12–14. For example, a Nc value

of 12 has been adopted for an XDP (Beard 1977, Bowman et al. 1995) and a DPA

- 69 -

(Richardson, 2008, O‟Loughlin et al. 2009) whereas Stenier et al. (2012) estimated Nc =

14 for a CPT-Lance. Higher values have also been reported by Gilbert et al. (2008) who

estimated Nc = 17 for a torpedo anchor and Freeman & Schuttenhelm (1990) estimated

Nc = 15 for an Ocean Waste Carrier. The large variation is to be expected because the

range of Nc for a CPT varies from 10–30 for fissured clays (Powell & Quarterman

1988). In addition, Nc has been reported to vary depending on penetration depth (Chung

& Randolph 2004, Long & Gudjonsson 2004). Potential reasons include the dependency

on secondary soil characteristics as described in Section 2.2.3.

Thin-shafted FFP such as the STING (Figure 2.4b) have utilised a lower value of Nc =

10 (Mulhearn et al. 1998). This value is somewhere in between the value of a surface

foundation, ~ 5 (Kusakabe et al. 1986, Skempton 1951) and a fully buried circular plate

penetrometer which range between 12.42 (Martin & Randolph 2001) and 13.90 (Lu et

al. 2001). The range of Nc is due to the soil failure mechanism which transitions from a

shallow mechanism to a deep failure mechanism when soil flows over the top on the

penetrometer.

2.6.7. Numerical analysis

A relatively large amount of numerical studies have been carried out to on FFPs,

primarily to ascertain the final embedment depth after dynamic penetration. These

studies provide some insight into the complex soil-structure interaction that occurs at

elevated velocities. For example, Einav et al. (2004) conducted large deformation finite

element (LDFE) to predict the embedment depth of a Deep Penetrating Anchor (DPA)

(see Figure 2.11a). An important observation in the numerical analysis showed that after

the anchor stopped penetrating, negative excess pore pressures were predicted towards

the end of the anchor (near the flukes) which could be related to the flow of material

directly in the wake of the advancing anchor.

- 70 -

Two recent LDFE methods known as the Arbitrary Lagrangian-Eulerian (ALE) method

and Coupled Eulerian-Lagrangian (CEL) method have been reported in the literature.

The ALE method has been used to model the dynamic penetration of the XBP shown in

Figure 2.5d (Aubeny & Shi 2006), STING shown in Figure 2.4b (Abelev et al. 2009b),

CPT (Nazem et al. 2012) and Torpedo anchor shown in Figure 2.11d (e.g. Carter et al.

2013, Sabetamal et al. 2013). The semi-logarithmic rate formula (Equation 2.16) has

been adopted in these studies; however the drag (Equation 2.9) has not been considered

which, in some cases has led to an overestimation of shear strength in the near surface

soil (Abelev et al. 2009b).

LDFE analyses using the CEL approach have been conducted on Torpedo anchors by

Hossain et al. (2013). The analyses explored a range of parameters including anchor

diameter, tip angle, impact velocity and soil strength. In this study two interesting

aspects of the soil flow mechanism were identified: (a) downward soil movement

(possibly indicating the presence of wall failure described in Section 2.2.6.2) and (b)

mobilisation of an end bearing mechanism (see Figure 2.12a) at the base of the anchor.

2.7. Summary

In the past two decades, offshore in situ testing has dramatically increased. This is

usually carried out with an underwater penetration rig and is associated with high costs,

especially in deep water. As an alternative, a number of full-shafted FFPs have been

proposed as a cost-effective and rapid form of assessing the undrained shear strength.

The viability and functionality of these FFP systems have been primarily investigated

using full scale field trials in a large range of water depths up to 5000 m.

The uptake of FFPs as a site investigation tool has been stagnated, this is mainly due to

the interpretation of su. Most FFP designs are similar to a CPT which are full-shafted

- 71 -

penetrometers with a commonly-used a 60° conical tip. Typically, It is challenging to

estimate the su profile with a conical-tipped FFP due to the difficulty in establishing an

appropriate bearing capacity factor which can vary ± 40 per cent depending on the soil

sensitivity (Lunne et al. 1985, 2001). This variation is significantly reduced when a full-

flow penetrometer is considered, where the potential range of the bearing capacity

factor (due to surface roughness) is less than ± 13 percent for a T-bar (Randolph &

Houlsby, 1984) and ± 16 per cent for a ball (Einav & Randolph, 2005). Therefore, there

is an opportunity to conduct an experimental study on a new FFP which combines the

advantages of full-flow penetrometers with the benefits associated with traditional

FFPs.

- 72 -

CHAPTER 3. DYNAMIC PENETRATION OF A

SPHERE IN CLAY

3.1. Abstract

This paper describes field tests in which a 250 mm steel sphere was allowed to free-fall

through water from drop heights of up to 2 m and dynamically embed the soft clay

underlying the water. Instrumentation housed within the sphere measured accelerations

in three orthogonal axes and rates of rotation about the same three axes. These data were

used to calculate velocities and displacements of the sphere during free-fall in water and

embedment in soil. Reasonable agreement was obtained between the measured velocity

profiles and velocity profiles predicted using a simple approach based on strain rate

dependent shearing resistance and fluid mechanics drag resistance.

In the context of the thesis, this chapter introduces the Instrumented Free-Fall Sphere

(IFFS), which is a new free-fall penetrometer design compared to the slender free-

falling penetrometers described in the literature review (Chapter 2). The chapter

describes the field testing method that is also used in the field tests described in

Chapter 7. Example field data are provided and compared with a theoretical model

that describes the motion in soil. This model is extended further in Chapters 6 and 7.

Morton, J. P. & O‟Loughlin, C. D., 2012. Dynamic penetration of a sphere in clay.

Proceedings of the 7th International Conference on Offshore Site Investigation and

Geotechnics, London, UK, pp. 223–230.

- 73 -

3.2. Introduction

Understanding the processes associated with dynamic penetration of rigid bodies from

water into soft soil is challenging. This is principally because of:

An ill-defined transition from water to soil and selection of appropriate material

responses in this zone,

extreme strain rate dependency at high penetration velocities, and

uncertainties regarding the soil-structure interface behaviour for complex

geometries.

The problem has a number of applications including installation of dynamically

installed anchors, free-fall gravity core samplers and in-situ characterisation tools.

Previous work in this arena include centrifuge studies reported by Poorooshasb and

James (1989), Richardson et al. (2006), O‟Loughlin et al. (2004b, 2009), field tests

reported by Freeman et al. (1984) Lieng et al. (2010), and numerical studies reported by

Einav et al. (2004) Nazem & Carter (2010), Raie & Tassoulas (2006). In these studies

the geometry of the rigid body tends to be rather complex, to the extent that a number of

simplifying assumptions are required in order to address the problem. In this paper the

geometry is simplified to a sphere, for which the soil mechanics is quite well behaved

(Randolph et al. 2000), permitting a more rigorous assessment of dynamic penetration

effects. Data from lake tests in which a substantially solid steel sphere dynamically

embedded very soft clay after freefall in water are presented. These data are then used to

validate an embedment model based on strain rate dependent shearing resistance and

fluid mechanics drag resistance.

- 74 -

3.3. Site description and soil properties

3.3.1. Site location and description

Tests were conducted in Lower Lough Erne, Northern Ireland (see Figure 3.1). The lake

is the third largest lake in Britain and Ireland with an area of 109 km2 and is part of the

complex Erne system, with a catchment of 4212 km2.

Figure 3.1. Site location and bathymetric map of Lower Lough Erne (after

Lafferty et al. 2006)

The lower lake, which is situated in County Fermanagh, is located in a deep glacial

trough. The lakebed is largely composed of fine-grained sediments and limestone debris

of boulder clay (Lafferty et al. 2006). At the test site the lakebed is extremely soft to

depths of at least 8 m (Colreavy et al. 2012) and the superficial deposits are principally

underlain by Carboniferous Limestone (Gibson, 1998).

3.3.2. Soil classification

A number of bulk excavated samples have been obtained from the test site from depths

up to 2.5 m and some preliminary classification tests have been carried out. Initial

results are outside the range normally expected for soft clays. Natural moisture contents

are high throughout with values in the range of 270–520%. The Atterberg limits are also

- 75 -

high with plastic limits of 130–180% and liquid limits of 250–315%. Plasticity index

values are in the range of 120–175%. In all cases, the moisture content was

approximately 1.5 times the liquid limit. Over the depth of interest, the unit weight is

constant with depth and only marginally higher than that of water at 10.5 kN/m3.

Scanning electron microscopy images of oven dried samples of the clay revealed

significant amounts of diatoms and other unidentified microfossils (Colreavy et al.

2012). The presence of these microfossils is known to influence soil behaviour, and

contributes to high liquid limits, high porosity, and unusual compressibility (Mitchell &

Soga, 2005). Similar unusual characteristics and index properties have also been

reported for Mexico City clay, which also contains significant amounts of microfossils

(Mesri et al. 1975, Diaz-Rodriguez et al. 1998).

3.3.3. Shear strength profiles

(Colreavy et al. 2012) present penetration data from T-bar and piezoball tests conducted

at the testing site to depths of up to 8m. The penetration tests were performed using a

light weight cone penetrometer testing rig mounted on a floating pontoon. All

penetration tests were conducted at the standard penetration rate of 20 mm/s in an

attempt to ensure undrained conditions. Typical undrained shear strength profiles are

provided on Figure 3.2 for the upper 2 m as this is the depth of interest for the dynamic

penetration tests. The data on Figure 3.2 have been interpreted from the net penetration

data reported by Colreavy et al. (2012) using the commonly adopted resistance factor N

= 10.5 (DeJong et al. 2004, Chung & Randolph, 2004). These strength profiles represent

an undrained shear strength gradient, k = 1.4 kPa/m which is in the range considered

typical for deep water sediments (Randolph, 2004).

- 76 -

Figure 3.2. Typical undrained shear strength profiles at the test site

3.4. Test equipment and testing procedures

3.4.1. Instrumented free-fall sphere

The custom-made Instrumented Free-Fall Sphere (IFFS, Figure 3.3a and Figure 3.3c) is

250 mm in diameter and consists of two hemispheres that are bolted together with an

internal vertically orientated cylindrical void to accommodate instrumentation and a

motion logger. This motion logger (Figure 3.3b) is housed inside the vertically-

orientated void and is protected from water ingress by an O-ring seal. The sphere and

data logger weighs 620N in air and has a submerged weight of 549N in water. Dynaline

Max® 12 mm diameter rope was used for deploying the IFFS and recovering it from the

lakebed after penetration.

- 77 -

Figure 3.3. (a) sphere separated to show internally housed data logger housed, (b)

motion logger and underwater housing, (c) sphere suspended over the water prior

to a drop

3.4.2. Motion logger

The motion logger (Figure 3.3b) was designed to capture the motion history of free-fall

projectiles for periods up to four hours. The motion logger is mounted in an underwater

housing 185 mm long and 42 mm in diameter that fits securely within the sphere‟s

internal hollow chamber. Motion is measured using a 3-component 13-bit accelerometer

(ADXL 345) and a 3-axis 16-bit gyro sensor (ITG 3200) logged by an ARM MBED

logger to a 2 GB micro SD card within the logger, sampling data at up to 800 samples

per second. These sensors measure accelerations on three orthogonal axes and rates of

rotation about the same three orthogonal axes. When the instrumentation housing is

located within the sphere, the z axis of the accelerometer is vertical (parallel to gravity)

and the x and y axes are horizontal (perpendicular to gravity). The accelerometer has a

resolution of 0.04 mm/s2 and detects acceleration up to ± 16g (± ~157 m/s2). The gyro

sensor has a resolution of ± 0.07°/s with a full range of ±2000°/s. Integrating rotation

rates (measured using the three component gyro) allowed tilt to be measured whilst the

- 78 -

IFFS was accelerating in water or decelerating in soil. These measurements were used

to adjust the accelerations measured on the z axis of the accelerometer so that they

correspond to the accelerations in the vertical plane.

3.4.3. Field testing procedure

The sphere was deployed from various drop heights so as to examine the embedment

response of the sphere over a range of impact velocities. Depending on the required

release height above the lakebed, the experiments were conducted either from a jetty, a

floating pontoon or a self-propelled hopper barge. Care was taken to ensure that each

installation site was at least 2 m (8 sphere diameters) from existing test sites. In each

test a portable crane (for the jetty and pontoon tests) or a fixed crane (for the barge tests)

was used to suspend the IFFS at the pre-selected release height, and release was

achieved using a quick release shackle. The embedment depth of the IFFS was

measured by lowering a submersible camera positioned on a weighted platform so as to

focus on known markings on the retrieval rope.

3.5. Test results and analysis

3.5.1. Acceleration profile

A typical acceleration trace, calculated from the horizontally-orientated x and y axes

and the vertically-orientated z axis is shown in Figure 3.4. The z axis acceleration trace

changes abruptly at 0.3s, from 0 m/s2 when hanging vertically in water to ~10 m/s2 after

the quick release shackle was pulled and the IFFS started to free-fall. During free-fall,

the net acceleration decreases as drag on the sphere increases with increasing velocity.

Impact at the lakebed is apparent at ~0.6s, when the sphere decelerates rapidly and then

comes to rest at ~1.4s before rebounding slightly. This rebound is thought to be due to

the stored elastic energy in the clay. Similar results have been reported in free-fall

experiments (e.g. Dayal & Allen, 1973, Chow & Airey, 2010).

- 79 -

The horizontally-orientated y axis accelerometer trace indicates that the sphere rotated

during penetration in soil as the y axis trace does not return to the same position as

before the test (when hanging vertically in water, ~0 m/s2). Rotation derived from the

horizontally-orientated x and y axes of the tri-axis accelerometer is compared with the

rotation derived from the x and y axes of the gyro sensor in Figure 3.4. During the test

the angle of rotation was less than 3° in the x axis, whereas the sphere rotated steadily to

15° in the y axis. This rotation highlights the need to adjust the accelerations measured

on the z axis of the accelerometer during free-fall in water and penetration of the

lakebed.

- 80 -

Figure 3.4. (a) x, y and z axis acceleration traces from a typical test, (b) x and y

axis rotation traces from the same test

- 81 -

3.5.2. Acceleration profile interpretation

Figure 3.5 shows the velocity and displacement of the IFFS which was obtained by

numerically integrating the z axis acceleration (once for velocity and twice for

displacement). As discussed earlier, velocities and displacements determined in this

manner are only representative of velocities and displacements in the vertical direction

when the z axis accelerations are adjusted for tilt. This adjustment is included as a pre-

processing step before numerical integration. Figure 3.5 shows velocities and

displacements that were calculated from the z axis acceleration profile shown in Figure

3.4.

Figure 3.5. z axis acceleration trace for the test shown on Figure 3.4 together with

corresponding velocity and displacement traces

- 82 -

3.5.3. Velocity and embedment depth profile

Figure 3.6 shows the velocity–embedment depth profiles for a 0.5 m and a 1 m release

height above the lakebed. The profiles show the velocity increasing from 0 m/s at the

release height to ~3 m/s and ~4 m/s for the 0.5 m and 1 m drop heights respectively. As

is to be expected for low drop heights (≤ 4 sphere diameters), the velocity does not

reach terminal velocity in water and actually continues to increase during initial

penetration of the soil. This is due to the very low resistance afforded by the very soft

sediments near the lakebed surface. At some depth the resistance available from the soil

exceeds the submerged weight of the sphere and the sphere begins to decelerate. Similar

profiles have been reported by O‟Loughlin et al. (2009) from centrifuge tests on

dynamically installed anchors. Ultimately the sphere comes to rest in the soil as

indicated by zero velocity (in the lakebed) on. This eventual embedment depth is 1.15 m

and 1.35 m for the 0.5 m and 1.0 m drop heights and is in agreement with independent

direct sight measurements using the underwater camera (see Figure 3.7). The

availability of continuous velocity profiles such as that on Figure 3.6 allows embedment

prediction models to be verified and calibrated. The second half of this paper describes

such an embedment model and uses the data reported here to assess its performance.

3.6. Embedment depth prediction

The embedment of bodies penetrating soil after free-fall in water may be quantified by

considering Newton‟s second law of motion and the forces acting on the body during

penetration. Several studies (e.g. True, 1976, Aubeny & Shi, 2006, Audibert et al. 2006,

O‟Loughlin et al. 2004a) have adopted such an approach, with variations on the

inclusion and formulation of the various forces acting on the body during penetration. A

similar approach is adopted here with the forces acting on the body during penetration

shown by Figure 3.8, leading to a governing equation of:

- 83 -

dbfs FFRWdt

zdm 2

2

3.1

where m is the sphere mass, z is the penetration depth below the lakebed, t is the time

after impact with the soil, Ws is the submerged weight of the sphere in clay, Rf is a

strain rate function, Fb is the bearing resistance and Fd is the drag resistance. The

inclusion of the strain rate function, Rf, is justified as the shear strength of clays is well

known to be a function of strain rate (Casagrande & Wilson, 1951, Graham et al. 1983).

Figure 3.6. Velocity profiles in water and soil for release heights of 0.5 m and 1 m

- 84 -

Figure 3.7. Measuring the embedment depth using markers on the retrieval rope

taken with the underwater camera

Figure 3.8. Forces acting on the IFFS during penetration in soil

Fb is formulated as:

pucb AsNF 3.2

- 85 -

where Nc is the tip-bearing capacity factor for a sphere taken as 10.5 (DeJong et al.

2004, Chung & Randolph 2004), Ap is the projected area of the sphere, and su is the

undrained shear strength at the sphere tip. The dependence of shearing resistance on

strain rate is accounted for in Equation 3.1 by scaling the bearing resistance using a

power strain rate law (Biscontin & Pestana, 2001), expressed as:

β

reff γ

γR

3.3

where β is the strain rate parameter, γ is the strain rate and refγ is the reference strain rate

associated with the reference value of undrained shear strength. In the field test the

undrained shear strength was measured using a 113 mm piezoball penetrating at 20

mm/s, such that v/D = 0.2s-1. As similar amounts of remoulding are to be expected

during dynamic penetration of the sphere as in static penetration of the piezoball, it

becomes reasonable to replace the strain rate terms in Equation 3.3 with v/D to give:

β

reffR

DvDv

3.4

where (v/D)ref = 0.2s-1. Back-analysis of dynamically installed anchor data from

centrifuge tests indicates that the strain rate parameter β is in the range 0.05 to 0.15 as

vav/D increases from 500 to 4250s-1 (O‟Loughlin et al. 2009). Values of vav/D in the

field tests reported here are typically equal to 20. As such, an appropriate β for use in

Equation 3.4 should be at the lower extreme of the range quoted above, and would

therefore be similar to parameters deduced from variable rate penetrometer tests (e.g. β

= 0.05, Chung et al. 2006; β = 0.06, Lehane et al. 2009). The model can be further

- 86 -

refined by accounting for the transition from a shallow to a deep failure mechanism

during shallow penetration and also for the buoyancy effects as the ball transitions from

water into soil (White et al. 2010). However these refinements are not warranted here as

the unit weight of the soil is close to that of water and the ball typically embeds by

several diameters.

The inclusion of the fluid mechanics drag term, Fd, in Equation 3.1 is warranted in view

of the very soft viscous clay that is often encountered at the clay surface (O‟Loughlin et

al. 2009). The fluid drag is formulated as:

2

21 vACF pDd 3.5

where CD is the drag coefficient, dictated largely by object geometry (blunt objects

exhibit a large CD and streamlined bodies exhibit smaller CD), ρ is the density of the

soil, Ap is the projected area and v is the velocity.

An appropriate CD for the sphere can be determined by considering Equation 3.1 during

freefall in water rather than embedment in soil. As such the bearing resistance term in

Equation 3.1 becomes zero, the submerged weight, Ws, (in Equation 3.1) and the

density, ρ, relate to water rather than soil. The governing equation in water then

becomes:

2

22

21

dtzdmvACW pds 3.6

The theoretical velocity profile for the sphere free falling in water can then be

constructed from a finite difference approximation of Equation 3.6. These theoretical

profiles are shown on Figure 3.9 together with experimental velocity profiles for drop

heights of 0.5m, 1 m and 2 m. Lower and upper bounds to the experimental data are

obtained using CD = 0.3 and CD = 0.4, with an average CD = 0.35 providing the best

- 87 -

overall fit to the measurements. This is in agreement with (Richardson, 2008) who

determined CD = 0.35 for a sphere using computational fluid dynamics and is very close

to the values shown on Figure 2.14, especially for Re > 2 x 105.

Figure 3.9. Measured and theoretical velocity profiles of the sphere free-falling in

water

The experimental velocity profiles of the sphere penetrating the lakebed are shown in

Figure 3.10. The velocity profiles correspond to a 0m, 0.5m, 1 m and 2 m release height

above the lakebed. Also shown on Figure 3.10 are the theoretical velocity profiles that

were constructed using a finite difference approximation of Equation 3.1 and the soil

and model parameters introduced and discussed earlier in the paper (i.e. Nc = 10.5, β =

0.05, CD = 0.35).

Both the measurements and the model show increasing sphere velocity during initial

penetration in the soil. This is to be expected as for these relatively low drop heights the

- 88 -

sphere will not have reached terminal velocity. The very low penetration resistance

afforded by the soil at shallow depth is lower than the submerged weight and the sphere

velocity continues to increase in the soil. At some depth the resistance from the soil

equals and then exceeds the submerged weight of the sphere and the velocity begins to

reduce. The depth at which this occurs is seen to reduce with increasing impact velocity,

as the difference between the terminal velocity and the impact velocity reduces. Similar

findings have been reported by (O‟Loughlin et al. 2009) from centrifuge tests on

dynamically installed anchors over a range of impact velocities.

The agreement between the model and the measured velocity profiles is quite good,

with the exception of the test where the IFFS was released from the mudline (impact

velocity = 0 m/s), and to a lesser extent the test released from 2 m above the lakebed

(impact velocity ≈ 5 m/s). This is considered to be due to the assumed linear shear

strength profile, k = 1.4 kPa/m which underestimates the actual undrained shear strength

between 0 and 0.8 m and slightly overestimates the actual undrained shear strength

between 0.8 and 2.0m.

- 89 -

Figure 3.10. Predicted and measured velocity profiles of the sphere penetrating the

lakebed

3.7. Conclusions

This paper has presented field tests undertaken to evaluate the behaviour of a simple

geometry dynamically penetrating soft clay. The tests, which were undertaken in a lake

in water depths up to 5m, involved the release of a substantially solid steel sphere, 250

mm in diameter, through drop heights of up to 2m. Impact velocities (at the mudline) of

up to 5.1 m/s were obtained in the field tests, resulting in embedment (to the base of the

sphere) of 1.52 m, equivalent to over 6 sphere diameters. Reasonable agreement was

obtained between the measured velocity profiles and velocity profiles predicted using a

simple approach based on strain rate dependent shearing resistance (β = 0.05) and fluid

mechanics drag resistance (CD = 0.35).

- 90 -

CHAPTER 4. STRENGTH ASSESSMENT DURING

SHALLOW PENETRATION OF A SPHERE IN CLAY

4.1. Abstract

Strength interpretation from the measured penetration resistance of full-flow

penetrometers, such as the T-bar and ball, is generally based on a constant bearing

capacity factor associated with a deep flow-round mechanism. This approach may

underestimate the strength of near-surface sediments, which is becoming increasingly

important for the design of offshore infrastructure such as pipelines, steel catenary risers

The theoretical model considered in Chapter 3 makes the limiting assumption that

shallow penetration effects may be ignored. This chapter avoids this limitation by

proposing a shallow penetration framework that more accurately determines the

penetration resistance acting on a ball penetrometer whether pushed in or free-falling

during shallow embedment. The framework is used in Chapters 7 and 8 where it is

applied to field and centrifuge experiments for the analysis of push-in and free-falling

ball penetrometer tests.

Morton, J. P., O‟Loughlin, C. D. & White. D. J., 2014. Strength assessment during

shallow penetration of a sphere in clay. Géotechnique Letters 4 (October-December),

pp. 262–266.

- 91 -

and mudmats. This paper describes a series of centrifuge experiments designed to

capture the change in the capacity factor of a ball penetrometer during shallow

penetration. A rigorous consideration of soil buoyancy is provided. This is an important

consideration in soils with a higher strength to self-weight ratio because a cavity is

formed by the passage of the ball and remains open to greater depths. The depth at

which a full-flow mechanism develops is related to the dimensionless strength ratio,

expressed as the ratio of the undrained shear strength to the effective unit weight and

penetrometer diameter. This observation forms the basis for proposed formulations that

describe the evolution of the bearing capacity factor with depth for different

dimensionless strength ratios. These formulations can be used to determine more

accurately the undrained shear strength of near surface soil over the range of

dimensionless strength ratios that is of interest to offshore applications.

4.2. Introduction

For a deeply embedded ball penetrometer, a full-flow mechanism that is symmetrical

above and below the ball is operative and the undrained shear strength, su, can be

interpreted from the measured bearing pressure, qm, according to su = qm/Nb-deep.

However, at shallow penetration depths a full-flow mechanism does not develop and su

should be interpreted from the measured penetration resistance using a bearing capacity

factor Nb-shallow < Nb-deep. Adopting an appropriate Nb-shallow, and accounting for its

evolution to Nb-deep with depth, is an important aspect of quantifying su over the upper 1

to 2 metres of the seabed. This is critical for the design of almost all shallowly

embedded offshore infrastructure (Puech et al. 2010) including subsea pipelines, steel

catenary risers and mudmats.

Large deformation finite element (LDFE) analyses on a T-bar (White et al. 2010, Tho et

al. 2012), spudcan (Hossain et al. 2005) and ball penetrometer (Zhou et al. 2013) have

- 92 -

shown that the transition depth from Nb-shallow to Nb-deep is dependent on the

dimensionless strength ratio, su/γ'D, where γ' is the effective unit weight of the soil and

D is the diameter of the penetrometer. Higher strength ratios are associated with a delay

in the transition to a steady Nb-deep. Correlations for the transition depth and Nb-shallow

have been derived and, in the case of the T-bar and ball, a basis for correcting

penetration data within the shallow zone has been proposed (White et al. 2010, Zhou et

al. 2013). However from the perspective of a spherical penetrometer, which is the focus

of this paper, the range of strength ratios previously examined (su/γ'D = 2.95 to 44.25,

Zhou et al. 2013) is narrower than the range that is of practical interest for offshore

problems. For example, a 250 mm diameter freefall spherical penetrometer (for

measuring the strength of the seabed, Morton & O‟Loughlin, 2012) penetrating very

soft soil is associated with low values of su/γ'D approaching 0.1 at one diameter

embedment, whereas an 80 mm piezoball penetrating the seabed with a crust strength of

~20 kPa is associated with high values of su/γ'D, approaching 40.

The motivation for this study was to experimentally capture the variation in Nb-shallow

with depth for a ball penetrometer embedding into clay over a wide range of su/γ'D. The

experimental data are combined with reinterpreted LDFE results (Zhou et al. 2013), and

form the basis of a new correlation which describes the evolution of Nb with depth.

4.3. Experimental details

The problem is addressed through centrifuge tests carried out at 100 g in the University

of Western Australia (UWA) beam centrifuge. The penetration resistance response and

the degree of hole-closure were analysed for nine penetrometer tests using a 11.3 mm

diameter ball with a 4.8 mm diameter shaft, penetrating a kaolin clay sample with a

progressively higher overconsolidation ratio. A constant penetration rate of 1 mm/s was

adopted such that the non-dimensional velocity, vD/cv ~ 130 (where v is the penetration

- 93 -

rate, D is the sphere diameter, and cv is the coefficient of vertical consolidation ~ 2.8

m2/yr (Cocjin et al. 2014) and the response is primarily undrained (House et al. 2001).

A video observed the progressive hole-closure during each test, and provided a means

of determining the depth at which the cavity, formed by the passage of the ball, closed

over. The experimental arrangement is shown in Figure 4.1.

4.4. Experimental procedure

4.4.1. Preparation of clay specimen

The sample was prepared by mixing kaolin powder with water in a vacuum mixer at a

moisture content equal to twice the liquid limit (120%). A drainage sand layer at the

base of the sample allowed two-way drainage during self-weight consolidation in the

centrifuge at 100 g, and vertical drains in the corners of the sample ensured there was no

hydraulic gradient over the height of the sample. A nominal 10 mm layer of free water

was maintained throughout testing.

- 94 -

Figure 4.1. (a) and (b) Experimental arrangement in the beam centrifuge

In order to investigate the range of su/γ'D of interest (spanning two orders of

magnitude), the local su was progressively increased by scraping 20 mm layer of clay

from the surface of the sample between each consecutive penetration test (see Fig 2).

This had the effect of increasing the overconsolidation ratio (OCR) of the clay and

- 95 -

increasing su relative to the (new) sample surface. In order to minimise potential

boundary effects, a minimum of 3 sphere diameters was allowed between each test site

and sample walls.

Figure 4.2. A scraped soil sample before a test

Figure 4.3. Ball penetrometer and cavity after a penetration test

- 96 -

4.5. Theoretical basis for interpretation of measured ball penetration

resistance

As the ball penetrates the soil, the measured bearing pressure, qm, includes the evolving

soil resistance, qs, expressed in terms of Nb, and the resistance due to soil buoyancy, qb.

p

buoybbsm A

FsNq uqq 4.1

where Fbuoy is the soil buoyancy force and Ap is the projected area of the sphere.

If the ball is deeply embedded such that the soil flows around the ball during

penetration, the buoyancy force can be calculated from Archimedes‟ principle, i.e. the

buoyancy force is the volume of the displaced soil multiplied by the effective unit

weight of the soil, Fbuoy = (πD3/6)γ'. However, during initial penetration the soil does

not flow around the ball. Instead, a cavity is created above the ball and the soil that

would have filled this void is instead accommodated by heave at the soil surface.

To capture this heave effect, a simple multiplier can be applied on Fbuoy, as proposed

previously for penetration of cylindrical and spherical geometries (e.g. Merifield et al.

2009, White et al. 2010, Chatterjee et al. 2012, Zhou et al. 2013). An alternative

approach to derive this multiplier directly, is to consider the work required to lift the

soil that is displaced by the incrementally advancing ball. This can be done by assuming

a cavity geometry formed by the advancing ball which is approximated here as an

inverted cone for all considered values of su/γ'D, prompted by camera observations (e.g.

see Figure 4.3) and supported by LDFE simulations (Zhou et al. 2013).

At a ball invert depth, w ≤ 0.5D (Figure 4.4a), all of the soil displaced by the advancing

ball is lifted to the soil surface. The work done then becomes the weight of the displaced

- 97 -

soil multiplied by the distance between the centroidal height of the embedded ball (a

spherical cap) and the soil surface. For 0.5D > w ≤ Hc + D, where Hc is the cavity depth,

only some of the soil displaced by the advancing ball is lifted to the surface. The

remainder fills part of the cavity created by the increment of ball penetration as shown

in Figure 4.4b. In this case, the work done is calculated by adding the gain in potential

energy by these two separate elements of soil, resulting in the profile of soil buoyancy

with embedment given in Figure 4.4c. The analytical closed-form solution to calculate

the soil buoyancy is shown in Appendix 1.

- 98 -

Figure 4.4. Schematic illustration of soil buoyancy due to (a) the sphere and (b) the

sphere and conical cavity (c) buoyancy function for a typical cavity depth

- 99 -

The net penetration resistance, qnet, can be calculated using the measured resistance, qm

from Equation 4.1, by correcting for the unequal pore pressure and overburden pressure

effects due to the shaft behind the ball (Chung & Randolph 2004):

p

sovmnet A

Aα1uσqq 4.2

where σv is the overburden pressure, u0 is the hydrostatic pore pressure, As is the shaft

area, and the parameter α is the net area ratio of the load cell core to the shaft area (α =

0.8 for the tests considered here).

For a deeply embedded ball, qnet, and hence the inferred su, corresponds to the mid-

height of the ball due to the depth symmetry of the flow-round mechanism. For a

shallowly embedded ball, where the full-flow mechanism is not fully developed, the

normalised operative depth, ŵop, is assumed to vary linearly up to the depth where a full

flow-round mechanism occurs, in a similar manner to that proposed for a shallowly

embedded T-bar (White et al. 2010):

0.5DDH

0.5D

wc

op

www 4.3

Omitted from the above theoretical framework, is any consideration of how Nb evolves

during shallow to deep penetration. This has been purposely excluded from the

preceding discussion and will be formulated later in the paper to reflect the

experimental results presented in the following section.

- 100 -

4.6. Results and comparisons

4.6.1. In-flight video camera observations

The camera was synchronised with the data acquisition system, such that visual

observations could be relayed to the measured penetration response. For lower values of

su/γ'D, where the open cavity depth was lower and could be captured by the camera, the

instances when soil flowed over the ball were consistent with the transitional depths

inferred from the penetration profiles shown in Figure 4.5 (and discussed in the

following section). This observation is at variance with LDFE results reported by Zhou

et al. (2013), which indicate that more penetration is required to establish a deep failure

mechanism after full flow of soil over the top of the ball.

Figure 4.5. Comparison of strength profiles from Equation 4.4 and qnet/Nb-deep

During shallow penetration, little or no heave was observed on the soil surface,

- 101 -

particularly for tests with higher values of su/γ'D. This is considered to be due to the

axisymmetric flow mechanism of the ball, which reduces the heave compared with

plane strain flow for a cylindrical T-bar or pipeline (Stanier & White, 2014). In light of

this observation, enhancement of Fbuoy in Equation 4.1 to account for heave of the soil

surface was included using the approach outlined in Section 4.5, and the approach

considered by Merifield et al. (2009), Randolph & White (2008) and Stanier & White

(2014) was not included.

4.6.2. Undrained shear strength profiles

As discussed earlier, the OCR of the sample was incrementally increased by scraping an

additional 20 mm from the sample surface between penetration tests. The strength of the

sample can then be assumed to vary with depth according to the following relationship

proposed by Ladd et al. (1977).

mOCRσsσs

ncv

uvu

4.4

where σ'v is the current vertical effective stress, determined from the γ' profile with

depth and the varying acceleration level within the centrifuge (93 to 105 g over the

depth of penetration) and m is the plastic volumetric strain ratio (Schofield and Wroth,

1968). The normally consolidated undrained strength ratio (su/σ'v)nc = 0.13, as

determined from ball penetrometer tests before scraping the soil surface (i.e. OCR = 1)

using the commonly adopted Nb-deep = 10.5 (Chung & Randolph 2004). The ratio

increases to (su/σ'v)nc = 0.15 if the measured resistance (which ignores shaft effects) is

considered, similar to (su/σ'v)nc = 0.16 reported by Richardson et al. (2009) and Hu et al.

(2014), and equivalent to an undrained strength gradient with prototype depth, k = 1

- 102 -

kPa/m, which is typical for UWA kaolin. The best agreement between Equation 4.4 and

the experimental qnet/Nb-deep profiles in the overconsolidated samples was obtained using

m = 1 (rather than the commonly reported m = 0.8), which reflects the minimal swelling

time permitted between each soil scrape and the subsequent penetration test.

4.6.3. Deep mechanism transition depth

The normalised transition depths, ŵdeep-op, are also shown on Figure 4.5, and were

selected as the depths where the qnet/Nb-deep experimental data were judged to have

reached su predicted using Equation 4.4. The final two profiles do not reach the

predicted su profile due to the proximity of the base drainage sand layer and ŵdeep-op are

approximated in these instances on the basis of the Nb variation with depth, discussed

below. Values of ŵdeep-op are also shown in Figure 4.6 alongside previously reported T-

bar, spudcan and ball data (White et al. 2010, Hossain et al. 2005 and Zhou et al. 2013

respectively), but re-interpreted to account for the definition of operative depth adopted

here. Further reinterpretation of the Zhou et al. (2013) ball data was made to ascertain

ŵdeep-op, assessed as when Nb became effectively constant (to within < 5% of the final

value), rather than reaching the limit, which is difficult to judge and is approached

asymptotically. The experimental ball data reported here, together with the reinterpreted

Zhou et al. (2013) ball data now form a unique relationship between ŵdeep-op and su/γ'D

for a ball penetrometer (where su is the undrained strength at ŵdeep-op), which can be

described using:

fu

cu

opdeep

eDγ's1

adDγ'

sbaw

4.5

where the fitting constants a = 16.3, b = 0.12, c = 1.3, d = 0.52, e = 4.9 and f = 1.5.

- 103 -

Figure 4.6. Effect of strength ratio su/γ'D on transition depth

4.6.4. Shallow bearing capacity factors

Figure 4.7 shows the experimental variation in Nb with depth for each penetration test,

obtained by dividing qnet by su from Equation 4.4. Each Nb profile follows the same

trend, commencing at zero at the soil surface and reaching a steady Nb-deep at the

transition depth, which is entirely dependent on su/γ'D. This variation in Nb with depth

can be fitted using Equations 4.6 and 4.7 (also shown on Figure 4.7).

p

opdeep

opdeepbshallowb w

wNN

4.6

and

- 104 -

0.01u

Dγ's0.49p

4.7

Figure 4.7. Measured variation in normalised bearing factor with normalised

embedment depth and equation fit

4.7. Conclusions

This paper reports on centrifuge tests in which a ball was penetrated into clay under

undrained conditions over su/γ'D = 0.07 to 2.7 (at the transition depth). The depth at

which an open cavity, formed by the passage of the ball, closed over is considered to

indicate the transitional depth, ŵdeep-op, where a full flow-round mechanism develops. A

novel analytical solution for the soil buoyancy in the case of an open conical hole has

been developed. This rigorous approach is necessary to avoid significant errors in the

determination of strength during shallow penetration in soils with a low strength to self-

weight ratio. For instance in a clay with zero mudline strength and an undrained shear

- 105 -

strength ratio, su/σ'v = 0.25, the buoyancy resistance increases to almost 70% of the

geotechnical resistance during shallow penetration, and is independent of the

penetrometer diameter at all penetration depths. Data from the centrifuge experiments,

combined with reinterpreted data from LDFE analyses, show a unique relationship

between ŵdeep-op and su/γ'D examined over the range su/γ'D ≈ 0.1 to 40. Equations that

describe the change in ŵdeep-op and the capacity factor, Nb, with su/γ'D are proposed.

These equations offer a more rigorous and reliable means of assessing soil strength in

the upper few metres of the seabed.

- 106 -

CHAPTER 5. MEMS ACCELEROMETERS FOR

MEASURING DYNAMIC PENETRATION EVENTS IN

GEOTECHNICAL CENTRIFUGE TESTS

5.1. Abstract

Micro-electro mechanical system (MEMS) accelerometers are small, inexpensive

sensors that have only recently been used in geotechnical centrifuge tests. This is unlike

piezoelectric accelerometers, which are by comparison large and expensive but have

been used extensively in geotechnical centrifuge tests over the past couple of decades.

This paper examines the response of a single-axis ±500 g MEMS accelerometer under

both static and dynamic conditions in a centrifuge environment. The potential for

MEMS accelerometers to be used to measure the depth of objects buried in soil is

This chapter describes novel FFP experiments that use accelerometers embedded in a

reduced scale centrifuge model anchor to measure the motion response during free-

fall and embedment in soil. The chapter is a precursor for Chapters 6 and 8 where

further experiments with MEMS accelerometers are undertaken in the field and the

centrifuge. The chapter highlights the potential error associated with FFP tilt which is

investigated in more detail in Chapter 6.

O‟Loughlin, C. D., Gaudin, C., Morton. J. P. & White, D. J., 2014. MEMS

accelerometers for measuring dynamic penetration events in geotechnical centrifuge

tests International Journal of Physical Modelling in Geotechnics, 14(2), pp. 31–39.

- 107 -

examined and the achievable resolution is discussed. Unlike piezoelectric

accelerometers, which only measure changes in acceleration, MEMS accelerometers

can measure both constant and changing accelerations. The merit of this feature is

demonstrated through tests in which MEMS and piezoelectric accelerometers are

embedded within a dynamically installed model anchor. The MEMS accelerometer is

capable of measuring accelerations during both the free-fall phase and the soil

embedment phase, whereas the piezoelectric accelerometer is only able to measure the

changing accelerations that dominate during the soil embedment phase. Velocity

profiles derived from numerical integration of the MEMS accelerations give mudline

anchor velocities that agree with independent measurements and anchor embedment

depths that agree with direct measurements.

5.2. Introduction

Micro-electro mechanical system (MEMS) accelerometers are simple and inexpensive

accelerometers that are commonly used in motion-activated user interfaces (such as

smartphones and game consoles) and protection systems (such as free-fall protection of

hard drives in laptops and airbag deployment in vehicles). MEMS accelerometers 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 spring elements. Accelerations acting on the proof

mass cause it to displace, and plates connected to the proof mass move between the

fixed plates. This displacement causes a differential capacitance that is measured by

integrated electronics and is output as a voltage that is proportional to the acceleration

acting on the proof mass. The operational principle is shown schematically in Figure

5.1. The Analog Devices ADXL001 MEMS accelerometer used in this paper is shown

- 108 -

in Figure 5.2a, where the internal view of the chip shows the MEMS sensing element

surrounded by the integrated electronics. A scanning electron microscope (SEM) image

of the sensing element in Figure 5.2b shows the proof mass, plates, springs and anchor

points. This sensing element measures only 0.5 x 0.5 mm with overall chip dimensions

of 5 x 5 x 2 mm. MEMS accelerometers have been used for full scale geotechnical

applications, including measurement of inclinations in boreholes (Bennett et al. 2009),

measurement of soil displacements associated with rapid uplift of footings (Levy &

Richards, 2012 and measurement of the motion response of dynamically installed

anchors during free-fall in water and penetration in soil (similar to this study)

(O‟Loughlin et al. 2013; Lavieri et al. 2011). The use of MEMS accelerometers in

geotechnical centrifuge modelling has increased in recent years, particularly for

measuring earthquake accelerations (Cilingir & Madabhushi, 2011; Stringer et al.

2010), and more recently for measuring rotations of structures during slow lateral

cycling and after dynamic shaking (Allmond et al. 2014). Bhattacharya et al. (2012) and

Stringer et al. (2010) compared the performance of MEMS accelerometers against

piezoelectric accelerometers, which have been widely used in geotechnical physical

modelling applications over the last two to three decades. These studies highlight the

attractiveness of MEMS accelerometers owing to their low mass, size and cost, but also

show that the MEMS accelerometer produces comparable performance during high-

frequency events and superior performance during low-frequency events.

- 109 -

Figure 5.1. Schematic representation of the operational principle of a MEMS

accelerometer

- 110 -

Figure 5.2. MEMS accelerometer: (a) Analog Devices ADXL001 MEMS

accelerometer with an internal view of the chip showing the MEMS sensor

surrounded by the integrated circuitry and (b) Scanning Electron Microscope

image of the ADXL001 accelerometer showing the proof mass, plates, springs and

anchor points

This paper investigates the performance of MEMS accelerometers in a geotechnical

centrifuge under both static and dynamic conditions, meaning both constant and varying

acceleration. The static tests compare the MEMS acceleration with that derived from the

rotational velocity of the centrifuge, before considering the potential for the MEMS

- 111 -

accelerometer to be used as a depth measurement for objects buried in soil. The

dynamic tests focus on embedment of objects free-falling into soil. First, the MEMS

accelerometer is compared with a piezoelectric accelerometer and second the MEMS

accelerometer is used to derive the motion response of an object free-falling through air

and embedding in soil.

5.3. Static centrifuge tests

The MEMS accelerometer performance was first examined under static conditions in

the beam geotechnical centrifuge at The University of Western Australia. In these initial

tests an ADXL001 ±500 g MEMS accelerometer was embedded within an epoxy-filled

void in a strip of aluminium that was rigidly connected to an actuator in an orientation

such that the axis of the MEMS accelerometer was aligned with the direction of the

gravitational field within the centrifuge. The vertical axis

of the actuator was then moved (at 1g) until the radius of the MEMS accelerometer was

equal to the radius at which the nominal g-level was set (1.54 m). A comparison of the

acceleration level measured by the MEMS accelerometer and the acceleration level

calculated using the measured angular velocity of the centrifuge is provided in Figure

5.3a during centrifuge acceleration from 0 to 100 g and back. The only signal

conditioning applied to the signal is a low pass resistor– capacitor (RC) passive filter

with a 3 dB cut-off frequency of 120 kHz and signal averaging in time blocks of 0.1 s.

A slight difference is observed in the measured quantities before and after spinning, as

the MEMS measures Earth‟s gravity and so outputs 1g, whereas the zero angular

velocity infers zero centripetal acceleration. The inset in Figure 5.3a, during which the

acceleration level is a constant 100g, shows good agreement between the acceleration

levels measured by the MEMS accelerometer and derived from the centrifuge‟s

rotational speed. The minimum resolution of the MEMS accelerometer is evident from

- 112 -

the step-ladder response at constant acceleration level. The minimum resolvable

acceleration resolution, ∆N, is given by:

max2N

VVN n

S

ADC 5.1

where VADC is the full scale input voltage range of the ADC (analogue to digital

converter) in the data acquisition system, VS is the full scale output range of the

accelerometer, n is the bit resolution of the ADC and Nmax is the full scale range of the

accelerometer. For these tests where VADC = 10 V, VS = 2.2 V, n = 16 and Nmax = 1000

g (±500 g accelerometer), ∆N is calculated as 0.07 g or 0.69 m/s2. This resolution can be

increased by using a sensor with a lower rated Nmax.

The data in Figure 5.3a are also shown in Figure 5.3b, but with the MEMS sensor

output plotted against the centrifuge acceleration level as the centrifuge acceleration is

increased from 0 to 100 g and back. Figure 5.3b indicates that the response is linear,

although with an apparently slight hysteresis which is attributable to the time delay in

the reporting of the centrifuge rotational speed to the data acquisition system.

- 113 -

Figure 5.3. Comparison between MEMS acceleration measurements and those

derived from the rotational speed of the centrifuge during spin up to 100 g and

down again: (a) time history, and (b) comparison

(a)

(b)

- 114 -

The vertical axis of the actuator was subsequently moved at 1 mm/s and the radius, R,

derived from the MEMS accelerometer measurements (Equation 5.2) was compared

with the radius established from the vertical displacements of the actuator (measured

using the motor encoder on the vertical axis of the actuator) (Figure 5.4). The agreement

highlights the potential for MEMS accelerometers to be used as a depth indicator in

centrifuge tests. To assess the minimum depth increment that can be detected, the

acceleration level in the centrifuge depends on the radius, R, according to:

by:

RNg 2 5.2

where ω is the angular velocity in rad/s. From Equations 5.1 and 5.2 the minimum

displacement resolution ∆R becomes:

maxmax2 22N

VV

NRN

VVgR n

S

ADCen

S

ADC

5.3

For the tests considered later in the paper, where N = 133.3 at an effective radius, Re =

1.574 m, VADC = 10 V, VS = 2.2 V and Nmax = 1000 g, ∆R is calculated as 0.82 mm.

This resolution will increase as the maximum range of the accelerometer decreases, and

also as the acceleration level in the centrifuge increases. Although this resolution is at

least two orders of magnitude worse than can be achieved using conventional

displacement potentiometers, the small size of the MEMS accelerometer opens up

possibilities for making depth measurements of objects that are buried within a soil

sample, out of reach of a conventional potentiometer. This has particular merit for

freefall penetrometers and dynamically installed anchors, as demonstrated in the

following section.

- 115 -

Figure 5.4. Comparison between position measurements derived from the MEMS

accelerometer and measured using the motor encoder on the vertical axis of the

actuator

5.4. Example application: dynamically installed anchors

Dynamically installed anchors (see Figure 5.5) are torpedo shaped devices used

offshore. They are designed so that, after release from a designated height above the

seafloor, they will penetrate to a target depth in the seabed using the kinetic energy

gained through free-fall. The key challenge in predicting the anchor capacity is the

calculation of the anchor embedment depth and hence the available shear strength in the

vicinity of the embedded anchor. This calculation is complicated by (a) the strain rates

at the anchor–soil interface, which are on the order of 25 s-1, three orders of magnitude

higher than that in a vane test (0.029 s-1; Einav & Randolph, 2006) and seven orders of

magnitude higher than typical laboratory testing rates of , 1%/h (2.8 × 10-6 s-1), and (b)

hydrodynamic effects associated with the very soft viscous clay at shallow penetration

- 116 -

and possible entrainment of a boundary layer of water adjacent to the anchor

(O‟Loughlin et al. 2013). These complexities necessitate careful calibration of anchor

embedment models to ensure that the framework for incorporating drag resistance and

for enhancing the soil strength as a function of strain rate is appropriate. To date it has

been necessary to calibrate such models using known starting and end conditions: a

measured (or calculated) anchor velocity at zero depth (i.e. at the mudline) and zero

anchor velocity at the final (measured) anchor embedment depth (e.g. O‟Loughlin et al.

2013). However, this approach is less than satisfactory as anchor motion data are not

available during dynamic penetration. The following section describes how MEMS

accelerometers embedded within dynamically installed model anchors can produce

these motion data in a centrifuge environment.

Figure 5.5. Dynamically installed anchors: (a) torpedo pile (after Araujo et al.

2004), (b) OMNI-Max anchor (after Shelton, 2007), (c) deep penetrating anchor

(Deep Sea Anchors, www.deepseaanchors.com/News.html)

- 117 -

5.5. Dynamic centrifuge tests

A single-axis ADXL001 ±500 g MEMS accelerometer was embedded in epoxy resin in

a void created in the shaft of a dynamically installed „OMNI-Max‟ anchor, described by

Shelton (2007) and Zimmerman et al. (2009) and tested at UWA within a „Joint

Industry Project‟ (Gaudin et al. 2013). The three sensor wires (power, ground and

signal) were recessed into the anchor shaft and exited the anchor at the top of the shaft

where they joined the anchor mooring line (see Figure 5.6). For comparison, and to

ensure that acceleration could still be measured if the MEMS accelerometer went out of

range during deceleration in the soil, a ±10,000 g piezoelectric accelerometer

(Piezotronics model 305A03) was located at the upper end of the anchor shaft. As

highlighted earlier, piezoelectric accelerometers are generally too large for this

application. In this instance the model anchor was scaled at 1:133.3 so that the shaft

diameter (7.2 mm) was the same as the diameter of the piezoelectric accelerometer

(after machining excess material from the casing and removing the accelerometer cable

connections).

Figure 5.6. Accelerometers installed in a model dynamically installed anchor

- 118 -

The tests were conducted in the UWA beam centrifuge at 133.3 g in both normally

consolidated calcareous silt and normally consolidated kaolin clay. Anchor installation

was achieved in flight by allowing the anchor to fall through a vertical installation guide

from 240 mm (measured to the anchor tip) above the soil sample (see Figure 5.7).

Example acceleration traces from both the MEMS and piezoelectric accelerometers

captured at 50 kHz are provided in Figure 5.8 for a test in the calcareous silt. As with

the static tests, the data were conditioned using a low pass RC passive filter with a 3 dB

cutoff frequency of 120 kHz, but with no time averaging. As piezoelectric

accelerometers only make dynamic measurements, detecting changes in acceleration,

the output from the piezoelectric accelerometer is zero up to the point where the anchor

is released. The delay in the signal returning to zero after the anchor comes to rest is due

to the discharge time of the piezoelectric charge. As MEMS accelerometers are capable

measuring both constant and changing accelerations, the output from the MEMS

accelerometer is non-zero both before and after the anchor drop. The slight difference in

output before and after the anchor drop is a measure of the higher acceleration level

associated with the increased radius of the anchor after falling from a point above the

mudline to a point beneath the mudline. As discussed previously, this allows MEMS

(but not piezoelectric) accelerometers to be used to measure the depth (or height) of a

buried object, as is the case here.

- 119 -

Figure 5.7. Dynamic anchor experimental arrangement showing the anchor

located in installation guide before release and embedded in the soil sample after

release

Anchor in soilSample

Water

Anchor in soilSample

Water

Anchor in guide

Mooring line

Anchor release mechanism

Mooring line

Anchor installation guide

Side view End view

- 120 -

Figure 5.8. MEMS and piezoelectric accelerometer data measured before, during

and after a dynamically installed anchor drop in a beam centrifuge: (a) entire

trace, (b) during freefall and embedment

(a)

(b)

- 121 -

Accelerations measured during the test are shown more clearly on Figure 5.8b, in which

the time scale has been clipped to show the anchor release and embedment in more

detail. The electrical noise on the MEMS accelerometer signal is approximately ±15 g,

compared with ±2.5 g on the piezoelectric accelerometer signal. As a ratio of the full

range output, the noise level of the MEMS accelerometer is approximately 100 times

higher than the piezoelectric accelerometer. The higher noise for the MEMS

accelerometer is well-known (e.g. Ratcliffe et al. 2008), and is higher in these dynamic

tests as no time averaging was applied to the data sampled by the ADC. However, the

±15 g noise level on the MEMS accelerometer is not considered to be problematic as

the data are eventually integrated to obtain velocities and displacements, effectively

averaging the noise error.

The saw tooth response at the beginning of the drop is a consequence of the anchor

release method. This was achieved inflight using a resistor which, when supplied with

current, heated and subsequently burned through a sacrificial anchor release cord. The

cord consists of multiple strands which likely parted sequentially, causing the changes

in acceleration over the first couple of milliseconds of the drop. During free-fall, the

MEMS accelerometer outputs approximately zero acceleration as expected, since there

is no force acting on the anchor except for a small component of Earth‟s gravity. This

acts at an inclination of tan-1(120) to the anchor giving a component of 0.01 g.

The piezoelectric accelerometer initially responds with the correct change in

acceleration of ,120g (i.e. the centrifuge acceleration level at the radius of the anchor

release location), but then starts to discharge towards zero as the acceleration during

free-fall is essentially constant and zero. This precludes the piezoelectric accelerometer

from being used as a means of establishing the free-fall velocity profile and hence the

anchor‟s velocity at the mudline.

- 122 -

5.6. Interpretation of accelerometer data

Numerical integration of acceleration yields velocity (after a single integration) and

displacement (after double integration), allowing a depth profile of acceleration and

velocity to be established for the anchor as it penetrates the soil. The integration is

undertaken within the rotating frame of reference, such that it is the anchor acceleration,

velocity and displacement relative to the soil surface that is considered.

As the precise release point can be difficult to identify (see Figure 5.8), the most

appropriate starting point for the numerical integration is when the anchor is stationary

relative to the soil at zero (relative) velocity and at a measured embedment depth. In this

test, the final anchor embedment depth was measured at 1g by clamping the release cord

trailing from the anchor at the soil surface, extracting the anchor, and then measuring

the distance from the clamp location to the anchor tip.

The integration is performed in reverse from the end of the test when the anchor is at

rest in the soil and so the relative acceleration is zero and the centrifuge (and hence

anchor) acceleration and radius is given by the MEMS accelerometer measurements.

The anchor acceleration relative to the soil during embedment and free-fall can then be

calculated using Equation 5.2, where the updated radius is determined from the at-rest

radius and the change in embedment calculated from the integration. The MEMS

acceleration data in Figure 5.8 are replotted in Figure 5.9a together with the centrifuge

(and soil) acceleration associated with the current anchor location during freefall and

embedment, and the anchor velocity and displacement relative to the soil.

The anchor velocity relative to the soil increases during free-fall due to the centrifuge

acceleration of the soil. At time t = 1.316 s the impact with the soil occurs which

corresponds to a peak anchor velocity, v = 22 m/s and an anchor displacement, ∆z =

80.4 mm (measured up from the at rest position). The directly measured anchor tip

- 123 -

embedment depth was 80.5 mm, which implies that the anchor reached its maximum

velocity and began to decelerate as soon as it impacted the soil surface. The anchor

comes to rest in the soil at t = 1.323 s before rebounding slightly. This effect is most

evident on the acceleration trace, with the numerical integration producing a lessening

effect on the velocity and displacement. This rebound has been reported in other studies

involving free-fall objects (e.g. Chow & Airey, 2010; Dayal & Allen, 1973; Morton &

O‟Loughlin, 2012), and is attributed to elastic rebound of the soil.

- 124 -

Figure 5.9. Interpretation of the MEMS accelerometer data in the rotating frame

of reference: (a) acceleration, velocity and displacement traces, (b) velocity profile

during freefall and embedment

Figure 5.9b shows the velocity profile determined from the acceleration data in Figure

5.9a, verified by independent measurements of the velocity during free-fall using

(a)

(b)

- 125 -

photoemitters and photo-receivers located on the anchor installation guide above the

sample surface (Richardson et al. 2006). The velocity profile implies a drop height of

222 mm. This compares to an anchor drop height of 240 mm that was set at 1g before

the test. However, this height is reduced by stretching of the release line as the anchor

becomes heavier during centrifuge spin up and as the line unravels during the burn. The

directly measured tip embedment depth (80.5 mm) is in excellent agreement with the

point of maximum velocity. This is to be expected for this sample owing to the surface

crust with a strength of about 20 kPa. The resulting penetration resistance is sufficient to

immediately cause a net deceleration, overcoming the effective weight of the anchor

caused by the acceleration of the soil relative to the anchor. As indicated earlier, for

these tests the MEMS accelerometer has a displacement resolution of 0.82 mm, which is

comparable to that achieved by the direct measurement using a scale rule marked in

millimetre divisions.

Figure 5.10 shows another velocity profile from an anchor installation in normally

consolidated kaolin clay with an undrained shear strength profile, su = 1.1zs, where zs is

the equivalent prototype depth of the soil sample. Unlike the calcareous silt example on

Figure 5.8b, the anchor velocity continues to increase after impact with the sample

surface. This is attributable to the very low resistance afforded by the very soft clay near

the sample surface. At about 30 mm penetration the resistance available from the soil

exceeds the submerged weight of the anchor and the anchor begins to decelerate.

- 126 -

Figure 5.10. Anchor velocity profiles for an anchor installation in kaolin clay with

su = 1.1z

As shown by Figure 5.7, in these experiments the vertical installation guide located

above the centreline of the soil sample prevents the anchor tilting during the freefall

stage, and applies the tangential force needed to keep the anchor rotating at the same

angular velocity as the centrifuge, which requires an increasing tangential velocity with

increasing radius. Once the anchor is embedded in the soil, this force must be applied by

the soil and there may be a tendency for the anchor to follow a curved trajectory through

the sample. This effect could be more significant when the soil is softer, such as in the

kaolin clay test (where the tip embedment was 2.5 times the anchor length compared

with 1.2 times the anchor length for the calcareous silt test) as the curved path is more

likely if the soil does not provide the resistance needed to accelerate the anchor

tangentially.

- 127 -

If the anchor tilts, the sensing axis of the MEMS accelerometer will not be coincident

with the centripetal acceleration vector, and the resulting acceleration measurements

will be lower than the centrifuge acceleration. The influence of tilt during anchor

embedment in soil is explored further in Figure 5.11, where various tilt angles are

artificially introduced to the MEMS acceleration trace measured in the kaolin clay test

at an anchor tip embedment of one anchor length. As the integration starts in reverse

from the at-rest position, the effect of tilt (applied to the acceleration trace only when

the anchor is free of the guide) has a cumulative effect all the way to the release point.

Figure 5.11 shows that tilt angles up to 10˚ have negligible effect on the velocity profile,

with a maximum reduction in impact velocity and final embedment depth of 1.4% and

1% respectively. However, for a tilt angle of 20˚ the reduction increases to 5.4% on

impact velocity and 3.7% on final embedment depth, and for a tilt angle of 30˚ impact

velocity reduces by 11.8% and final embedment depth reduces by 8.1%. Although the

final tilt angle can be quantified by using a dual or tri-axis accelerometer, the agreement

between the MEMS measurements and those made independently indicates that the

anchor tilt angle was negligible in the tests considered here.

Commercially available MEMS accelerometers are generally limited to a maximum

rated acceleration level of ±500 g. As such the sensor signal can become saturated

during impact events at higher centrifuge acceleration levels or for samples with higher

strengths. Figure 5.9a shows MEMS acceleration levels that are above the 500 g

maximum rated range of the sensor, which will have an effect on the linearity of the

sensor output at these „excess‟ acceleration levels. These two limitations need to be

considered in the experimental design.

- 128 -

(a)

(b)

- 129 -

Figure 5.11. Effect of anchor tilt during embedment in soil (kaolin clay with su =

1.1z): (a) 5 degree tilt, (b) 10 degree tilt, (c) 20 degree tilt, (d) 30 degree tilt

(c)

(d)

- 130 -

5.7. Concluding remarks

MEMS accelerometers are small, inexpensive sensors that have rarely been used in

geotechnical centrifuge modelling despite widespread use in consumer devices and for

industrial applications. Unlike piezoelectric accelerometers, MEMS accelerometers

measure both constant and varying acceleration. This feature has been exploited in the

paper to measure the motion response of a dynamically installed anchor as it free-falls

through air and embeds within a centrifuge soil sample. The capability of the MEMs

device to detect constant acceleration levels allows the centrifuge acceleration

component, and its variation with radial position, also to be detected. The motion

response can be used to produce acceleration or velocity depth profiles for the anchor;

such profiles are important for the calibration and validation of embedment prediction

tools for dynamically installed anchors.

- 131 -

CHAPTER 6. IN-SITU MEASUREMENT OF THE

DYNAMIC PENETRATION OF FREE FALL

PROJECTLES IN SOFT SOILS USING A LOW COST

INERTIAL MEASUREMENT UNIT

6.1. 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

This chapter describes a novel theoretical framework for interpretation of in situ IMU

data. This framework is then used in Chapter 7 in the analysis of the IFFS test data

and a reduced, simplified form of the framework is used in the analysis of the IFFS

centrifuge data.

Blake, A. P., O'Loughlin, C. D., Morton, J. P., O' Beirne, C., Gaudin, C. & White, D.

J., 2015. In-situ measurement of the dynamic penetration of freefall projectiles in soft

soils using a low cost inertial measurement unit. Geotechnical Testing Journal,

ASTM. DOI: 10.1520/GTJ20140135.

- 132 -

soft clay. The paper 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.

6.2. 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.

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

- 133 -

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 6.1.

- 134 -

(a)

(b)

Figure 6.1. 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

- 135 -

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 & 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 free-falling projectiles in clay (O‟Loughlin et al. 2014; Chow et

al. 2014), earthquake accelerations (Cilingir & 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 paper, 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 soil strength in the vicinity of

- 136 -

the embedded projectile and (for the case of the anchors) how this strength will be

mobilised during loading.

This paper 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.

6.3. Free-falling projectiles

6.3.1. Deep penetrating anchors

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 6.2), 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.

- 137 -

Figure 6.2. Deep penetrating anchor

6.3.2. Dynamically embedded plate anchors

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

6.3). 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

6.3.

- 138 -

Figure 6.3. Dynamically embedded plate anchor

6.3.3. Instrumented free-falling sphere

The instrumented free-falling sphere (IFFS) has been proposed as an in-situ

characterisation tool for soft soils (Morton & O‟Loughlin 2012; O‟Loughlin et al.

2014). 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 6.4).

- 139 -

Figure 6.4. Instrumented free-falling sphere

6.3.4. 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 6.5) 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 6.6 (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

- 140 -

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.

Figure 6.5. Inertial Measurement Unit

- 141 -

Figure 6.6. Body frame of reference

6.4. 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 paper, 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, ωbz and

accelerations Abx, Aby, Abz measured by the IMU will not be coincident with the inertial

ωby

ωbx

ωbz

yb xb

zb

Aby Abx

Abz

- 142 -

frame (see Figure 6.7). 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 6.7. Resultant tilt angle, μ, defined in the inertial frame

6.4.1. Rotation

The body frame rotation rates ωbx, ωby, ωbz were transformed from the body frame to the

inertial frame to correspond with rotation rates about the inertial frame ωx, ωy, ωz using

an angular velocity transformation matrix (AVTM), ibT (Fossen 2011):

bz

by

bxi

b

z

y

x

T

6.1

y

zb

μ

z

x

xb

yb

- 143 -

bbbb

bb

bbbi

b

b

T

cos/coscos/sin0sincos0

tancostansin1 6.2

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 dttt00 )()( 6.3

t

bybb dttt00 )()( 6.4

Similarly, the rotation angle ψb about the body frame axis zb was established by

numerical integration of ωbx:

t

bzbb dttt00 )()( 6.5

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 x, y and z

respectively (Euler angles) to be established:

t

x dttt00 )()( 6.6

t

y dttt00 )()( 6.7

- 144 -

t

z dttt00 )()( 6.8

6.4.2. Acceleration

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)

ibR (Nebot & Durrant-Whyte 1999; Jonkman 2007; King et al. 2008; Fossen 2011):

bz

by

bxib

z

y

x

AAA

RAAA

6.9

xyzib RRRR 6.10

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:

1000cossin0sincos

zR 6.11

cos0sin010

sin0cos

yR 6.12

- 145 -

cossin0sincos0

001

xR 6.13

Multiplication of the Rz(-ψ), Ry(-θ) and Rx(-ϕ) rotation matrices gives the DCM:

coscossincossinsinsincossincossinsinsincoscossincossincoscossinsinsinsincossincoscoscos

ib

R

6.14

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):

gAAA

aaa

z

y

x

z

y

x

00

6.15

The resultant linear acceleration, a (acceleration in the direction of motion), was

calculated as:

gAAAa zyx 222 6.16

- 146 -

6.4.3. Velocity and distance

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 dttavtv00 )()( 6.17

t

yyy dttavtv00

)()( 6.18

t

zzz dttavtv00 )()( 6.19

The resultant projectile velocity v, was calculated using the following expression:

222zyx vvvv 6.20

The resultant projectile velocity v, was numerically integrated to establish the distance

travelled by the projectile along its trajectory s:

t

dttvsts00 )()( 6.21

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 )()( 6.22

- 147 -

6.4.4. 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 6.7) angles using the following

expressions:

gAby

acc1sin (King et al. 2008) 6.23

gAbx

acc1sin (King et al. 2008) 6.24

gAbz1cos (Stephan et al. 2012) 6.25

6.5. 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

6.8.

- 148 -

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 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/m3 at about 3.5 m (limit of the sampling depth). Figure 6.9a

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 cm2 cone), and

piezoball bearing factors Nball = 11.5 (50 cm2 ball) and Nball = 12.2 (100 cm2 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 ratio of

remoulded to intact soil resistance 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 & 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

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 & 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 6.9b 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

- 149 -

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 6.8. Test sites locations

Lough Erne

Firth of Clyde

- 150 -

Figure 6.9. Undrained shear strength profiles: (a) Firth of Clyde and (b) Lough

Erne

6.6. Test procedure

Testing was conducted using the RV Aora, a 22 m research and survey vessel in Firth of

Clyde (Figure 10a) and either a fixed vessel berthing jetty or a 15 m self-propelled

barge (Figure 10b) 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 6.11 for the DEPLA tests using the

RV Aora) and involved the following stages:

1. The IMU was powered up and secured in the projectile.

- 151 -

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 6.12).

Figure 6.10. (a) RV Aora and (b) Self-propelled barge

- 152 -

Figure 6.11. DEPLA field test procedure

h iz,d

rop

WaterSeabed

(a)

Follower line

DEPLA

Plate line

RV AoraCrane Pulley

Drum WinchQuick release shackleA-frame

Water

(b)

Follower linePlate line

Pulley

Drum Winch

Crane

A-frameRV Aora

Seabed

Water

(c)

Follower line

ROV cable

ROV

Plate line

z iz,e

PulleyCrane

A-frame

Load cell

Pulley

RV Aora

Seabed

- 153 -

Figure 6.12. Image capture from ROV camera showing the follower retrieval line

at the seabed

6.7. 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.

6.7.1. Rotation

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 6.13 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

- 154 -

(DEPLA), 5.95 m (IFFS), and 3 m (DPA). In Figure 6.13 ϕacc and θacc are rotations

relative to the inertial frame deduced from the horizontally orientated y- and x-axes

accelerometers using Equations 6.23 and 6.24, ϕb, θb and ψb are rotations about the body

frame axes xb, yb and zb established using Equations 6.3, 6.4 and 6.5, and ϕ, θ and ψ are

the pitch, roll and yaw rotations about the inertial frame axes x, y and z derived using

Equations 6.6, 6.7 and 6.8.

In Figure 6.13a, 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 6.13b shows that the IFFS rotated about all three axes during freefall 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

- 155 -

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.

In contrast, rotations measured during the DPA free-fall and embedment phases (Figure

6.13c) 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.

- 156 -

Figure 6.13. Projectile rotations during free-fall through water and soil

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

- 157 -

6.7.2. Acceleration

Figure 6.14 shows acceleration profiles for the same tests as shown in Figure 6.13. In

Figure 6.13 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. Az is the acceleration measurement in the direction of gravity). In Figure 6.14a 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

6.15). 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/s2). 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/s2). 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 & Allen 1973; Chow & Airey 2010; Morton & O‟Loughlin 2012;

O‟Loughlin et al. 2014), 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 (az = 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 6.14b) is broadly similar to that of the

DEPLA, with the expected change in acceleration upon release and the subsequent

- 158 -

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 6.14c shows the acceleration response for the DPA test. The response is

qualitatively similar to those shown in Figure 6.14a and Figure 6.14b 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 where

relatively small for this test.

- 159 -

Figure 6.14. Projectile accelerations during free-fall through water and soil

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

- 160 -

6.7.3. Velocity profiles

Figure 6.15 shows velocity profiles for free-fall in water and embedment into soil for

the three tests considered previously and shown in Figures 13 and 14. The velocity vz,

and distance sz, (i.e. depth) relative to the inertial frame were established using

Equations 6.19 and 6.20. The velocity vbz, and distance zbz, were also derived from

Equations 6.19 and 6.20, albeit with abz = Abz + g, instead of az and Az. vbz and zbz

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 6.15a 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 + 2.8z. Figure 6.15b 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).

Figures 15a and 15b 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 6.15c) 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 6.15 are direct measurements of the final embedment depths based on

- 161 -

mudline observations of markings on the retrieval line using a ROV in 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 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.

- 162 -

Figure 6.15. Projectile velocity profiles corresponding to free-fall through water

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

- 163 -

6.7.4. Verification of the IMU derived measurements

Independent verification of the IMU-derived measurements of the projectile

displacement (Equation 6.22) 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 6.16. 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 6.15). 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.

- 164 -

Figure 6.16. Comparision of IMU derived displacement measurements with those

obtained using a draw wire sensor

6.7.5. 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, O‟Loughlin et al. 2014; 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 6.13 and Figure 6.14 play an important role in the validation

and calibration of such predictive models. An example comparison is provided in Figure

6.17 for the DEPLA, where the predictions are based on an analytical model described

- 165 -

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 & Wilson 1951, Graham et al. 1983, Sheahan et al. 1996), whilst

also accounting for drag resistance and the buoyant weight of the displaced soil.

Consideration of these resistance components leads to the following governing

equation:

dbearbearffrictfrictfbs FFRFRFWdt

sdm ,,2

2

6.26

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, Rf,frict and Rf,bear the

strain rate function associated with the bearing and frictional resistance and Fd is drag

resistance, formulated as:

2

21 vACF psdd 6.27

where s is the submerged density of the soil, Cd is the drag coefficient, Ap is anchor

projected (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

- 166 -

et al. 2012, Einav & Randolph, 2006, White 2012, Sahdi et al. 2014). O‟Loughlin et al.

(2013) and Blake & O‟Loughlin (2015) further showed that 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 6.28

pubear ANsF 6.29

where α is an interface friction ratio (of limiting shear stress to undrained shear

strength), As is anchor shaft area, N is the bearing capacity factor for the projectile tip or

fluke, and su is the undrained shear strength averaged over the contact area, Ap or As.

The reference undrained shear strength adopted in Equation 6.28 is the idealised profile

shown in Figure 6.9a, which is enhanced using a power law strain rate function

(Biscontin & Pestena 2001; Peuchan & Mayne 2007; Randolph et al. 2007; O‟Loughlin

et al. 2013b) expressed as:

refreff dv

dvnR

6.30

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 associated with

the measurement of the undrained shear strength. The factor n in Equation 6.30

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 =

- 167 -

1 for tip resistance (Zhu & Randolph, 2011) and as a function of (adopted from

Equation 8b in Einav & Randolph, 2006) for estimating rate effects in shaft resistance

according to:

22 l

l nβnn 6.31

where nl is 1 for axial loading.

The predictions on Figure 6.17 were obtained using bearing capacity factors of N = 7.5

for the leading and trailing edges of the flukes (analogous to a deeply embedded strip

footing) and N = 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

video capture of the drop sites (see Figure 6.12) 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 & 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. Vaid & Campenella 1977, Graham et al. 1983,

Lefebvre & 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 6.17 indicates that the inclusion of a fluid-mechanics drag resistance

term is appropriate for projectiles penetrating soft clay at high velocities. There is

- 168 -

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

drag resistance required α = 0.38, which is inconsistent with results from the cyclic

piezoball remoulding tests and gave a much poorer match.

Figure 6.17. DEPLA velocity profile derived from the IMU data measured at the

Firth of Clyde test site and corresponding theoretical profile

6.8. Conclusions

This paper 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‟

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

tanc

e, s

(m)

- 169 -

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 paper showed that for projectiles that tilt during free-

fall, 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 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.

- 170 -

CHAPTER 7. ESTIMATION OF SOIL STRENGTH BY

INSTRUMENTED FREE-FALL SPHERE TESTS

7.1. Abstract

Dynamic penetration of rigid bodies into soil is a complex problem as it involves

inertial effects and extreme strain rates that enhance the soil strength. The dynamic

response of a sphere in soft clay is considered in this paper through field tests in which a

0.25 m diameter steel sphere was allowed to free-fall in water and dynamically

penetrate the underlying soft soil. The test data, which were collected in a lake and a

nearshore environment, relate to sphere velocities of up to 8 m/s, reaching sphere invert

This chapter extends Chapter 3 by further reporting field test data measured in IFFS

tests at two soft soil sites. The measurement technique introduced in Chapter 6 for

interpreting the IMU data is used to interpret the net penetration resistance acting on

the IFFS that extends the simplified model described in Chapter 3 and also

incorporates the shallow penetration framework described in Chapter 4. The

framework allows the undrained shear strength, su, to be devised from the IFFS net

penetration resistance and the chapter concludes by comparing these measurements

with those obtained in equivalent pushed-in ball penetration tests.

Morton, J. P., O‟Loughlin, C. D. & White. D. J., 2015. Field testing an in situ

freefalling spherical penetrometer in soft soil. Submitted to Géotechnique.

- 171 -

embedments close to 10 diameters. An interial measurement unit located within the

sphere measured the motion response of the sphere during free-fall and penetration in

soil. The resulting acceleration data were used within a simple framework that accounts

for both geotechnical shearing resistance and fluid mechanics drag resistance, but cast

in terms of a single capacity factor that can be expressed in terms of the non-Newtonian

Reynolds number. The merit of the framework is demonstrated by using it as a forward

model in a series of inverse analyses that calculate the undrained shear strength profile

from acceleration data measured in free-fall sphere tests. The good match between these

profiles and those obtained from push-in piezoball penetrometer tests point to the

potential for an Instrumented Free-Fall Sphere (IFFS) to be used a tool for charcterising

the near surface strength of soft seabeds.

7.2. Introduction

Geotechnical aspects of offshore and near-shore infrastructure projects often involve the

assesment of complex soil-structure interactions involving strain rates than are often

several orders of magnitude higher than those associated with laboratory element tests.

The most extreme examples relate to dynamic impact events such as submarine

landslide runout on pipelines and installation of free-fall projectiles such as soil

samplers, penetrometers for soil strength estimation and dynamically installed anchors.

Such problems require an assessment of the net dynamic penetration resistance, which

can generally be resolved into two separate components. The first is the strain-rate

dependant geotechnical component (comprising the bearing resistance and for slender

projectiles, frictional resistance) which represents the strength-dominated domain, and

the second is the fluid dynamics component, which is the drag resistance that represents

the inertia-dominated domain. Morton & O‟Loughlin (2012) adopt this „summation‟

- 172 -

approach to assess the net penetration resistance on a sphere, Fresist, as it dynamically

penetrates soft soil:

popucpDresist AsNvACF ,2

21

7.1

The first term on the right-hand side of Equation 7.1 is the fluid dynamics drag force

(Fd), where CD is the drag coefficient (dictated by object geometry and roughness), ρ is

the density of the medium, Ap is the full projected area of the sphere, D2/4, and v is the

velocity. The second term is the geotechnical bearing force (Fsu,op) in which Nc is the

bearing capacity factor and su,op is the strain-rate-enhanced undrained shear strength of

the soil.

In other studies, only the first term of Equation 7.1 has been considered and the drag

coefficient, CD has been defined as a function of the non-Newtonian Reynolds number,

Renon-Newtonian, also known as the Johnson number, (Zakeri et al. 2008; Zakeri, 2009;

Zakeri et al. 2009):

2

Re vNewtoniannon 7.2

where τ is the mobilised shear stress within the flowing material, referred to hereafter as

the operative shear strength, su,op.

When only the drag (CD) term is used, the variation of CD with Re-non-Newtonian (Equation

7.2) introduces an influence from the mobilised shear stress (or shear strength).

However, this approach is not favoured for two reasons. Firstly, in the theoretical limit

of a weightless medium, such an approach predicts zero resistance regardless of the

- 173 -

shear stress or strength. Secondly, it also predicts zero resistance at zero velocity,

meaning that an object can never reach a stationary equilibrium, which is the end point

of all free-fall penetrometer tests.

The influence of strain rate on the mobilised strength due to viscous effects is

captured in the second term of Equation 7.1 by a power law relationship for the soil

strength (Briaud et al. 1984; Biscontin & Pestana, 2001; Peuchen & Mayne, 2007;

Jeong et al. 2009; Lehane et al. 2009):

refrefuopu v

vss

,, 7.3

where su,ref is the undrained shear strength at the reference strain rate refγ and β is the

strain rate parameter.

However, if only the strength (Nc) term of Equation 7.1 is used, poor predictions are

made for high speed penetration events (e.g. Georgiadis, 1991). This is because drag

forces are increasingly significant, as identified at shallow embedment (particularly for

bluff bodies) by O‟Loughlin et al. (2013b) and Blake & O‟Loughlin (2015).

Instead, a superposition approach is favoured with separate terms for the resistance

associated with the self-weight of the medium and the shear stress (or strength) within

the medium (Equation 7.1). With this hybrid relationship the total resistance can be

expressed as a single bearing capacity factor, N:

popu

resist

AsFN

,

7.4

- 174 -

where Fresist is the resistant force acting on the sphere given by Equation 7.1 (i.e. Fresist =

Fd + Fsu,op). It follows from Equations 7.1, 7.2 and 7.6 that a relationship between N and

Renon-Newtonian may then be expressed as:

cNewtoniannonD NCN Re21 7.5

This methodology captures the impact force of submarine slide debris on a pipeline in

both numerical (Randolph & White, 2012) and experimental studies (Sahdi et al. 2014).

The aim of this paper is to consider an analogous framework for the dynamic

penetration of a sphere in soft soil, using data from free-fall tests at two soft soil sites.

7.3. Bearing capacity factor

Throughout the analysis described in this paper, the sphere bearing capacity factor, Nc,

is dependent on interface roughness and the dimensionless soil strength ratio, su/γ'D

(White et al. 2010; Morton et al. 2014), where γ' is the unit weight of the soil and D is

the sphere diameter. Plasticity limit analyses provide bounds on Nc for a deeply

embedded sphere that lie in the range Nc-deep = 10.98 to 11.6 for a fully smooth sphere

and Nc-deep = 15.10 to 15.31 for a fully rough sphere (Randolph et al. 2000). The

transition from a reduced Nc value at shallow penetration where the failure mechanism

extends to the soil surface, to the limiting value for deep embedment associated with a

full-flow mechanism, has recently been investigated for a sphere over a wide range of

su/γ'D ≈ 0.07 to 40 and can be expressed by the following power function (Morton et al.

2014):

p

opdeep

opdeepbc w

wNN

ˆˆ

7.6

- 175 -

where the transitional depth at which Nc = Nc-deep is given by

f

u

cu

opdeep

eD

s

adD

sbaw

/1

ˆ

7.7

and the fitting constants a = 16.3, b = 0.12, c = 1.3, d =0.52, e = 4.9, f = 1.5 and p =

0.49.

7.4. Site description and soil properties

The Instrumented Free-Fall Sphere (IFFS) tests were conducted at two sites; the first

was Lough Erne which is a freshwater lake in County Fermanagh, Northern Ireland and

the second was Firth of Clyde, which is located off the West coast of Scotland between

the mainland and the Isle of Cumbrae. Water depths at the testing locations were 1 to 12

m at Erne and approximately 50 m at Clyde. Classification tests were conducted on bulk

and tube samples retrieved from Erne and on piston samples retrieved from Clyde. The

lakebed at Erne is very soft clay with moisture contents in the range 270 to 520%,

typically about 1.5 times the liquid limit, and a fines fraction of 95%. The measured

bulk unit weight of the Lough Erne clay is only marginally higher than water at γ = 10.8

kN/m3. This is due to the very high proportion of diatoms that are evident from

scanning electron microscopic images of the soil (Colreavy et al. 2012). These diatoms

have an enormous capacity to hold water in the intraskeletal pore space (Tanaka &

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. The seabed at Clyde is also very

soft, with moisture contents in the range 50 to 100% (close to the liquid limit) and a

- 176 -

fines fraction of 80%. Consistency limits plot above or on the A-line on the Casagrande

plasticity chart, indicating a clay of intermediate to high plasticity. The bulk 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 7.1a and b shows profiles of undrained shear strength, su, with depth derived

from piezoball penetrometer tests and a combination of in situ vane tests (Erne),

laboratory vane tests (Clyde) and fall cone tests (Clyde) The piezoball tests used a

penetrometer with a ball diameter of 113 mm in Erne and ball diameters of 113 and 80

mm in Clyde. An exact 10:1 ratio between the ball and shaft areas was maintained for

all penetration tests by using different shafts for each piezoball. A penetration rate of 20

mm/s (as is standard for cone testing) was adopted in an attempt to obtain undrained

conditions during penetration. The best agreement between the piezoball profiles and

the other strength measurements was obtained using Nc-deep = 11.8 for Clyde and Nc-deep

= 8.5 for Erne.

- 177 -

Figure 7.1. Undrained shear strength profiles in: (a) Lough Erne and (b) Firth of

Clyde

- 178 -

7.5. Test equipment and testing procedures

7.5.1. Instrumented free-fall sphere

The custom-made Instrumented Free-Fall Sphere (IFFS, see Figure 2a and 2c) is 250

mm in diameter and consists of two mild steel hemispheres that bolt together. An O-ring

located between the two hemispheres prevents the ingress of water to protect an inertial

measurement unit (IMU, described in the following section) that can be located within

vertical cylindrical voids in each hemisphere. The sphere and IMU have a dry mass, m

= 51.25 kg and an effective mass when submerged in fresh water of 43.07 kg. A 12 mm

diameter Dyneema SK75 rope was used for deploying and recovering the IFFS.

Figure 7.2. Instrumented free-fall sphere: (a) sphere separated with IMU located

within internal void, (b) IMU and (c) assembled sphere prior to a free-fall test in

Erne

7.5.2. Inertial measurement unit

The inertial measurement unit (IMU) used in the free-fall sphere tests (shown in Figure

7.2b) is a fully self-contained motion logger designed to capture the motion history of

free-fall projectiles. The IMU includes a 16-bit three component MEMS rate gyroscope

- 179 -

(ITG 3200) and a 13-bit three-axis MEMS accelerometer (ADXL 345). The gyroscope

has a resolution of 0.07 °/s with a measurement range of +/- 2000 °/s. The

accelerometer has a resolution of 0.04 m/s2 with a measurement range of +/-16 g. Data

are logged by an mbed micro controller with an ARM processor to a 2 GB SD card at

400 Hz. Internal batteries are capable of powering the logger for up to 4 hours. In the

free-fall sphere tests the IMU was contained in a watertight aluminium tube 185 mm

long and 42 mm in diameter that fitted securely within the internal cylindrical void in

the sphere.

The MEMS accelerometer measures both linear and gravitational acceleration

(depending on the sphere orientation) in three orthogonal body-frame axes that are

common to both the IMU and the sphere. In order to distinguish the sphere‟s linear

acceleration component from the acceleration detected by the sensor, which may be

slightly rotating, it is important to transform the body frame acceleration measurements

to accelerations that are coincident with the Earth-fixed inertial frame using rotation

matrices, described in detail by Blake et al. (2015). Linear accelerations corresponding

to the inertial frame z-axis (i.e. in the direction of Earth‟s gravity) were numerically

integrated to establish the sphere velocity and displacement. Figure 7.3 shows a typical

acceleration measurement of the sphere, from a hanging position to free-falling in water,

through penetration in soil, before a slight rebound due to soil elasticity (Dayal & Allen,

1973; Chow & Airey, 2010) as the sphere comes to rest. The importance of

transforming the measured accelerations to the inertial frame is evident in Figure 7.3 as

the body frame velocity is non-zero and the displacement is not constant when the

sphere is at rest. For the remainder of the paper the inertial frame z-axis linear

acceleration is referred to as the vertical acceleration.

- 180 -

Figure 7.3 IMU measurements and their interpretation from a typical free-fall

sphere test in Erne

7.5.3. Field testing procedure

The field testing program involved 87 tests that were carried out from a fixed pontoon

or a hopper barge in Erne (Figure 7.2c) and from a research and survey vessel in the

Clyde. The drop height, impact velocities and final embedments are summarised in

Table 1. The IFFS was released from various heights above the mudline to permit an

assessment of the embedment response over a range of travelling velocities. The release

height in the 72 Erne tests was varied between 0 and 9.23 m, with resulting impact

velocities in the range 0 to 7.9 m/s and sphere invert embedments of up to d = 2.373 m

(~9.5 sphere diameters). In the 15 Clyde tests, the IFFS release height was varied

between 1 and 20 m, which resulted in impact velocities between 4.0 and 6.3 m/s and

- 181 -

sphere invert embedments of up to d = 0.782 m (~ 3.13 sphere diameters). Embedment

depths were established from the IMU data as described previously, with direct

measurements based on mudline observations of markings on the retriveal line using a

remotely operated vehicle in Clyde and a drop camera in Erne. These direct

measurements were made to confirm that the analyses had not produced any gross error,

as the accuracy in the mudline obervations was much lower that was possible from the

IMU data.

- 182 -

Table 7.1. Free-fall sphere test data from the Erne tests

Test no.

Release height (m)

Impact velocity (m/s)

Sphere invert embedment depth

(m) 1 0.50 2.69 1.017 2 0.50 0.01 1.115 3 0.50 2.79 1.150 4 1.00 4.01 1.439 5 1.00 3.87 1.595 6 1.00 3.89 1.500 7 1.00 3.88 1.484 8 3.00 5.79 2.378 9 4.00 6.63 2.055 10 5.00 6.80 2.152 11 2.00 5.25 1.841 12 3.00 6.15 1.974 13 1.00 3.74 1.097 14 0.50 2.46 1.128 15 1.00 2.46 1.128 16 2.00 5.22 1.855 17 3.00 6.11 1.846 18 3.00 6.44 2.117 19 0.50 2.64 1.243 20 1.00 3.80 1.295 21 1.66 5.19 1.572 22 3.00 6.20 1.995 23 0.44 2.58 1.088 24 0.54 2.81 1.116 25 0.61 2.97 1.148 26 0.50 2.72 1.148 27 0.02 0.55 1.192 28 1.04 3.94 1.308 29 1.07 3.93 1.303 30 2.13 5.38 1.718 31 1.82 5.23 1.620 32 2.22 5.70 1.456 33 2.00 4.86 1.613 34 2.62 6.15 1.732 35 2.55 6.30 1.956 36 2.84 6.35 1.804 37 2.61 6.51 1.788 38 2.47 6.14 1.803 39 3.39 6.63 1.889 40 0.50 2.67 0.918 41 0.84 3.68 1.183

- 183 -

42 1.85 5.34 1.543 43 2.00 5.35 1.518 44 3.00 6.19 1.747 45 2.81 6.43 1.766 46 0.00 0.00 0.791 47 0.00 0.00 0.732 48 0.00 0.00 0.646 49 3.00 6.31 1.815 50 3.00 6.45 1.780 51 3.31 6.58 1.878 52 0.00 0.00 0.889 53 0.00 1.32 0.903 54 0.00 0.00 0.748 55 1.00 3.86 1.392 56 2.00 5.06 1.491 57 3.00 6.05 1.805 58 0.00 0.00 0.754 59 1.00 3.72 1.276 60 2.00 5.09 1.588 61 3.00 5.96 1.735 62 3.93 6.60 1.972 63 4.49 6.62 2.081 64 4.73 6.75 1.964 65 8.95 7.88 2.373 66 6.28 7.41 2.181 67 4.00 6.36 1.927 68 4.00 5.97 1.732 69 3.99 6.42 1.820 70 9.27 7.47 2.127 71 6.78 7.91 2.108 72 3.70 6.78 2.052

- 184 -

Table 7.2. Free-fall sphere test data from the Clyde tests

Test no.

Release height (m)

Impact velocity (m/s)

Sphere invert embedment depth

(m) 1 2 5.04 0.611 2 3 0.00 0.076 3 15 5.14 0.653 4 9 6.05 0.812 5 4.5 3.95 0.408 6 15 5.84 0.740 7 9 5.92 0.660 8 4.5 5.62 0.624 9 4.5 5.82 0.702 10 20 6.20 0.782 11 15 5.70 0.670 12 15 5.92 0.630 13 3 5.15 0.610 14 2 4.45 0.580 15 1 3.99 0.569

7.6. Forces acting on a sphere during free-fall in water

The hydrodynamic forces acting on the IFFS during free-fall in water (i.e. before it

impacts the underlying soil) are shown schematically in Figure 7.4a, which leads to the

following equation that governs the motion response during free-fall in water:

AMdsw FFFdtdvm 7.8

where dv/dt is the vertical acceleration (t is time), m is the mass of the IFFS, FSW is the

submerged weight of the IFFS in water, Fd is the drag force introduced earlier in

Equation 7.1. The final term on the right hand side of Equation 7.8 is the added mass

force, FAM =

Cmmwater that occurs during non-stable flows where an object is

accelerating or decelerating (Lamb, 1932). The added mass (Cmmwater) can be

interpreted as the mass of fluid displaced by the sphere that is accelerated with the

- 185 -

sphere, which is higher than the displaced mass, mwater by an amount controlled by the

added mass coefficient, Cm. Omitted from Figure 7.4a and Equation 7.8 is the drag force

that develops on the deployment and retrieval rope. This is intentional so that the

framework (Equation 7.5) can be considered in terms of a single contact area, Ap, for

both the geotechnical and fluid mechanics force components. The average drag

resistance on the rope (modelled using a friction drag coefficient of 0.015 (Blake &

O‟Loughlin, 2015) and the contact area between the rope and the water/soil) never

exceeded more than 2% of the total resistance.

- 186 -

Figure 7.4. Forces acting on a sphere: (a) free-falling in water, (b) during dynamic

penetration in soil

As the vertical acceleration is measured, the evolution of the drag coefficient, CD,

during free-fall in water can be established from Equation 7.8. This is shown on Figure

- 187 -

7.5 for three Erne tests over a calculated range of Reynolds numbers, Re = 104 to 106

together with the empirical correlation between CD and Re proposed by Morrison

(2013), that is based on the benchmark experimental summarised by Schlichting (1955).

At lower values of Re, between Re = 104 and 105 the back-figured experimental CD is

significantly higher than empirical values of CD, and reduces rapidly with increasing Re.

However, this range of Re corresponds with the start of free-fall, with sphere velocities

less than 0.5 m/s and free-fall displacements less than 0.02 m (0.08 diameters). A

similar rapid variation in CD has been associated with the hydrodynamic response of

dynamically installed „torpedo anchors‟, immediately after the onset of free-fall

(Fernandes et al. 2006). However, at Re > 106, when the sphere velocity is > 0.5 m/s

and the free-fall displacement is typically about 0.02 m, the experimental CD values

agree well with the Morrison (2013) relationship using Cm = 0.5, which is exactly

coincident with the theoretical value for a sphere (e.g. Sumer & Fredsoe, 1997,

Pantaleone & Messer, 2011).

- 188 -

Figure 7.5. Measured and theoretical evolution of the drag coefficient, CD, during

free-fall in water

7.7. Forces acting on a sphere during dynamic penetration in soil

When the IFFS impacts the mudline the geotechnical resistance must be considered. As

discussed earlier, the combination of fluid mechanics and geotechnical resistance may

be considered in terms of a single resistance factor, N, as given by Equation 7.7. The

forces acting on the sphere during penetration in soil are shown schematically in Figure

7.4b, which leads to a modified form of Equation 7.8 that governs the motion response

of the sphere in soil.

AMresistss FFFdtdvm 7.9

- 189 -

where FSS is the submerged weight in soil, calculated as the product of the soil unit

weight, γ, and the volume of displaced soil, Vdisp, which is the volume of the IFFS

currently embedded in the soil in addition to the volume of a cavity that may form in the

wake of the advancing sphere. This cavity may be approximated as (for example) an

inverted cone, with a depth given by Equation 7.5 (Morton et al. 2014). FAM is the added

mass force, FAM =

Cmmsoil. Fresist is the combined resistant force acting on the IFFS

given by Equation 7.1, Fd + Fsu,op. The latter term requires an estimation of the

operational shear strength, su,op, which may be calculated using Equation 7.3, but where

the strain rates are approximated as γ ⁄ (Hurst & Murdoch, 1991; Lehane et al.

2009; O‟Loughlin et al. 2009) such that:

refrefuopu DV

Dvss//

,, 7.10

where (v/D)ref is a proxy for the reference strain rate associated with the reference value

of undrained shear strength, su,ref. As shown on Figure 7.1a and b, the majority of the su

measurements were made using a 113 mm diameter piezoball penetrated at 20 mm/s, so

(v/D)ref = 0.18 s-1. The added mass in Equation 7.9 may be calculated in the same way

as for free-fall in water, using Cm = 0.5 but with the density, ρ, of soil rather than of

water, corresponding to an added mass of 4.2 kg in Erne and 6.1 kg in Clyde.

Rearranging Equation 7.9 allows N to be calculated directly from the vertical

acceleration measurements, as shown against Renon-Newtonian on Figure 7.6 for all the test

data included in Table 1 (Erne, Figure 7.6a) and Table 2 (Clyde, Figure 7.6b). The

operational strength, su,op, which is reflected in both the vertical and horizontal axes of

Figure 7.6a and b, was calculated using β = 0.07, which is at the lower end of the ranges

- 190 -

β = 0.05 to 0.17 quoted by Jeong et al. (2009) and β = 0.05 to 0.15 quoted by

O‟Loughlin et al. (2013b). β = 0.07 is consistent with a 17% increase in undrained shear

strength for every log cycle increase in shear strain rate, and is within the 10 to 20%

range commonly reported (e.g. Vaid & Campenella, 1977; Graham et al. 1983; Lefebvre

& Leboeuf, 1987).

The breadth of the range in β quoted in the literature may be (at least in part) attributed

to the order of magnitude difference between the operational strain rate and the

reference strain rate used in Equation 7.3. The maximum v/D values associated with the

IFFS tests considered here is v/D = 31 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 and laboratory vane tests, which tend 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) or β = 0.055

(Biscontin & Pestana, 2001).

The upper bound on Renon-Newtonian in Figure 7.6a approaches 104 in Erne and over 102 in

the Clyde. These upper bounds correspond with tests where the impact velocity was

close to the terminal velocity of the sphere and are lower for Clyde due to the higher

soil strength at the mudline when the sphere velocity is high. The lower bound Renon-

Newtonian = 10-4 and 10-1 for Erne and Clyde respectively were selected to omit the final

10 mm of embedment, where the sphere velocity was typically 0.2 m/s and the back

figured values of N begin to reduce suddenly, indicating that other forces (perhaps from

impact of the following rope or shackle) are involved.

Figure 7.6a and b highlights that at Renon-Newtonian < ~3, soil strength rather than drag is

the dominant form of resistance, whereas for Renon-Newtonian > ~3, N increases linearly

with Renon-Newtonian highlighting the dominance of drag resistance. The annotation

indicating the normalised depth (d/D) and approximate sphere velocity in Figure 7.6

- 191 -

shows that this occurs at shallow depths where the magnitude of su,op (linked to the

geotechnical component) is very low and the velocity is very high. Although the sphere

velocity is linked to both the geotechnical and fluid mechanics terms, the drag has a

much higher velocity dependency than the geotechnical term, and therefore is the

dominant form of resistance at Renon-Newtonian > ~3.

- 192 -

Figure 7.6. Relationship between N and Renon-Newtonian for: (a) Erne and (b) Clyde

- 193 -

Also shown on Figure 7.6 is the theoretical response described by Equation 7.5, which

provides a good fit to both datasets using CD = 0.26, and equivalent data for a cylinder

from experiments in a flume (Zakeri et al. 2008; Zakeri et al. 2011) and in a centrifuge

(Sahdi et al. 2014). These data exhibit the same trends as for the sphere, albeit that the

cylinder data plot above the sphere data at Renon-Newtonian > ~3, as at high Reynolds

numbers CD for a cylinder is higher than that for a sphere (Schlichting, 1955).

At Renon-Newtonian < ~3, N for the sphere occupies a wider range than for the cylinder as

the sphere. This is because the sphere embedment depth is changing, reflecting the

variation in Nc during shallow penetration (all cylinder tests were at a single

embedment). In Erne, the sphere penetrated to (on average) between 1 and 2 m,

equivalent to 4 to 8 diameters, which is sufficient in this soft soil to establish a full-flow

deep failure mechanism. The approximate range of Erne N values at Renon-Newtonian < ~3

is typically Nc-deep = 6 to 9.5, with an arithmetic mean of Nc-deep = 7.5, which is to be

expected given that the strength data input to the analyses were determined using Nc-deep

= 8.5 (as described earlier). Over the same range of at Renon-Newtonian, Clyde N values do

not reach Nc-deep = 11.8 used to interpret the piezoball data. The approximate range of

Clyde N values at Renon-Newtonian < ~3 is typically Nc-deep = 7 to 10, with an arithmetic

mean of Nc-deep = 9. This difference to the piezoball analysis is to be expected as the

final sphere embedment depth at Clyde is typically 0.65 m (2.6 sphere diameters),

which for this soil with a relatively high su/γ'D, is insufficient to develop a deep failure

mechanism (i.e. Nc < Nc-deep).

7.8. Soil strength estimation using free-fall sphere data

The merit of the framework described in the previous section is now explored by

performing a simple inverse analysis on vertical acceleration data measured during IFFS

- 194 -

tests to obtain su profiles for both sites. Rearranging Equation 7.9, the undrained shear

strength, su, at a given depth is:

refp

soilmss

u

vvNA

mcmdtdvF

s 7.11

In Erne, a deeply embedded soil flow mechanism is assumed (as described previously),

such that the buoyancy force may be calculated using the sphere volume and N may be

determined using Equation 7.5 with Nc = 8.5 and CD = 0.26. In Clyde the dynamic

sphere penetration is relatively shallow and the dimensionless strength ratio su/γ'D is

relatively high (~1.8 at the final sphere embedment depth), such that both the variation

in Nc with embedment depth and the cavity formed by the penetrating sphere require

consideration. It is assumed here that the transitional embedment depth at which Nc =

Nc-deep may be calculated using Equation 7.5 and the cavity geometry is an inverted cone

(Morton et al. 2014). This gives a transitional embedment depth at which Nc = Nc-deep

(measured to the invert of the sphere) of approximately 1.2 m. This allows the variation

in Nc with depth to be determined (using Equation 7.6) and the inverted cone volume to

be calculated. The variation in Nc is included in Equation 7.5 and the changing

buoyancy force during penetration is calculated in Equation 7.11 using the current

cavity geometry.

The resulting su profiles are compared with their piezoball counterparts in Figure 7.7

using operative rather than invert depth (White et al. 2010, Morton et al. 2014). The

Clyde piezoball data have been interpreted using the same shallow analysis described

for the dynamic data. The excellent agreement between the push-in and free-fall su

profiles is extremely encouraging and points both to the merit of the relatively simple

- 195 -

framework proposed, which is made possible by the simple sphere geometry, and also

to the potential of the instrumented free-fall sphere as an effective tool for

characterising the near-surface strength of soft seabeds.

Figure 7.7. Comparison of undrained shear strength profiles derived from free-fall

sphere acceleration data and push-in piezoball penetration resistance: (a) Erne

and (b) Clyde

(a)

(b)

- 196 -

7.9. Conclusions

Acceleration data, measured in field tests where a steel sphere was allowed to free-fall

through water and penetrate the underlying soft soil, have been used in the development

of a theoretical framework that describes the forces acting on a sphere during dynamic

embedment in soil. The framework is cast in terms of both fluid mechanics drag

resistance and geotechnical shear resistance, but formulated in terms of a single capacity

factor that approaches the conventional bearing capacity factor at low strain rates, but

may be up to two orders of magnitude higher when the sphere approaches its terminal

velocity. Experimental data gathered from 87 separate free-fall sphere tests at two sites

are interpreted. The merit of the framework was demonstrated through a simple inverse

analysis in which the undrained shear strength measured during free-fall sphere tests

was calculated from the measured acceleration data, accounting for buoyancy created by

the passage of the advancing sphere and a reduced bearing capacity factor at shallow

embedment. The resulting undrained shear strength profiles were shown to be in

excellent agreement with those derived from piezoball penetrometer tests, indicating the

potential of the free-fall sphere as a simple yet effective tool for characterising the near-

surface strength of soft seabeds.

- 197 -

CHAPTER 8. CENTRIFUGE MODELLING OF AN

INSTRUMENTED FREE-FALL SPHERE FOR

MEASUREMENT OF UNDRAINED STRENGTH IN

FINE-GRAINED SOILS

8.1. Abstract

This paper describes centrifuge tests in which a model free-fall sphere was allowed to

free-fall in water before dynamically embedding within reconstituted samples of kaolin

clay and two offshore natural clays. Instrumentation within the sphere measured

accelerations along three orthogonal axes. The resultant acceleration was used to

calculate sphere velocities and displacements. This allowed the penetration resistance

acting on the sphere to be expressed in terms of a single capacity factor that captures

soil resistance from both shearing and drag, and varies uniquely with the non-

Newtonian Reynolds number. Undrained shear strength profiles obtained from a simple

The final technical chapter extends the experience described in all previous technical

chapters by using the approaches developed in the thesis to further demonstrate the

merit of the IFFS through centrifuge experiments on kaolin clay and two natural soils

offshore soils.

Morton, J. P., O‟Loughlin, C. D. & White. D. J., 2015. Centrifuge modelling of an

instrumented free-fall sphere for measurement of undrained strength in fine grained

soils. Canadian Geotechnical Journal. DOI: 10.1139/cgj-2015-0242.

- 198 -

inverse analysis of the acceleration data show good agreement with those obtained

using conventional push-in penetrometer tests.

8.2. Introduction

Assessments of seabed strength for design are often based on interpretations of the

penetration resistance of a cone, T-bar or ball penetrometer as it is pushed into the soil.

However, this type of test requires a large seabed frame to provide reaction forces to

advance the penetrometer. A more cost effective alternative is to allow a penetrometer

to free-fall through the water column, so that the total energy of the penetrometer at the

seabed mudline provides the force to drive it into the seabed. Experience with such

systems has typically been with slender, shafted projectiles with a conical tip (i.e.

similar to a cone penetrometer). In some cases these devices have flukes near the rear of

the penetrometer for stability during free-fall (e.g. see Figure 8.1). Although the ease

and speed of installation make free-fall penetrometers an attractive option, their

adoption is hampered by difficulties in interpreting soil strength at the very high strain

rates that are prevalent during the dynamic penetration. These strain rates are up to eight

orders of magnitude higher than in a standard laboratory element test. Hydrodynamic

aspects further complicate the analysis. These include drag resistance that not only

occurs during free-fall in water, but also during dynamic embedment in soil, and the

potential for water to become entrained at the projectile-soil interface. The entrainment

of water has been indicated in numerical studies (Sabetamal et al. 2014) and free-fall

penetrometer and anchor field tests (Seifert et al. 2008; Jeanjean et al. 2012). These

complexities present challenges for free-falling penetrometers in firstly, isolating a

value of „dynamic strength‟ from the total resistance experienced by the projectile

during embedment and secondly, scaling this dynamic strength to a „static‟ value

appropriate for design.

- 199 -

In soft fine-grained soils, as an alternative to a free-falling CPT, a spherical free-fall

penetrometer can be used, of the form shown in Figure 8.2 (Morton and O‟Loughlin

2012; Morton et al. 2015). A spherical free-fall penetrometer is attractive as it benefits

from the advantages of full-flow penetrometers, which have more tightly bracketed

bearing capacity factors for deriving the undrained shear strength, su, from the net

penetration resistance (Chung and Randolph 2004). In the context of free-fall

penetrometers, the analysis for a sphere is simpler than for a cone as it does not include

the somewhat complicated frictional resistance that occurs along the shaft of a cone

penetrometer. This resistance is uncertain because it may be affected by water

entrainment (Richardson et al. 2009; Jeanjean et al. 2012) and has been shown to

exhibit different strain-rate dependency compared to the tip (Dayal et al. 1975; Steiner

et al. 2014; Chow et al. 2014). Finally, unlike a cone penetrometer, a sphere may freely

rotate without affecting the projected area and in turn the mobilised shear resistance.

- 200 -

Figure 8.1. Examples of free-fall shafted penetrometers with conical tips:(a) CPT-

Lance (courtesy of Dr. Nina Stark); (b) Nimrod (courtesy of Dr. Nina Stark); (c)

FFCPT (Furlong et al. 2006); (d) CPT Stinger (after Young et al. 2011)

Figure 8.2. Free-fall sphere ready for release (Morton and O’Loughlin 2012)

- 201 -

This paper examines the merit of a free-falling sphere to quantify near surface soil

strength using centrifuge data from instrumented free-fall sphere tests in three different

clays. A simple inverse analysis of the free-fall data - using a model that couples the

drag and strain rate-enhanced shear resistance in terms of a single capacity factor

framework - provides profiles of su with depth that are compared with equivalent

profiles obtained from push-in ball penetrometer tests.

- 202 -

8.3. Penetration of a sphere in soil

Figure 8.3. Forces acting on the sphere during penetration in soil

The forces acting on the IFFS during penetration in soil (i.e. after it impacts the

mudline) are shown schematically in Figure 8.3, which leads to the following equation

that governs the motion response in soil:

resistAMSS FFFma 8.1

where a is the linear acceleration (v is velocity and t is time), m is the mass of the

sphere, FSS is the submerged weight in soil. This is calculated as the dry weight of the

sphere minus the product of the soil unit weight, γ, and the volume of displaced soil,

Vdisp. This volume is that embedded in the soil, with an adjustment for any cavity that is

currently growing in the wake of the advancing sphere (Morton et al. 2014). FAM is the

added mass force, FAM = aCmmsoil, that occurs during non-stable flows when an object is

- 203 -

accelerating or decelerating (Lamb, 1932). The added mass (Cmmsoil) can be interpreted

as the mass of surrounding soil that is accelerated with the sphere, which results in a

force controlled by the added mass coefficient (Cm). The final term on the right hand

side of Equation 8.1 is a combined dynamic resistance force acting on the sphere. Fresist

is conventionally considered as the sum of the strain-rate enhanced bearing resistance,

, and the drag resistance, FD:

2

21 vACAsNFFF pDpopucDsresist opu

8.2

where Nc is a bearing capacity factor, Ap is the projected area of the sphere, su-op is the

mobilised, or operative shear strength at the strain rate relevant to the velocity v, CD is

an drag coefficient and ρ is the density of the soil. However, Sadhi et al. (2014) and

Morton et al. (2015) showed that these terms can be combined via a superposition

approach in which the strength and drag resistance components are expressed as a single

resistance force:

popuresist ANsF 8.3

where N is a geometry-dependant capacity factor that is dependent on the non-

Newtonian Reynolds number, Renon-Newtonian, (Zakeri et al. 2008):

opuNewtoniannon s

v

,

2

Re 8.4

- 204 -

The well know influence of strain rate, on the mobilised strength due to viscous

effects is described by a power law enhancement of the soil strength (Briaud et al.,

1984; Biscontin and Pestana, 2001; Peuchen and Mayne, 2007; Jeong et al., 2009;

Lehane et al. 2009; O‟Loughlin et al. 2013b):

refrefuopu ss

,, 8.5

where su,ref is the undrained shear strength at the reference strain rate ref and β is a strain

rate parameter defining the strength of the rate effect for any given soil. As shown by

Lehane et al. 2009, O‟Loughlin et al. (2013b) and Chow et al. (2014), the operational

shear strain rate around a penetrometer may be approximated by v/D, such that Equation

8.5 may be written as:

refrefuopu Dv

Dvss//

, 8.6

It follows from Equations 8.2, 8.3 and 8.4 that a relationship between N and Renon-

Newtonian may then be expressed as:

cNewtoniannonD NCN Re21 8.7

Morton et al. (2015) showed that this framework is capable of describing the resistance

acting on a sphere as measured in field tests with an IFFS. The model requires three

main input parameters: the drag coefficient, CD, for the sphere over the range of non-

Newtonian Reynolds numbers applicable to dynamic penetration in soil, the strain rate

- 205 -

parameter, β, and a reference strain rate, ref. The selection of these parameters are

considered later in the paper.

Equation 8.7 also requires an assessment of the bearing capacity factor, Nc, and its

evolution with depth, as for push-in penetrometer tests such as the CPT, T-bar or ball.

Hossain et al. (2005), White et al. (2010), Tho et al. (2012) and Morton et al. (2014)

showed that this evolution depends on the dimensionless soil strength ratio, su/γ'D,

where γ' is the effective unit weight of the soil and D is the diameter of the penetrating

body. The transition from a reduced Nc value at shallow penetration where the failure

mechanism extends to the soil surface, to the limiting value for deep embedment

associated with a full-flow mechanism, has recently been investigated for a sphere over

a wide range of su/γ'D (Morton et al. 2014). This approach, expressed by the following

power function, has also been shown to work well in two other natural soils (Morton et

al. 2015):

p

opdeep

opdeepbc w

wNN

ˆˆ

8.8

where the transitional depth at which Nc = Nc-deep is given by:

f

u

cu

opdeep

eD

s

adD

sbaw

/1

ˆ

8.9

and the fitting constants a = 16.3, b = 0.12, c = 1.3, d = 0.52, e = 4.9, f = 1.5 and p =

0.49.

- 206 -

The above framework may be used with vertical acceleration data, measured during

free-fall sphere tests, to obtain undrained shear strength profiles. Rearranging Equation

8.1, enables su to be calculated at any depth during penetration from

refp

soilmssu

DvDvNA

mcmaFs

//

8.10

The merit of Equation 8.10 in quantifying su is considered in the remainder of the paper,

through consideration of instrumented free-fall sphere centrifuge tests in kaolin clay,

Laminaria clay (from the Timor Sea, Erbrich and Hefer 2002) and West Africa clay

(from the Gulf of Angola).

8.4. Experimental details

8.4.1. Measurement technique

Estimation of su from free-fall sphere tests requires measurements of the acceleration in

the direction of motion. In equivalent field experiments reported by Morton and

O‟Loughlin (2012), O‟Loughlin et al. (2014) and Morton et al. (2015), an inertial

measurement unit (IMU) measured acceleration along three orthogonal axes and

rotation rates about the same three axes. Interpretation of the IMU data (Blake et al.

2015) required that the measurements were considered within a fixed frame of reference

to establish the projectile velocity and displacement along the projectile‟s direction of

motion. This approach becomes complex in a centrifuge environment as the gyroscope-

measured rotation rates include a component of the centrifuge rotation rate. However, if

the direction of motion can be maintained vertical, the interpretation can be reduced to

considering the acceleration along the depth (vertical) axis of a fixed reference frame.

This was achieved in the centrifuge tests considered here by allowing the sphere to drop

- 207 -

through a vertical installation guide located above the soil surface. The acceleration

along the vertical direction of motion can then be determined from the resultant

acceleration calculated from the acceleration components measured along the three

orthogonal axes shown in Figure 8.4a using a tri-axis accelerometer.

- 208 -

Figure 8.4. Model instrumented sphere shown: (a) during fabrication showing the

void in the sphere for the tri-axis MEMS accelerometer (b) after fabrication

alongside a centrifuge scale push-in ball penetrometer

- 209 -

8.4.2. Instrumented free-fall sphere

The 20 mm diameter stainless steel sphere used for the free-fall centrifuge experiments

(see Figure 8.4) had a mass of 31 grams. The centrifuge tests were conducted at 12.5 g,

such that the centrifuge model IFFS is equivalent to the 250 mm IFFS used in the field

trials referred to previously and shown in Figure 8.3. The model sphere was

instrumented with a tri-axis ADXL377 +/- 200 g MEMS accelerometer, measuring

approximately 5 × 5 × 2 mm. A commercially-available ±200 g accelerometer was

chosen as the only device offering three-axis sensing within the size constraints, and

providing an analogue output. The sensor was located within a cylindrical void in the

sphere (see Figure 8.4a) that was subsequently filled with epoxy. The accelerometer was

aligned with the body frame of the sphere as shown in Figure 8.4a. The body frame is a

reference frame with three orthogonal axes xb, yb and zb. Before release in the centrifuge

tests, the sphere was oriented such that the zb axis was approximately parallel to the

centripetal acceleration vector in the centrifuge. The accelerometer measures

accelerations Abx, Aby and Abz along the three orthogonal axes of the body reference

frame. As discussed previously, these data are used to calculate the resultant

acceleration, Ar:

222bzbybxr AAAA 8.11

The linear acceleration, a, of the sphere then becomes:

rAa r2 8.12

where ω is the angular velocity of the centrifuge and ω2r is the centripetal acceleration

level at a distance, r, from the centrifuge‟s axis of rotation to the sphere location.

- 210 -

8.4.3. Soil preparation technique

The offshore clay samples were reconstituted by adding minimal amounts of water to

reconstitute the soils as slurries and the kaolin clay sample was prepared by mixing

kaolin powder with water to form a slurry at a moisture content of 120%. The slurries

were then mixed continuously for about 24 hours and then transferred to sample boxes,

each with internal dimensions 598 mm long, 117 mm wide and 300 mm deep. These

three sample boxes were then nested side by side in a larger strongbox to allow „in-

flight‟ consolidation of the three samples at the same time. The acceleration field acts

radially in the centrifuge so the testing was performed only with the sample located

along the longitudinal centreline of the beam centrifuge strongbox. Using three

narrower internal sample boxes facilitated a greater test plan area by interchanging the

sample boxes to put the „test box‟ in the middle.

A 10 mm layer of sand was placed at the base of each box (i.e. before pouring the

slurry) and the end walls of each box were then lined with a drainage blanket to ensure

that there was no hydraulic gradient between the top and base of the sample. The

samples were normally consolidated under self-weight consolidation in the centrifuge at

12.5 g for approximately 14 days, during which time a 30 mm water layer was

maintained at the sample surface to ensure saturation. After consolidation, the surface of

each soil sample was scraped to create a slightly overconsolidated soil with a level

surface and final sample heights of 225 mm for the kaolin and West Africa Clay and

160 mm for the Laminaria clay.

8.4.4. Centrifuge test details and procedures

The centrifuge test programme included a combination of free-fall sphere tests and

constant rate of penetration ball and T-bar penetrometer tests to determine the reference

- 211 -

undrained shear strength. A minimum of 3 sphere diameters was allowed between rigid

sample box walls and adjacent test sites to minimise potential boundary effects.

Figure 8.5. Experimental arrangement for the push-in ball and instrumented free-

fall sphere tests

- 212 -

The experimental arrangement is shown in Figure 8.5. The vertical installation guide

was located along the centreline of the strongbox with the exit located 2 mm above the

sample surface. Prior to each free-fall test, the sphere was located in the installation

guide at a height between 50 and 110 mm above the soil surface and the water level was

increased to ensure that the sphere was fully submerged at the release height. A release

cord attached to the top of the sphere kept the sphere at the preselected release height in

the guide. The centrifuge was then spun up to 12.5 g, and after a reconsolidation period

of one hour, the sphere was released in-flight by supplying current to a resistor causing

it to heat and burn through the release cord. The data during the release and embedment

were acquired at 50 kHz using the high speed logging mode of the data acquisition

system (Gaudin et al. 2009).

8.5. Test results and discussion

8.5.1. Penetrometer tests and soil properties

Strength characterisation tests were conducted in-flight, using a ball penetrometer 18

mm in diameter located at the end of a 5 mm diameter shaft. The measured resistance

was corrected for the unequal area due to the presence of the shaft. This correction was

almost negligible for the ball penetrometer tests due to the very low probe/shaft area

ratio (0.07). The reference ball penetrometer tests were conducted at a penetration rate

of 0.3 mm/s in kaolin and 0.25 mm/s in the natural soils. This gave a dimensionless

velocity, V = vd/cv 140 (where v is the velocity, d is the ball diameter and cv is the

vertical coefficient of consolidation in each soil type) ensuring that the response was

undrained (Finnie and Randolph 1994; House et al. 2001; Chung et al. 2006). The

values of cv were based on Rowe Cell data at a vertical effective stress, σ'v = 6 kPa

(Richardson et al. 2009; Bienen et al. 2015; White, 2015), which is the average of the

- 213 -

range of σ'v of interest in the free-fall tests (i.e. the vertical effective stress at half of the

mean embedment depth achieved in the free-fall sphere tests). This gave cv = 1.23 m2/yr

for kaolin and 1 m2/yr for the Laminaria and West Africa clays. The plastic limit (PL)

and liquid limit (LL) for each soil are: Laminaria clay PL = 41% and LL = 89%, West

Africa clay PL = 50% and LL = 165% and kaolin clay PL = 27% and LL = 61%.

The undrained shear strength, su, was derived from the net penetration resistance, qnet

using the commonly adopted bearing factor Nc-deep = 10.5 once a full-flow mechanism

was established. The net resistance is the measured penetration resistance corrected for

pore pressure on the ball shoulder (at the u2 position) and overburden pressure effects

(Chung and Randolph 2004), with an unequal area ratio of 0.08. The full-flow

mechanism was observed when video capture of the penetrometer tests showed that the

conical cavity created during shallow penetration closed over the top of the ball. In the

kaolin, Laminaria and the West Africa soils, a normalised cavity depth of 0, 1.7 and 2

was observed, corresponding to a of 0, 2.2 and 2.5 respectively. At shallower

penetration depths, the net penetration resistance was interpreted within the shallow

penetration framework proposed by Morton et al. (2014). The resulting su profiles are

shown in Figure 8.6a using operative rather than invert depth (White et al. 2010;

Morton et al. 2014). The increase in su due to the soil scrape may be described using the

well-known dependence of soil strength on overconsolidation ratio (Ladd et al. 1977):

m

ncv

uvu OCRss

8.13

where σ'v is the current vertical effective stress, determined from the γ' profile with

depth (shown alongside the moisture content (w) profiles with depth in Figure 8.6b) and

- 214 -

the slightly varying acceleration level with radius within the centrifuge (12.5 to 13.7 g

over the depth of penetration in soil), m is the plastic volumetric strain ratio (Schofield

and Wroth 1968) taken as m = 0.8.

- 215 -

Figure 8.6. Profiles of: (a) undrained shear strength with depth (from ball

penetrometer tests) and (b) moisture content and effective unit weight with depth

established from post-testing sample cores

- 216 -

The undrained strength ratio, su/σ'v, for each soil type is also shown on Figure 8.6a. For

kaolin clay su/σ'v = 0.16, equivalent to an undrained strength gradient with prototype

depth, k = 1 kPa/m, which is typical for normally consolidated centrifuge samples of

UWA kaolin clay (Chow et al., 2014; Morton et al., 2014; Hu et al., 2014). The

Laminaria and West Africa su profiles can both be represented by su/σ'v = 0.35,

equivalent to k = 2.2 and 1.1 kPa/m for Laminaria and West Africa clay respectively

owing to the different effective unit weight profile for the two soils (see Figure 8.6b).

8.5.2. Free-fall tests

A typical acceleration time history from a IFFS test in kaolin clay is shown in Figure

8.7 together with the resulting velocities and displacements established from numerical

integration of the acceleration data.

The resultant acceleration, Ar, contains a component of the centripetal acceleration at

the current sphere location. Ar is therefore ~11.9 g before release and slightly higher at

~13 g after the sphere has come to rest in the soil at a larger radius from the centrifuge

axis. These acceleration levels are consistent with the centripetal acceleration calculated

at the sphere release height and embedment depth. The velocity and displacement data

were obtained by numerically integrating the linear acceleration of the sphere, a (given

by Equation 8.12), relative to the soil surface, once for velocity and twice for

displacement. Following the logic outlined by O‟Loughlin et al. (2014), the integration

sequence has been performed in reverse from the end of the test when the sphere is

embedded in the soil and the linear acceleration is zero (O‟Loughlin et al. 2014). The

acceleration trace on Figure 8.7 is characterised by a reduction in acceleration at the

release height as the sphere begins to „free-fall‟. The acceleration does not reach zero

- 217 -

due to frictional resistance between the sphere and the walls of the guide, and drag

resistance on the sphere as it travels through the water. Oscillations in the acceleration

trace during the free-fall stage are attributed to sporadic knocking of the sphere against

the guide walls. At approximately t = 0.05 s the sphere impacts the soil with a velocity,

v = 2.8 m/s. The low resistance afforded by the soil at shallow embedment allows the

sphere to continue to accelerate, reaching a peak velocity, v = 3.4 m/s at t = 0.065 s and

a sphere invert embedment, z = 48 mm (measured from the soil surface). The sphere

comes to rest in the soil at t = 0.118 s, evident as when the acceleration returns to the

centripetal acceleration, which causes the velocity to become zero and the displacement

to reach a constant value. The point of impact with the soil surface can be difficult to

establish from the acceleration trace alone, particularly for impact with soft soils. As

outlined by O‟Loughlin et al. (2014), integrating linear acceleration in reverse from the

end of the test allows the velocity and displacement to be calculated incrementally.

Impact with the soil surface can then be established as the point at which the calculated

displacement equals the direct measurement of the embedment depth (obtained after the

test when the centrifuge was at rest). In this example the final sphere embedment depth

was 128 mm, equivalent to over 6 diameters and the sphere orientation changed 4° and

15° in the Xb and Yb axes (i.e. perpendicular to gravity) as it penetrated the water

column and soil. This level of sphere spin indicates the sphere spin during soil

penetration is not significant.

The acceleration spike evident in Figure 8.7 at t = 0.055 s, just after impact with the soil

surface, is noteworthy. This is thought to be the compression wave caused from the

initial impact, travelling back past the sphere after reflection from the rigid sample

boundary, an artefact of conducting the tests in a laboratory that would not be present in

the field. Consideration of the time from initial impact to the apparent point of reflection

- 218 -

on Figure 8.7, and the depth from the soil surface to the base of the sample, indicate that

the wave is travelling at approximately 100 m/s, which is consistent with the speed of P

wave propagation in soft soils. Similar observations have been made from 1 g free-fall

penetrometer tests in uniform strength kaolin clay (Chow, 2012). Although the reflected

wave does not appear to significantly affect the calculated velocity or displacement, it

manifests as an increase in resistance and has therefore been excluded in the subsequent

interpretation.

Figure 8.7. Accelerometer, velocity and displacement traces in a typical free-fall

sphere centrifuge test in kaolin clay

Figure 8.8 shows time histories of the acceleration, velocity and displacement for

typical tests in each soil type. Unlike Figure 8.7 in which the acceleration includes both

the centripetal acceleration and the acceleration associated with the motion of the

sphere, Figure 8.8 plots linear acceleration, a. Determined using Equation 8.12, a, is the

sphere acceleration relative to the centripetal acceleration at the current sphere position

within the rotating frame of reference. Therefore, the sphere acceleration is zero before

- 219 -

release, changing abruptly to about 8g immediately after release (consistent with the

centripetal acceleration at the sphere release height) and zero when the sphere has come

to rest in the soil.

Velocity and displacement traces derived from the linear acceleration data are provided

in Figure 8.8b and Figure 8.8c respectively.

- 220 -

Figure 8.8. Example time histories of: (a)linear acceleration, (b) velocity and (c)

displacement for the three soil types

- 221 -

As discussed earlier in the paper, the measurement approach is based on the vertical

trajectory followed by the sphere during free-fall in water and embedment in soil. As

shown by Figure 8.5, the vertical installation guide located above the centreline of the

soil sample ensured that the sphere fell vertically in water. The guide applied the

tangential force needed to keep the sphere rotating at the same angular velocity as the

centrifuge, which requires an increasing tangential velocity with increasing radius.

Once the sphere is embedded in the soil, this force must be applied by the soil and there

may be a tendency for the sphere to follow a curved trajectory through the soil,

particularly for softer soil (O‟Loughlin et al. 2014). This was checked in a number of

tests by sectioning the soil sample after the test to reveal the final location of the sphere

relative to the point of impact. An example (from a test in Laminaria clay) illustrating

this process is provided in Figure 8.9, where it can be seen that the lateral displacement

of the sphere was approximately 4 mm. For a typical sphere embedment of 100 mm (5

sphere diameters), a lateral displacement of 4 mm increases the distance travelled in the

soil by less than 0.1 mm (0.1% error) compared with the vertical penetration depth. This

is sufficiently small to be considered negligible and ignored, effectively validating the

measurement approach.

- 222 -

Figure 8.9. Post-test analysis of the sphere trajectory and measurement of the final

embedment depth

8.5.3. Interpretation of free-fall acceleration data

The acceleration data are now considered within the simple framework outlined earlier

in the paper that accounts for both geotechnical shearing resistance and fluid mechanics

drag resistance. Rearranging Equation 8.7 allows N to be established as a function of

Renon-Newtonian. This is shown on Figure 8.9a, b and c for the Laminaria, West Africa and

kaolin clays respectively. The highest values of Renon-Newtonian correspond with initial

penetration, where the soil strength is low and the sphere velocity is high. Differences in

the maximum Renon-Newtonian values in Figure 8.9 reflects differences in strength and

impact velocities between tests in each soil. At these high values of Renon-Newtonian, N is

typically between 20 and 300, 2 to 30 times higher than the (geotechnical) bearing

capacity factor, Nc. As penetration progresses, Renon-Newtonian reduces as the soil strength

increases (as reflected in Figure 8.6a) and the sphere velocity reduces. This has the

effect of decreasing the dynamic resistance component, resulting in reducing values of

- 223 -

N. At deeper embedments Renon-Newtonian reduces to between 5 and 10. At depth N

approaches the limiting value of Nc = 10.5 that was used at deep embedments to

interpret the undrained shear strength from the net penetration resistance measured in

ball penetrometer tests.

The data on Figure 8.10 require an assessment of the added mass force, FAM, and the

operative strength, su,op. FAM was calculated using Cm = 0.5, as established both

theoretically and experimentally for a sphere (Sumer and Fredsoe 1997; Pantaleone and

Messer, 2011). The operative strength, su,op, which is reflected in both the vertical and

horizontal axes of Figure 8.10, was calculated from Equation 8.5 using β = 0.07, which

approximates to an 18% change in soil strength per log cycle change in strain rate.

- 224 -

Figure 8.10. Relationship between N and Renon-Newtonian for a sphere in the three soil

types: (a) Laminaria soil (b) West Africa clay (c) kaolin

- 225 -

Also shown on Figure 8.10 are data from equivalent field experiments using the 0.25 m

diameter IFFS shown in Figure 8.2 in a soft lakebed (Morton et al. 2015) and equivalent

data for a cylinder from experiments in a flume (Zakeri et al. 2008; Zakeri et al. 2011)

and in a centrifuge (Sahdi et al. 2014). Collectively the data indicate that the threshold

Renon-Newtonian at which drag resistance dominates over soil strength is between 3 and 10,

similar to the range reported by Zhu and Randolph (2011), Randolph and White (2012)

and Sahdi et al. (2014) for debris flow impacting on submarine piplelines. As indicated

through the annotations on Figure 8.10, drag resistance dominates (at Renon-Newtonian > 3

to 10) during shallow penetration at normalised embedment depths, d/D < 1, at

combinations of high velocities (v = 2.5 to 3.5 m/s) and low operative strengths (su,op = 0

to 0.5 kPa). The centrifuge data can be described using Equation 8.7, which is also

shown on Figure 8.10, using Nc = 10.5 and CD = 0.26 as established from IFFS field

tests reported by Morton et al. (2015). The centrifuge data eventually reach N = 10.5 at

low ReNon-Newtonian, which is higher than N = 8.5 for the field data as this number was

adopted in the interpretation of the piezoball data used as the measurement of su.

Within Figure 8.10, there is a small region (10 < Renon-Newtonian < 30) that exhibits some

lower values of N. These correspond to very shallow embedments at which the bearing

factor is reduced due to near-surface effects.

Figure 8.10 also indicates the post-analysed normalised cavity height. In the kaolin clay

sample the cavity depth was calculated as 0.1 diameters, such that soil flow around the

advancing sphere occurred almost immediately, at very shallow depth. This is consistent

with the very low dimensionless strength ratio, su/γ'D = 0.17 at the transitional depth for

kaolin clay, which leads to a calculated cavity depth, d/D = 0.1 using Equation 8.9. In

the Laminaria and West Africa clays the dimensionless strength ratios at the transition

depth are similar - su/γ'D = 0.97 and 0.94 respectively. Although the undrained strength

- 226 -

profiles for the Laminaria and West Africa clays differs by a factor of approximately 2,

the lower effective unit weight of the Laminaria clay results in almost identical

dimensionless strength ratios for the two soils. Although this results in deeper cavities

for the tests in Laminaria and West Africa clays, calculated using Equation 8.9 as

= 2.0 and 2.3 for Laminaria and West Africa clay respectively. In both soils

the sphere penetrated beyond the cavity depth such that a full-flow failure mechanism

was established.

The merit of the framework is now explored by using the linear, acceleration described

in Equation 8.12 to obtain su profiles for each soil, based on the dynamic free fall data.

In kaolin, a deeply embedment soil flow mechanism is assumed, such that the

submerged weight of the sphere, FSS, is the weight of the sphere in water minus the

effective weight of the displaced soil, which may be calculated using the volume of the

sphere. The capacity factor, N, may then be determined using Equation 8.7 with Nc =

10.5 and CD = 0.26. As discussed in the previous section, su/γ'D ~ 0.95 is higher in the

West Africa and Laminaria clays at the final sphere embedment depth, which is

sufficiently high that both the variation in Nc with embedment depth and the soil

buoyancy associated with the cavity formed by the penetrating sphere require

consideration (Morton et al. 2014). The transitional embedment depth at which Nc = Nc-

deep calculated using Equation 8.9 compared well to the post-analysed which

were 2.2 and 2.5 for Laminaria and West Africa clay respectively and were within 10%

of the calculated measurements. These transitional embedment depths allow the

variation in Nc with depth to be determined using Equation 8.8.

The resulting su profiles obtained from Equation 8.10 are compared with the push-in ball

penetrometer profiles in Figure 8.11 using operative rather than invert depth (White et

al. 2010; Morton et al. 2014). Shallow embedment effects have been considered in the

- 227 -

same way for the push-in tests as the free-fall tests. Interruptions to the su profiles on

Figure 8.11 correspond with the acceleration spikes observed in Figure 8.8 that are

attributed to compression wave reflections and have been removed before conducting

the inverse analysis. Figure 8.11 shows good agreement between the push-in and free-

fall su profiles, despite some disparity at very shallow embedment. The level of

agreement observed on Figure 8.8 confirms the merit of the interpretation framework,

which is made possible by the simple sphere geometry, and also to the potential of the

free-fall sphere to be used as an effective tool for characterising the near-surface

strength of soft seabeds. As discussed previously, the framework extends the

interpretation of a push-in ball penetrometer test using only two additional geotechnical

parameters, CD and β.

The value of CD for a bluff body depends on the flow regime and the object geometry,

as is well established in fluid mechanics. For a sphere, CD does not vary significantly

over the range of non-Newtonian Reynolds numbers in Figure 8.10 where drag is

significant (Schlichting et al. 2000). The consistent scatter around the trend line in

Figure 8.10 indicates that a more refined approach with CD varying with Renon-Newtonian

would not capture the data set any better.

The range of β reported in the literature is quite wide, spanning β = 0.05 to 0.17 (Jeong

et al. 2009), and a choice of is required in the interpretation of dynamic penetration

problems. The wide reported range is partly due to different strain rate ranges being

considered and a variety of test conditions, both in situ and in a laboratory. For instance,

laboratory element tests involving variable strain rates tend to be at relatively low strain

rates (typically 1%/h or 2.8 × 10-6 s-1; Boukpeti et al. 2012) and measure intact strength

with no strength reduction from softening. In contrast, variable rate full-flow

penetrometer tests usually involve strain rates in the range 0.2 to 20 s-1 and include

- 228 -

compensating effects of strain softening that limit the strength increases associated with

increasing strain rates. IFFS tests include a similar degree of strain softening to a

variable rate full-flow penetrometer test. They also only involve slightly greater strain

rates because the increased velocity is partly compensated by the larger diameter

(maximum v/D = 31 and 130 s-1 for field and centrifuge free fall ball tests respectively).

Therefore, β values for an IFFS test should be guided by those measured in variable rate

full-flow penetrometer tests, which typically give β = 0.05 to 0.09 (Low et al. 2008;

Chung et al. 2006). The effect of adjusting β over this range is explored in Figure 8.12

for Laminaria clay. It is apparent that the lower bound β = 0.05 corresponds with the

largest departure from the base case profile established using β = 0.07, although by less

than 20%. This uncertainty is comparable to the differences that are commonly linked to

uncertainty between different laboratory strength tests (Bienen et al. 2010).

- 229 -

Figure 8.11. su profiles from free-fall sphere and push-in penetrometer tests in: (a)

Laminaria soil (b) West Africa soil (c) kaolin

- 230 -

Figure 8.12. Effect of varying β parameter on free-fall sphere su profile

8.6. Conclusion

This paper has presented centrifuge tests undertaken to investigate the potential for a

new in situ tool – the Instrumented Free-Fall Sphere (IFFS) – to measure the undrained

shear strength profile of soft soil. The centrifuge experiments, which were conducted in

kaolin clay and two reconstituted offshore clays, from Laminaria (Timor Sea) and West

Africa (Gulf of Guinea) clay gave sphere embedments that were up to 5.5 diameters,

which would provide strength measurements over the upper 1.5 m of the seabed for a

practically-dimensioned field tool.

Interpretation of sphere acceleration measurements to quantify the soil strength relied

on a framework cast in terms of both drag resistance and geotechnical shear resistance,

but formulated in terms of a single capacity factor. This capacity factor reduces to the

- 231 -

conventional geotechnical bearing capacity factor at low strain rates, but has been

shown in these experiments to be over an order of magnitude larger when the sphere is

penetrating at high velocities.

The merit of the simple framework was demonstrated through an inverse analysis in

which the undrained shear strength was calculated from the measured acceleration data,

accounting for buoyancy created by the passage of the advancing sphere and a reduced

bearing capacity factor at shallow embedment. Good agreement was obtained between

the dynamically measured su profile and the su profile measured during push-in ball

penetrometer tests. The level of agreement shown, and the relative insensitivity to the

choice of strain rate parameter, suggests that the IFFS is an effective tool for

characterising the near-surface strength of soft seabeds. This is in part due to the fact

that the IFFS, unlike a slender FFP, requires no correction for shaft effects and the

associated strain rate difficulties. Given the positive research findings, it is desired that

future experiments would extend the validation of dynamic versus static su data to a

wider range of soils and focus on practical aspects of deployment and recovery of the

device. Another option that is of interest is the optimization of size and weight of the

IFFS for various soils types, potentially creating an „IFFS family‟ for different expected

soil conditions.

- 232 -

CHAPTER 9. CONCLUSIONS

9.1. Summary

This thesis has introduced a new free-falling penetrometer as a cost effective and

accurate in situ site investigation tool. The new penetrometer is spherical in design and

comprises instrumentation to capture the motion history of the sphere during free-fall in

water and penetration in soil. The Instrumented Free-Fall Sphere (IFFS) represents a

step-change in free-falling penetrometer design, departing from the traditional slender

full-shafted projectile geometry. The IFFS combines the benefits of free-falling

penetrometers with the advantages of full-flow penetrometers which can more

accurately derive the undrained shear strength, su. This is primarily because: (a) it is

unnecessary to eliminate or correct for the shaft resistance in order to directly infer the

soil strength from the bearing resistance, and (b) there is greater uncertainty associated

with shaft resistance rate effects than bearing resistance. It is envisaged that significant

cost savings could be realised by utilising the IFFS in conjunction with other projects,

such as geophysical surveys that typically occur earlier in the site investigation

campaign. This is made possible due to the ease and speed of operation and the

potential to deploy the IFFS from smaller vessels, relative to the conventional approach

of using push-in penetrometers that require larger vessels and seabed frames.

The thesis primarily focused on assessing the viability of the new penetrometer design

by carrying out an experimental study through field tests and centrifuge experiments.

The field tests were conducted on a 0.25 m diameter IFFS in two locations: (i) an inland

- 233 -

lake, Lower Lough Erne in the Northwest of Ireland and (ii) an offshore site in the Firth

of Clyde which is located off the coast of Scotland in the Irish Sea. The sphere was

instrumented with a custom-made, low cost, 6 degree of freedom (6DoF) inertial

measurement unit (IMU) to measure the sphere acceleration and associated velocity and

displacement. The initial field tests provided the „proof of concept‟ and provided a basis

for understanding the behaviour during free-fall in water and dynamic embedment in

soil.

The centrifuge experiments were carried out in two phases - the first focussing on the

shallow penetration effects of an 11.3 mm diameter ball penetrometer (investigated in

Chapter 4). Shallow penetration effects alter the relationship between penetration

resistance and undrained shear strength. If ignored, they can lead to an under-estimation

of the undrained shear strength in the near surface soil. A series of formulations were

proposed to determine more accurately the undrained shear strength of near surface soil

over the range of dimensionless strength ratios that are of interest to offshore

applications.

The second phase of centrifuge testing focussed on the dynamic embedment of 20 mm

diameter model IFFS in three different soils: (i) Laminaria soil recovered from the

Timor Sea, (ii) West Africa clay recovered from the Gulf of Angola and (iii) kaolin

clay. The dynamic centrifuge experiments were carried out on a 1:12.5 reduced scale

model of the IFFS employed in the field experiments. The experiments used a new

theoretical framework to measure the undrained shear strength. The framework

considered the geotechnical strain rate enhanced shear resistance and the fluid

mechanics drag resistance, but couples both components in a single capacity factor. This

capacity factor approaches the conventional geotechnical bearing capacity factor at low

strain rates, but may be up to two orders of magnitude higher when the sphere

- 234 -

approaches its terminal velocity. The resulting undrained shear strength profiles were

shown to be in excellent agreement with those derived from piezoball penetrometer

tests, indicating the potential of the IFFS as a simple yet effective tool for characterising

the near-surface strength of soft seabeds.

9.2. Main findings

9.2.1. Free-fall and dynamic embedment in soil

This thesis has made a particular contribution in the area of dynamic penetration in soft

soils. Although a relatively large body of work has been carried out on numerous FFPs,

the investigation of high-quality FFP field and centrifuge experiments are rare. The

main findings from the dynamic embedment of the sphere in soil are summarised here:

A final embedment depth of over 9.5 sphere diameters (2.373 m) was achieved

in the Lough Erne site and 3.13 sphere diameters (0.782 m) in the Firth of

Clyde. The embedment depth was verified at each test site with an underwater

camera which observed markings on the retrieval line at the mudline. These

embedment depths exceed the depth of interest for geotechnical design of

shallow subsea infrastructure, which is typically 0.5 m.

In the centrifuge experiments, the sphere impacted the soil surface at velocities

up to 3.3 m/s. This resulted in embedment depths in the range of 3–7 sphere

diameters (60–128 mm) depending on the soil strength and impact velocity.

The assessment of the dynamic resistant forces acting on the sphere led to an

estimation of the strain rate parameter. The appropriateness of the strain rate

parameter was demonstrated through a parametric study by varying β within the

typical range reported from variable rate penetrometer tests (β = 0.05 to 0.09).

- 235 -

Higher strain rates were achieved in the centrifuge tests, with a maximum v/D =

130 s-1 compared to a maximum v/D = 31 s-1 in the field. In both sets of

experiments, a best-fit strain rate parameter of β = 0.07 was calculated.

Analysis of the motion data recorded by the IMU during dynamic embedment

in soil led to the development of a new theoretical framework that describes the

forces acting on a IFFS during dynamic embedment in soil. The framework is

cast in terms of both fluid mechanics drag resistance and geotechnical shear

resistance, but formulated in terms of a single capacity factor. It was found that

for low non-Newtonian Reynolds Numbers, Renon-Newtonian, the capacity factor

approaches the conventional bearing capacity factor, but may be up to two

orders of magnitude higher at high Renon-Newtonian.

The theoretical framework was used to calculate the undrained shear strength in

5 soils, 2 field soils (Chapter 7) and 3 soils in the centrifuge (Chapter 8). The

calculated su profiles in each soil compared very well to the undrained shear

strength profile measured in push-in ball or piezoball penetrometer tests,

verifying the merit of the IFFS as a time-effective and reliable site investigation

tool for soft soils.

9.2.2. MEMS accelerometers in the centrifuge

The thesis has made an original contribution to the area of geotechnical centrifuge

testing using MEMS accelerometers. The thesis described in detail two measurement

systems based on a single-axis and triple-axis accelerometer. To the authors knowledge

these are first reported experiments using MEMS accelerometers to measure the

dynamic forces on a FFP in a geotechnical centrifuge. The initial experiments involved

a comparison between a MEMS and piezoelectric single-axis accelerometer measuring

the motion response of a dynamically installed anchor as it free-falls through air and

- 236 -

embeds within a centrifuge soil sample (described in Chapter 6). The main findings

from the direct comparison between the two accelerometers are summarised here:

It was found that the performance of the MEMS accelerometer was superior to

the piezoelectric accelerometer during the free-fall phase and similar in

performance during the embedment phase in soil.

An assessment of the anchor tilt angle was explored in Chapter 6 and it was

shown that tilt angles up to 10˚ have negligible effect on the velocity profile,

with a maximum reduction in impact velocity and final embedment depth of

1.4% and 1% respectively. For a tilt angle of 20˚, the impact velocity reduces

by 5.4% and the final embedment depth 3.7% and for a tilt angle of 30˚, the

impact velocity reduces by 11.8% and final embedment depth reduces by 8.1%.

The second major contribution using a MEMS accelerometer was described in Chapter

8 where the vertical penetrating resistance on a model IFFS in the centrifuge was

measured using a triple-axis accelerometer. The tri-axis accelerometer was required

because the sphere can undergo excessive tilt whilst in free-fall in water and embedment

in soil. The method used the resultant acceleration computed from the three component

accelerations to calculate the vertical penetration resistance. The method was shown to

be able to accurately capture the forces acting on the sphere during penetration in soil.

9.2.3. Shallow penetration framework

The shallow penetration effects of a ball penetrometer were assessed through a suite of

centrifuge tests (described in Chapter 4) and a series of equations have been proposed

that offer a more rigorous and reliable means of assessing soil strength in the upper few

metres of the seabed. The tests were carried out with an 11.3 mm diameter ball

penetrometer penetrating kaolin clay under undrained conditions over a range of

- 237 -

normalised strength ratios, su/γ'D = 0.07 to 2.7. The tests aimed to capture the influence

of two mechanisms that are often overlooked: (i) the varying soil buoyancy with

penetration depth and (ii) the reduced bearing factor, Nc-shallow, arising from the shallow

failure mechanism. It was shown that the shallow failure mechanism can significantly

influence the estimation of su and can extend several diameters from the mudline. For

example, using a 0.25 m diameter IFFS, the correction can extend to a depth of up to ~1

m and for a push-in ball penetrometer test using a 113 mm diameter ball, the shallow

correction may extend to a depth of up to ~0.5 m. The framework comprises a series of

simple formulae that are discussed briefly here:

A correlation for the transition depth of a ball penetrometer, which is the depth

at which soil flows over the top of the ball, has been derived. The transition

depth indicates the depth where the soil flow mechanism changes from a

shallow heave-dominated flow to a deep flow round failure mechanism. The

original centrifuge data (Chapter 4) were combined with reinterpreted data from

LDFE analyses to form a unique relationship between the transition depth and

the normalised strength ratio over the range su/γ'D ≈ 0.07 to 40.

A correlation for the shallow bearing capacity factor has been derived; this

formulation accounts for the varying bearing capacity factor from a low number

at the soil surface that gradually increases to the limiting bearing capacity

factor, Nc-deep at the transition depth. The bearing factor component of the

correction is particularly important in stiff soils, due to the deeper penetration

required to mobilise the deep flow round failure mechanism.

A novel analytical closed-form solution for the varying soil buoyancy force

with depth has been proposed (Appendix 1). The formulation assesses the soil

buoyancy in the case where an open conical hole develops due to the passage of

- 238 -

the sphere. This is an important consideration in very soft soil with zero

mudline strength where the buoyancy resistance increases to almost 70% of the

geotechnical resistance during shallow penetration, and is independent of the

penetrometer diameter at all penetration depths.

- 239 -

REFERENCES

Abelev, A., Tubbs, K. & Valent, P., 2009b. Numerical investigation of dynamic free-

fall penetrometers in soft cohesive marine sediments using a finite difference

approach. Proceedings of OCEANS 2009: Marine Technology for Our Future:

Global and Local Challenges, MTS/IEEE Biloxi, Biloxi, MS, pp. 1–10.

Abelev, A. & Valent, P., 2009. Strain-rate dependency of strength and viscosity of soft

marine deposits of the Gulf of Mexico. Proceedings of OCEANS 2009: Marine

Technology for Our Future: Global and Local Challenges, MTS/IEEE Biloxi,

Biloxi, MS, pp. 1–9.

Achenbach, E., 1971. Influence of surface roughness on the cross-flow around a circular

cylinder. Journal of Fluid Mechanics, 46 pp. 321–335

Achenbach, E., 1972. Experiments on the flow past spheres at very high Reynolds

numbers. Journal of Fluid Mechanics. 54, pp. 565–575.

Akal, T. & Stoll, R. D., 1995. An expendable penetrometer for rapid assessment of

seafloor parameters. Proceedings of OCEANS 1995: Challenges of Our

Changing Global Environment, MTS/IEEE, San Diego, CA, pp. 1822–1826.

Allen, H. S., 1900. The motion of a sphere in a viscous fluid. Philosophical Magazine,

50, pp. 519–534.

Allmond, J. D., Hakhamaneshi, M., Wilson, D. W., & Kutter, B. L., 2014, Advances in

measuring rotation with MEMS accelerometers. Proceedings of the Eight

- 240 -

International Conference on Physical Modelling in Geotechnics, Perth,

Australia, 14–17 January 2014. 1, pp. 353–359.

Andersen, K. H., Murff, J. D., Randolph, M. F., Cluckey, E. C., Erbrich, C. T., Jostad,

H. P., Hansen, B., Aubeny, C., Sharma, P. & Supachawarote, C., 2005. Suction

anchors for deepwater applications. Proceedings of the 1st International

Symposium on Frontiers in Geotechnics, Perth, Australia, pp. 3–30.

Aubeny, C. P. & Shi, H., 2006. Interpretation of impact penetration measurements in

soft clays. Journal of Geotechnical and Geoenvironmental Engineering, ASCE,

132(6), pp. 770–777.

Audibert, J. M. E., Movant, M. N., Won, J. Y, & Gilbert, R. B., 2006. Toprepdo piles:

laboratory and field research. Proceedings of the 16th International Offshore

and Polar Engineering Conference, San Francisco, California, USA, pp. 135–

142.

Bacon, D. L. & Reid, E. G., 1924. The resistance of spheres in wind tunnels and in air.

Nut. Adv. Corn. Aero. Report. No. 185.

Bakić, V., 2004. Experimental investigation of a flow around a sphere. Thermal

Science, 8(1), pp. 63–81.

Balderas Meca, J., 2004. Rate effects of rapid loading in clay soils. PhD thesis, The

University of Sheffield.

Baligh, M.M., Azzouz, A.S., Wissa, A.Z.E., Martin, R.T. & Morrison, M.H., 1981. The

piezocone penetrometer: cone penetration testing and experience. Proc. ASCE

Conference on Cone Penetration Testing, St Louis, pp. 247–263.

- 241 -

Beard, K. V. & Pruppacher, H. R., 1969. A determination of the terminal velocity and

drag of small water drops by means of a wind tunnel. Journal of Atmospheric

Sciences, 26, pp. 1066–1072.

Beard, R. M., 1977. Expendable doppler penetrometer: a performance evaluation,

Technical Report R855, Civil Engineering Laboratory, Port Hueneme,

California.

Beard, R. M., 1981. A penetrometer for deep ocean seafloor exploration. Proceedings of

OCEANS 1981: The ocean an international workplace, MTS/IEEE, Boston, MA,

United States, pp. 668–673.

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., & Thevanayagam, S., 2009, Design and

characterization of a compact array of MEMS accelerometers for geotechnical

instrumentation, Smart Structures and Systems, 5(6), pp. 663–679.

Bhattacharya, S., Murali, K., Lombardi, D., Crewe, A. & Alexander, N., 2012.

Economic MEMS based 3-axis water proof accelerometer for dynamic geo-

engineering applications. Soil Dynamics and Earthquake Engineering, 36, pp.

111–118.

Bienen, B. Ragni, F., Cassidy. M. J. & Stanier, S. A., 2015. Effects of consolidation

under a penetrating footing in carbonate silty clay. Journal of Geotechnical and

Geoenvironmental Engineering, 04015040.

Bienen, B., Cassidy. M. J., Randolph, M. F. & Tel. K. L., 2015. Characterisation of

undrained shear strength using statistical methods. 2nd International Symposium

on Frontiers in Offshore Geotechnics, Perth, Australia, pp. 661–666.

- 242 -

Biscontin, G. & Pestana, J. M., 2001. Influence of peripheral velocity on vane shear

strength of an artificial clay. Geotechnical Testing Journal, ASTM, 24(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

Eighth International Conference on Soil Mechanics and Foundation

Engineering, Moscow, pp. 111–159.

Blake, A. P., O‟Loughlin, C.D., Morton, J. P., O‟Beirne, C., Gaudin, C. & White, D. J.,

2015. In-situ measurement of the dynamic penetration of free-fall projectiles in

soft soils using a low cost inertial measurement unit. Geotechnical Testing

Journal, ASTM. DOI: 10.1520/GTJ20140135.

Blake, A. P. & 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, pp.

621–628.

Blake, A. P. & O‟Loughlin, C. D., 2015. Installation of dynamically embedded plate

anchors as assessed through field tests. Canadian Geotechnical Journal. 0, 0,

10.1139/cgj-2014-0327

Boukpeti, N., White, D. J., Randolph, M. F. & Low, H. E., 2012. Strength of fine-

grained soils at the solid-fluid transition. Géotechnique, 62(3), pp. 213–226.

Bowman, L., March, R., Orenberg, P., True, D. & Herrmann, H., 1995. Evaluation of

dropped versus static cone penetrometers at a calcareous cohesive site.

Proceedings of OCEANS 1995: Challenges of Our Changing Global

Environment, MTS/IEEE, San Diego, CA, pp. 1846–1858.

- 243 -

Boylan, N., Long, M., Ward, D., Barwise, A. & Georgious, B., 2007. Full flow

penetrometer testing in Bothkennar clay. Proceedings of the 6th International

Conference, Society for Underwater Technology, Offshore Site Investigation and

Geotechnics (SUT-OSIG), London, pp. 177–186.

Brandão, F. E. N., Henriques, C. C. D., Araújo, J. B., Ferreira, O. C. G. & Amaral, C.

D. S., 2006. Albacora Leste field development - FPSO P-50 mooring system

concept and installation. Proceedings of the Offshore Technology Conference,

Houston, Texas, USA, OTC 18243.

Briaud, J. L., Garland, E. & Felio, G. Y., 1984. Rate of loading parameters for vertically

loaded piles in clay. In Proceedings of the 16th Annual Offshore Technology

Conference, Houston, Texas, USA, paper OTC 4694, pp. 407–412.

Britto, A. M. & Kusakabe, O., 1982. Stability of unsupported axisymmetric excavations

in soft clay. Géotechnique 32(3), pp. 261–270.

Britto, A. M. & Kusakabe, O., 1983. Stability of axisymmetric excavations in clays.

Journal of Geotechnical and Geoenvironmental Engineering, ASCE, 109(5), pp.

666–681.

Brown, M. J. & Hyde, A. F. L., 2008b. Rate effects from pile shaft resistance

measurements. Canadian Geotechnical Journal, 45(3), pp. 425–431.

Brown, P. P. & Lawler, D. F., 2003. Sphere drag and settling velocity revisited. Journal

of Environmental Engineering, 129, pp. 222–231.

Carter, J. P., Nazem, M. & Airey, D. W., 2013. Analysis of Dynamic Loading and

Penetration of Soils - Application to site investigation and ground improvement.

International Conference on Geotechnical Engineering, ICGE 12, Hammamet,

Tunisia.

- 244 -

Casagrande, A. & Wilson, S., 1951. Effect of rate of loading on the strength of clays

and shales at constant water content. Géotechnique, 2(3), pp. 251–563.

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. & Zielinski, A., 1978. Use of free fall penetrometer in sea

floor engineering. Proceedings of OCEANS 1978: The Ocean Challenge,

MTS/IEEE, Washington, D.C., pp. 686–691.

Chatterjee, S., Randolph, M. F. & White, D. J., 2012. The effects of penetration rate and

strain softening on the vertical penetration resistance of seabed pipelines.

Géotechnique 62(7), pp. 573–582.

Chow, S. H. & Airey, D. H., 2010. Laboratory free falling penetrometer test into clay.

Proceeding of the 2nd International Symposium on Frontiers in Offshore

Geotechnics, Perth, Australia, pp. 265–270.

Chow, S. H., 2013. Free-falling penetrometer tests in clay. PhD thesis, The University

of Sydney, Sydney, Australia.

Chow, S. H., O‟Loughlin, C. D. & Randolph, M. F. (2014). Soil strength estimation and

pore pressure dissipation for free-fall piezocones in soft clay. Géotechnique,

64(10), PP. 817–827.

Chung, S. F. & Randolph, M. F., 2004. Penetration resistance in soft clay for different

shaped penetrometers. In A. V. da Fonseca & P. W. Mayne, eds. Proc., ISC-2 on

Geotechnical and Geophysical Site Characterization. Portugal: Millpress,

Rotterdam, pp. 671–677.

- 245 -

Chung, S. F., Randolph, M. F. & Schneider, J. A., 2006. Effect of penetration rate on

penetrometer resistance in clay. Journal of Geotechnical and Geoenvironmental

Engineering, ASCE, 132(9), pp. 1188–1196.

Cilingir, U. C. & Madabhushi, G., 2011. A model study on the effects of input motion

on the seismic behavior of tunnels. Soil Dynamics and Earthquake Engineering,

31, pp. 452–462.

Cocjin, M. J., Gourvenec, S. M., White, D. J. & Randolph, M. F., 2014. Tolerably

mobile subsea foundations – observations of performance. Géotechnique,

64(10), pp. 817–827.

Colp, J., Caudle, W. & Schuster, C., 1975. Penetrometer system for measuring in situ

properties of marine sediment. Proceedings of OCEANS 1975, MTS/IEEE, San

Diego, CA, pp. 405–411.

Colreavy, C., O‟Loughlin, C. D. & 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, pp. 597–602.

Coussot, P., Laigle. D., Arratano, M., Deganutti, A. & Marchi, L., 1998. Direct

determination of rheological characteristics of debris flow. J. Hydraul. Eng. 124,

pp. 865–868.

CYR, R., Ardus, D. A. & Champ, M. A., 1990. Sea bed sampling with an expendable

acoustic penetrometer system. In D. A. Ardus & M. A. Champ (eds), Ocean

Resources, 2, Kluwer Academic: Netherlands, pp. 45–56.

Davies, C. N., 1945. Definitive equations for the fluid resistance of spheres.

Proceedings of the Physical Society, London, 57, pp. 259–270.

- 246 -

Dayal, U. & Allen, J. H., 1973. Instrumented cone penetrometer. Canadian

Geotechnical Journal, 10, pp. 397–409.

Dayal, U., 1974. Instrumented impact cone penetrometer. PhD thesis, Memorial

University of Newfoundland.

Dayal, U., Allen, J. & Jones, J., 1975. Use of an impact penetrometer for the evaluation

of the in-situ strength of marine sediments. Marine Georesources &

Geotechnology, 1(2), pp. 73–89.

Deep Sea Anchors., 2011. DPA™ fabricated at Western Ship Yard in Klaipeda,

Lithuania [image online] available: http://www.deepseaanchors.com/News.html

[accessed 4 November 2011].

DeJong, J. T., Yafrate, N. J., DeGroot, D. J. & Jakubowski, J., 2004. Evaluation of the

undrained shear strength profile in soft layered clay using full-flow probes.

Proceedings of the 2nd Inernational Confewrence on Site Characterisation,

Porto. Rotterdam, pp. 679–686.

DeJong, J., Yafrate, N., DeGroot, D., Low, H. E. & Randolph, M. F., 2010.

Recommended practice for full-flow penetrometer testing and analysis,

Geotechnical Testing Journal, ASTM, 33(2), pp. 137–149.

Deglo de Besses, B. D., Magnin, A. & Jay, P., 2003.Viscoplastic flow around a cylinder

in an infinite medium. Journal of Non-Newtonian Fluid Mechanics, 115, pp. 27–

49.

Denness, B., Berry, A., Darwell, J. & Nakamura, T., 1981. Dynamic seabed penetration.

Proceedings of OCEANS 1981: The ocean; an international workplace,

MTS/IEEE, Boston, MA, United States, pp. 662–667.

- 247 -

Dennis, S. C. R. & Walker, J. D. A., 1971. Calculation of the steady flow past a sphere

at low and moderate Reynolds numbers. Journal of Fluid Mechanics., 48, pp.

771–789.

Díaz-Rodríguez, J. A., Lozano-Santa Cruz, R., Davila-Alcocer, V. M., Vallejo, E. &

Girón, P., 1998. Physical, chemical, and mineralogical properties of Mexico

City sediments: A geotechnical perspective. Canadian Geotechnical Journal,

35(4), pp. 600–610.

Diaz-Rodriguez, J. A. & Martinez-Vasquez, J. J., 2005. Strain rate behaviour of Mexico

City Soils. Proceedings of the 16th International Conference on Soil Mechanics

and Geotechnical Engineering, Osaka, Japan, 2, pp. 333–336.

Douglas, B. L. & Wapner, M. P., 1996. Test results of the doppler penetrometer

seafloor sediment profiling system. Proceedings of OCEANS 1996: Prospects

for the 21st Century, MTS/IEEE, Fort Lauderdale, Florida, United States, 2, pp.

722–727.

Einav, I., Klar, A., O'Loughlin, C. D. & Randolph, M. F., 2004. Numerical modelling of

Deep Penetrating Anchors. Proceedings of the 9th Australian and New Zealand

Conference on Geomechanics, Auckland, New Zealand, Vol 2, pp. 818–824.

Einav, I. & Randolph, M. F., 2005. Combining upper bound and strain path methods for

evaluating penetration resistance. International Journal for Numerical Methods

in Engineering, 63(14), pp. 1991–2016.

Einav, I. & Randolph, M. F., 2006. Effect of strain rate on mobilised strength and

thickness of curved shear bands. Géotechnique, 56(7), pp. 501–504.

- 248 -

Elsworth, D. & Lee, D. S., 2005. Indentation of a free-falling lance penetrometer into a

poroelastic seabed. International Journal for Numerical and Analytical Methods

in Geomechanics, 29(2), pp. 141–162.

Fage, A. 1936 Experiments on a sphere at critical Reynolds numbers. Brit. Aero. Res.

Counc. Rep.& Mem. pp.1766.

Feng, X., Randolph, M.F., Gourvenec, S. & Wallerand R., 2013. Design approach for

rectangular mudmats under fully three-dimensional loading. Géotechnique,

64(1), pp. 51–63.

Fernandes, A. C., Araujo, J. B., Lima de Almeida, J. C., Machado, R. D. & Matos, V.,

2006. Torpedo anchor installation hydrodynamics. Journal of Offshore

Mechanics and Artic Engineering, 128(11), pp. 286–293.

Finnie, I. M. S. & Randolph, M. F., 1994. Punch-through and liquefaction induced

failure of shallow foundations on calcareous sediments. Proceedings of the

International Conference on behaviour of offshore structures, Boston, pp. 217–

230.

Finnemore, E. & Franzini, J., 2001. Fluid Mechanics With Engineering Applications –

McGraw-Hill Education.

Fossen, T. I., 2011, Handbook of marine craft hydrodynamics and motion control, John

Wiley & Sons, Ltd., Chichester, United Kingdom.

Freeman, T.J. & 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.

- 249 -

Freeman, T. J., Murray, C. N., Francis, T. J. G., McPhail, S. D. & Schultheiss, P. J.,

1984. Modelling radioactive waste disposal by penetrator experiments in the

abyssal Atlantic Ocean. Nature, 310, pp. 130–133.

Freeman, T. J., Murray, C. N. & Schuttenhelm, R. T. E., 1988. The Tyro 86 penetrator

experiments at Great Meteor East. In Oceanology’88: Proceedings of an

international conference. London: Graham & Trotman, pp. 217–226.

Freeman, T. J. & 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.

Furlong, A., Osler, J., Christian, H., Cunningham, D & 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 Undersea

Defence Technology Pacific Conference 2006, San Diego, US, pp. 1–13.

Gaudin, C., O‟Loughlin, C. D., Hossain, M. S & 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, Vol. 6, OMAE2013-10115.

Gaudin, C., White, D. J., Boylan, N., Breen, J., Brown, T., De Catania, S. & Hortin, P.,

2009. A wireless high-speed data acquisition system for geotechnical centrifuge

model testing. Measurement Science and Technology, 20(9): 095709,

http://dx.doi.org/10.1088/0957-0233/20/9/095709.

Georgiadis, M., 1991. Landslide drag forces on pipelines, Soils and Foundations.

Japanese Society of Soil mechanics and Foundation Engineering. 31(1), pp.

156–161.

- 250 -

Gibson, G. C. & Coyle, H. M., 1968. Soil damping constants related to common soil

properties in sands and clays, Technical Report 125-1,Texas Transportation

Institute.

Gibson, C. E., 1998. Lough Erne. In: Moriarty, C. (Ed.), Studies of Irish Lakes and

Rivers. Marine Institute, Dublin, pp. 64–72.

Gilbert, R. B., Morvant, M. & Audibert, J., 2008. Torpedo piles joint industry project –

model torpedo pile tests in kaolinite test beds, Report No. 02/08B187, Offshore

Technology Research Center, Texas A & M University.

Goin, K. L. & Lawrence, W. R., 1968. Subsonic drag of spheres at Reynolds numbers

from 200 to 10,000. AIAA J., 6, pp. 961–962.

Govardhan, R. & Williamson, C. H. K. 2005 Votex-induced vibration of a sphere.

Journal of Fluid Mechanics. 531, pp. 11–47.

Graham, J., Crooks, J. H .A. & Bell, A. L., 1983. Time effects on the stress-strain

behaviour of natural soft clays. Géotechnique, 33(3), pp. 327–340.

Gunn, R. & Kinzer, G. D., 1949. The terminal velocity of fall for water droplets in

stagnant air. J. Meteorol., 6, pp. 243–250.

Hasanloo, D., Pang, H. & Yu, G., 2012. On the estimation of the falling velocity and

drag coefficient of torpedo anchor during acceleration. Ocean Engineering, 42,

pp. 135–146.

Hartman, M., Trnka, O. , & Svoboda, K.,1994. Free settling of nonspherical particles.

Industrial & Engineering Chemistry Research,. 33, pp. 1979–1983.

Hossain, M. S, Kim, Y. & Wang, D., 2013. Physical and numerical modelling of

installation and pull-out of dynamically penetrating anchors in clay and silt.

- 251 -

Proceeding of the 32nd International Conference on Offshore Mechanics and

Arctic Engineering, Nantes, France, Vol. 6, OMAE2013-10322.

Hossain M. S., Hu Y., Randolph M. F. & White D. J., 2005. Limiting cavity depth for

spudcan foundations penetrating clay. Géotechnique, 55(9), pp. 679–690.

Hossain, M. S., Kim, Y. & Gaudin, C., 2014. Experimental investigation of installation

and pullout of dynamically penetrating anchors in clay and silt. Journal of

Geotechnical and Geoenvironmental Engineering, 140(7), pp. 04014026.

Houlsby, G.T. & Wroth, C.P., 1983. Calculation of stresses on shallow penetrometers

and footings. Proceedings of IUTAM/IUGG Symposium on Sebed Mechanics.

Newcastle upon Tyne, pp. 107–112.

House, A. R., Oliveira, J. R. M. S. & Randolph, M. F., 2001. Evaluating the coefficient

of consolidation using penetration tests. International Journal of Physical

Modelling in Geotechnics, 1(3), pp. 17–25.

Houwink, R., 1952. Elasticity, plasticity and structure of matter. Cambridge University

Press.

Hu, P., Stainer, S. A., Cassidy, M. J. & Wang, D., 2014. Predicting peak resistance of

spudcan penetrating sand overlying clay. Journal of Geotechnical and

Geoenvironmental Engineering ASCE, 140(2).

Hurst, R. B. & Murdoch, S., 1991. Measurement of sediment shear strength for mine

impact burial predictions. Report of the 18th Meeting of the Mines and Mine

Countermeasures Technology Panel GTP-13.

Hyde, A. E. L., Anderson, W. F. & Robinson, S. A., 2000. Rate effects in clay soils and

their relevance to statnamic pile testing. In Statnamic Loading Test: Proceedings

of the 2nd International Statnamic Seminar, Tokyo, Japan, pp. 303–309.

- 252 -

Jeanjean, P., Spikula, D. & Young, A., (2012). Technical vetting of free-fall cone

penetrometer. Proceedings of 7th international offshore site investigation and

geotechnics conference: Integrated geotechnologies – present and future, pp.

15–18. London UK: Society for Underwater Technology.

Jeong, S. W., Leroueil, S. & Locat, J., 2009. Applicability of power law for describing

the rheology of soils of different origins and characteristics. Canadian

Geotechnical Journal, 46, pp. 1011–1023.

Jones, D. A. & Clarke, D. B., 2008. Simulation of flow past a sphere using the

FLUENT code. Defence Science and Technology Organisation for the

Australian Government Department of Defence. DSTO-TR-2232.

Jonkman, J. M., 2007, Dynamics modelling and loads analysis of an offshore floating

wind turbine. National Renewable Energy Laboratory Technical Report

NREL/TP-500-41958, National Renewable Energy Laboratory, Golden,

Colorado, United States.

Kee, R. & Ims, B. W., 1984. Geotechnical hazards associated with leg penetration of

jack-up rigs. Seabed Mechanics, pp. 169–174. London: Graham & Trotman.

Kelleher, P. & Randolph, M. F., 2005. Seabed geotechnical characterisation with a ball

penetrometer deployed from the portable remotely operated drill. Proceedings of

International Symposium on Frontiers in Offshore Geotechnics (ISFOG), pp.

365–371.

King, K., Yoon, S. W., Perkins, N. C., & Najafi, K., 2008, Wireless MEMS inertial

sensor system for golf swing dynamics, Sensors and Actuators, 141(2), pp. 619–

630.

- 253 -

Kunitaki, D. M. K. N., De Lima, B. S. L. P., Evsukoff, A. G. & Jacob, B. P., 2008.

Probabilistic and fuzzy arithmetic approaches for the treatment of uncertainties

in the installation of torpedo piles, Mathematical Problems in Engineering,

Article ID 512343.

Kusakabe, O., Suzuki, H. & Nakase, A., 1986. An upper bound calculation on bearing

capacity of a circular footing on a non-homogeneous clay. Soils and

foundations, 26(3), pp.143–148.

Ladd, C. C., Foot, R., Ishihara, K., Schlosser, F. & Poulos, H. G., 1977. Stress–

deformation and strength characteristics. Proceedings of the 9th International

Conference on Soil Mechanics. Found. Engng, Tokyo 2, pp. 421–494.

Lamb, H., 1932. Hydrodynamics. Cambridge University Press.

Lavieri, R. S., Tannuri, E. A., Fujarra, A. L. C, & Corr a, D. C., 2011. Aided Inertial

Measurement System Applied to Torpedo Pile Trajectory Reconstruction. 30th

International Conference on Ocean, Offshore and Arctic Engineering,

Rotterdam, The Netherlands, 1, pp. 59–66.

Lafferty, B., Quinn, R. & Breen, C., 2006. A side-scan sonar and high-resolution chirp

sub-bottom profile study of the natural and anthropogenic sedimentary record of

Lower Lough Erne, northwestern Ireland. Journal of Archaeological Science, 33

pp. 756–766.

Lefebvre, G. & LeBoeuf, D., 1987. Rate effects and cyclic loading of sensitive clays.

Journal of Geotechnical Engineering, 113(5), pp. 476–489.

Lehane, B. M., O‟Loughlin, C. D., Gaudin, C. & Randolph, M. F., 2009. Rate effects on

penetrometer resistance in kaolin. Géotechnique, 59(1), pp. 41–52.

- 254 -

Levacher, D., 1985. Penetrometre a chute libre: impact et penetration dans des argiles

reconstituees. Canadian Geotechnical Journal, 22(1), pp. 129–135.

Levy, F. M., & Richards, D. J., 2012, Rapid soil displacements from MEMS

accelerometers, Proceedings of the 30th IMAC Conference on Structural

Dynamics, Jacksonville, Florida, United States, 30 January–2 February 2012. 1,

pp. 197–208.

Liebster, H. 1927. Uber den widerstand von kugeln. Annalen der Physik (Leipzig), 82,

541 562.

Lieng, J. T., Hove, F. & Tjelta, T. I., 1999. Deep Penetrating Anchor : subseabed

deepwater anchor concept for floaters and other installations. Proceedings of the

9th International Offshore & Polar Engineering Conference, Brest, France, pp.

613–619.

Lieng, J. T., Tjelta, T . I. & 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, OTC 20758.

Liu, X. D., Sun, J., Yi, J. T., Lee, F. H., Gu, H. & Chow, Y. K., 2014. Physical and

Numerical Modeling of the Performance of Dynamically Installed Anchors in

Clay. 33rd International Conference on Ocean, Offshore and Arctic Engineering

Volume 1A: Offshore Technology San Francisco, California, USA.

Long, M. & Gudjonsson, G.T., 2004. T-bar testing in Irish soils. Proceedings of the 2nd

International Conference on Site Characterisation, Porto, pp. 719–726.

Low, H. E., Randolph M. F. & Kelleher, P., 2007. Comparison of pore pressure

generation and dissipation rates from cone and ball penetrometers. Proceedings

of the 6th International Conference of the Society for Underwater Technology,

- 255 -

Offshore Site Investigation and Geotechnics (SUT-OSIG), London, pp. 547–

556.

Low, H. E., Randolph, M. F., DeJong, J. T. & Yafrate, N. J., 2008. Variable rate full-

flow penetration tests in intact and remoulded soil. Proceedings of the 3rd

International Conference on Site Characterization, Taipei, pp. 1087–1092.

Low, H.E., Randolph, M. F., Lunne, T., Andersen, K. H. & Sjursen, M.A. 2011. Effect

of soil characteristics on relative values of piezocone, T-bar and ball penetration

resistances. Géotechnique, 61(8), pp. 651–664.

Lu, Q., Hu, Y. & Randolph, M. F., 2000. FE analysis for T-bar and spherical

penetrometers in cohesive soil. Proceedings of the 10th International Offshore

and Polar Engng Conference. Seattle, 2, pp. 617–623.

Lu, Q., Hu, Y. & Randolph, M. F., 2001. Deep penetration in soft clay with strength

increasing with depth. Proceedings of the 11th International Offshore and Polar

Engng Conference. Stavanger 2, pp. 453–458.

Lunne, T., Christoffersen, H. P. and Tjelta, T. I., 1985. Engineering use of piezocone

data in North Sea clays. Proceedings of the 11th International Conference on

Soil Mechanics and Foundation Engineering, San Francisco, 2, pp. 907–912.

Lunne, T., Robertson, P. K. and Powell, J. J. M., 1997. Cone Penetration Testing in

Geotechnical Engineering, Blackie Academic and Professional, London.

Lunne, T. 2001. In situ testing in offshore geotechnical investigations. Proceedings of

the International Conference on In Situ Measurement of Soil Properties and

Case Histories, Bali, pp. 61–81.

- 256 -

Lunne, T., Berre, T., Andersen, K. H. & Strandvik, S., 2006. Effects of sample

disturbance and consolidation procedures on measured shear strength of soft

marine Norwegian clays. Canadian Geotechnical Journal, 43(7), pp. 726–750.

Lunne, T. & Andersen, K. H., 2007. Soft clay shear strength parameters for deepwater

geotechnical design. Proceedings of the 6th International Offshore Site

Investigation and Geotechnics Conference: Confronting New Challenges and

Sharing Knowledge, London, UK, pp. 151–176.

Lunnon, R. G., 1928. Fluid resistance to moving spheres. Proceedings of the Royal

Society, London, 118, pp. 680–694.

Martin, C. M. & Randolph, M. F., 2001. Applications of the lower and upper bound

theorems of plasticity to collapse of circular foundations. Proceedings of the

International Conference on Computer Methods and Advances in

Geomechanics. Tuscon, pp. 1417–1428.

Martin, C. M. & Randolph, M. F., 2006. Upper-bound analysis of lateral pile capacity in

cohesive soil. Geotechnical Testing Journal, ASTM, 56(2), pp. 141–145.

Maxworthy, T., 1965. Accurate measurements of sphere drag at low reynolds numbers.

Journal of Fluid Mechanics., 23, pp. 369–372.

Maxworthy, T., 1969. Experiments on the flow around a sphere at high Reynolds

numbers. J. Appl. Mech., Trans. A.S.M.E. 36, pp. 598–607.

Medeiros., C. J., 2002. Low cost anchor system for flexible risers in deep waters.

Proceedings of Offshore Technology Conference. Houston, Texas, pp. 1–5.

Medeiros., C. J., Hassui, L. H. Machado, R. D., (1997). Pile for Anchoring Floating

Structures and Process for Installing the Same. United States Patent Number

6,106,199.0

- 257 -

Merifield R., White D. J. & Randolph M.F., 2009. The effect of surface heave on the

response of partially-embedded pipelines on clay. Journal of Geotechnical and

Geoenvironmental Engineering, ASCE, 135(6) pp. 819–829.

Mesri, G., Rokhsar, A. & Bohor, B. F., 1975. Composition and compressibility of

typical samples of Mexico City clay. Géotechnique, 24(3), pp. 527–554.

Meyerhof, G. G., 1972. Stability of slurry trench cuts in saturated clay. Proceedings of

the Speciality Conference on Performance of Earth and Earth Supported

Structures, 1, pp.1451–1466.

Migliore, H. J. & Lee, H. J., 1971. Seafloor penetration tests: presentation and analysis

of results. Naval Civil Engineering Laboratory Technical Note N-1178. Port

Hueneme, CA, pp. 60.

Millikan, C. B. & Klein, A. L., 1933. The effect of turbulence. An investigation of

maximum lift coefficient and turbulence in wind tunnels and in flight. Aircraft

Eng.

Mitchell, J. K. & Soga, K., 2005. Fundamentals of soil behavior. New York: Wiley.

Moller, W.,1938. Experimentelle untersuchungen zur hydrodynamik der kugel. Phys.

Z., 39, pp. 57–80.

Morrison, F. A., 2013. An Introduction to Fluid Mechanics. New York: Cambridge

University Press.

Morison, J. R., O'Brien, M. P., Johnson, J. W. & Schaaf, S. A., 1950. The force exerted

by surface waves on piles. Petroleum Transactions, 189, pp. 149–154.

Morton, J. P., & O‟Loughlin, C. D., 2012, Dynamic penetration of a sphere in clay,

Proceedings of the 7th International Conference in Offshore Site Investigation

- 258 -

and Geotechnics, London, United Kingdom, 12–14 September 2012. pp. 223–

230.

Morton, J. P., O‟Loughlin, C. D. & White, D. J. (2014). Strength assessment during

shallow penetration of a sphere in clay. Géotechnique Letters 4, October–

December, pp. 262–266.

Morton, J. P., O‟Loughlin, C. D. & White, D. J., 2015, An experimental assessment of

the dynamic resistance of a free-falling sphere in soft clay, in preparation.

Mulhearn, P. J., Poeckert, R. H. & 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.

Mulhearn, P. J., Morgan, S., Poeckert, R. H. & Holtkamp, R. H., 1999. Penetrometer

comparison trials, Sydney Harbour, March 1998, TTCP Report, DSTO,

Australia.

Mulhearn, P. J., 2003. Influences of penetrometer tip geometry on bearing strength

estimates. International Journal of Offshore and Polar Engineering, 13(1), pp.

73–78.

Mulukutla, G. K., 2009. Determination of geotechnical properties of seafloor sediment.

PhD thesis, University of New Hampshire.

Murff, J. D. & Coyle, H. M. (1973). Low velocity penetration of kaolin clay. Journal of

Soil Mechanics & Foundations Div, 99(sm5).

Murff, J. D., Wagner, D. A. & Randolph, M. F., 2009. Pipe penetration in cohesive soil.

Géotechnique, 39(2), pp. 213–229.

- 259 -

Nakase, A. & Kamei, T., 1986. Influence of strain rate on undrained shear

characteristics of Ko-consolidated cohesive soils. Soils and Foundations, 26(1),

pp. 85–95.

Nazem, M. & Carter, J. P., 2010. Parametric study of a free-falling penetrometer in

clay-like soils. Proceedings of the International Symposium on Frontiers in

Geotechnics. ISFOG 2010, Perth, pp. 293–298.

Nazem, M., Carter, J. P., Airey, D. W. & Chow, S. H., 2012. Dynamic analysis of a

smooth penetrometer free falling into uniform clay. Géotechnique, 62(10), pp.

893–905.

Nebot, E. & Durrant-Whyte, H., 1999, Initial calibration and alignment of low cost

inertial navigation units for land vehicle applications, Journal of Robotics

Systems, 16(2), pp. 81–92.

Noureldin, A., Karamat, T. B., & Georgy, J., 2012, Fundamentals of inertial navigation,

satellite-based positioning and their integration, Springer-Verlag, Berlin,

Germany.

O‟Loughlin, C. D., Gaudin, C., Morton, J. P. & White D. J., 2014, 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., Blake, A. P., Wang, D., Gaudin, C. & 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, OMAE2013-11571.

- 260 -

O‟Loughlin, C. D., Richardson, M. D., Randolph, M. F. & Gaudin, C., 2013b.

Penetration of dynamically installed anchors in clay. Géotechnique, 63(11) pp.

909–919.

O‟Loughlin, C. D., Richardson, M. D. & Randolph, M. F., 2009. Centrifuge tests on

dynamically installed anchors. Proceedings of the ASME 28th International

Conference on Ocean, Offshore and Arctic Engineering, Honolulu, Hawaii, pp.

1–9.

O'Loughlin, C. D., Randolph, M. F. & Einav, I., 2004a. Physical modelling of deep

penetrating anchors. Proceedings of the 9th Australian and New Zealand

Conference on Geomechanics, Auckland, New Zealand, Vol. 2, pp. 710–716.

O'Loughlin, C. D., Randolph, M. F. & 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, United States Naval Academy, Annapolis,

Maryland.

Øye, I., 2000. Simulation of trajectories for a deep penetrating anchor. CFD Norway

Report 250:2000, CFD Norway AS, Trondheim, Norway.

Pantaleone, J. & Messer, J., 2011. The added mass of a spherical projectile. American

Journal of Physics, 79(12), pp. 1202–1210.

Pazwash, H. & Robertson, J.M., 1969. Fluid-dynamic consideration of bottom

materials. Conference on Ocean Engeneering, ASCE - 1970.

Pazwash, H. & Robertson, J.M., 1975. Forces on bodies in Bingham fluids, Journal of

Hydraulic Research. 13, pp. 35–55.

- 261 -

Pettyjohn, E. S. & Christiansen, E. B., 1948. Effect of particle shape on free-settling

rates of isometric particles. Chem. Eng. Prog. 44, pp. 157–172.

Puech, A., Orozco-Caldero´n, M. & Foray, P., 2010. Mini T-bar testing at shallow

penetration. Proceedings of Frontiers in Offshore Geotechnics, Perth, pp. 305–

310.

Peuchen, J. & Mayne, P., 2007. Rate effects in vane shear testing. Proceedings of the

6th International Conference on Offshore Site Investigation and Geotechnics:

Confronting New Challenges and Sharing Knowledge. London, UK, pp. 187–

194.

Peuchen, J., Adrichem, J. & Hefer, P., 2005. Practice notes on push-in penetrometers

for offshore geotechnical investigation. pp. 19–21.

Poorooshasb, F. & James, R. G., 1989. Centrifuge modelling of heat-generating waste

disposal. Canadian Geotechnical Journal, 26, pp. 640–652.

Powell, J. & Quarterman, R., 1988. The interpretation of cone penetration tests in clays,

with particular reference to rate effects. In Penetration Testing 1988. Balkema,

Rotterdam, pp. 903–909.

Pruppacher, H. R. & Steinberger, E. H., 1968. Experimental determination of the drag

on a sphere at low Reynolds numbers. J. Appl. Phys. 39(9), pp. 4129–4132.

Quirós, G.W. & Young, A.G., 1988. Comparison of field vane, CPT and laboratory

strength data at Santa Barbara Channel site. Vane Shear Strength Testing of

Soils: Field and Laboratory Studies, ASTM STP 1014: pp. 306–317.

Raie, M. & Tassoulas, J. L., 2006. Numerical modeling of torpedo anchors. Final

project report – Phase I, Minerals Management Service, Project No. 557.

- 262 -

Raie, M. S. & Tassoulas, J. L., 2009. Installation of torpedo anchors: numerical

modelling. Journal of Geotechnical and Geoenvironmental Engineering,

135(12), pp. 1805–1813.

Randolph, M. F. & White, D. J., 2012. Interaction forces between pipelines and

submarine slides – a geotechnical viewpoint. Ocean Engineering, Vol. 48, pp.

32–37.

Randolph, M. F., Low, H. E. & 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.

Randolph, M. F., 2004. Characterisation of soft sediments for offshore applications.

Proceedings of 2nd International Conference on Geotechnical and Geophysical

Site Characterization - ISC‟2, Porto. Rotterdam, pp. 209–232.

Randolph, M. F. & 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, pp. 147–152.

Randolph, M. F., Martin, C. M. & Hu, Y., 2000. Limiting resistance of a spherical

penetrometer in cohesive material. Géotechnique, 50(5), pp. 573–582.

Randolph, M. F. & White, D. J., 2008. Pipeline embedment in deep water: processes

and quantitative assessment. Proceedings of Offshore Technology Conference,

Houston, 1, pp. 1–16.

Randolph, M. F., Hefer, P. A., Geise, J. M. & Watson, P. G., 1998. Improved seabed

strength profiling using T-bar penetrometer. Proceedings International

- 263 -

Conference on Offshore Site Investigation and Foundation Behavior - New

Frontiers, Society for Underwater Technology. London, UK, pp. 251–262.

Randolph, M. F. & Houlsby, G.T., 1984. The limiting pressure on a circular pile loaded

laterally in cohesive soil. Géotechnique, 34(4), pp. 613–623.

Ratcliffe, C., Heider, D. Crane, R., Krauthauser, K., Yoon, M.K. & Gillespie, J.W.,

2008. Investigation into the use of low cost MEMS accelerometers for vibration

based damage detection. Composite Structures, 82, pp.61–70.

Richardson, M. D., O‟Loughlin, C. D., Randolph, M. F. & Cunningham T. J., 2006.

Drum centrifuge modeling of dynamically penetrating anchors. Proceedings of

International Conference on Physical Modelling in Geotechnics, London. 1, pp.

673–678.

Richardson, M. D., 2008. Dynamically installed anchors for floating offshore

structures. PhD thesis, The University of Western Australia.

Richardson, M. D., O‟Loughlin, C. D., Randolph, M. F. & Gaudin, C., 2009. Setup

following installation of dynamic anchors in normally consolidated clay.

Journal of Geotechnical and Geoenvironmental Engineering ASCE, 135(4), pp.

487–496.

Rimon, Y. & Cheng, S. I., 1969. Numerical solution of a uniform flow over a sphere at

intermediate Reynolds numbers. Phys. Fluids, 12(5), pp. 949–959.

Pazwash, H., & Robertson, J. M., 1971. Fluid-Dynamic Consideration of Bottom

Materials. Journal of the Hydraulics Division, 97(9), pp. 1317-1329.

Roos, F. W. & Willmarth, W. W., 1971. Some experimental results on sphere and disk

drag. AIAA J., 9, 285–291.

- 264 -

Sabetamal, H., Nazem, M. & Carter, J. P., 2013. Numerical analysis of torpedo

anchors. ComGeo III, pp. 621–632.

Sahdi, F., Gaudin, C., White, D. J., Boylan, N. & Randolph, M. F., 2014. Centrifuge

modelling of active slide – pipeline loading in soft clay. Géotechnique, 64(1),

pp. 16–27.

Salençon, J. & Matar, M., 1982. Capacité portante des fonda- tions superficielles

circulaires. Journal de Mécanique théorique et appliquée, 1(2), pp. 237–267.

Schlichting, H., 1955. Boundary Layer Theory. New York: McGraw-Hill.

Schlichting, H., Gersten, K. & Gersten, K., 2000. Boundary-layer theory. Springer

Science & Business Media.

Schlue, B. F., Moerz, T. & Kreiter, S., 2010. Influence of shear rate on undrained vane

shear strength of organic harbor mud. Journal of Geotechnical and

Geoenvironmental Engineering, ASCE, 136(10), pp. 1437–1447.

Schmid, W. E., 1969. Penetration of objects into the ocean bottom, Report No. AD

695434, Naval Civil Engineering Laboratory, Port Hueneme, Califorfnia.

Schmiedel, J., 1928. Experimentelle untersuchungen uber die fallbewegung von kugeln

und scheiben in reibenden flussigkeiten. Phys. Z., 29, pp. 593–610.

Schofield, A. N. & Wroth, C. P., 1968. Critical state soil mechanics. London: McGraw-

Hill.

Scott, R. F., 1970. In-place ocean soil strength by accelerometer. Journal of the Soil

Mechanics and Foundations Division, 96(1), pp. 199–211.

Shakespear, G. A., 1914. Experiments on the resistance of the air to falling spheres.

Philos. Mag. 28, pp. 728–734.

- 265 -

Sharma, A., 2007, CMOS systems and circuits for sub-degree per hour MEMS

gyroscopes, PhD thesis, Georgia Institute of Technology, Atlanta, Georgia,

United States.

Sheahan, T. C., Ladd, C. C. & Germaine, J. T., 1996. Rate-dependent undrained shear

behavior of saturated clay. Journal of Geotechnical Engineering, 122(2), pp.

99–108.

Shelton, J.T., 2007. OMNI-Max anchor development and technology. In: Proceedings

of the Ocean Conference, Vancouver, Canada.

Skempton, A. W., 1951. The bearing capacity of clays. In Proc. Building Research

Congress. pp. 180–190.

SNAME., 1997. Recommended practice for site specific assessment of mobile jack-up

units, Rev. 1 Jersey City, NJ: Society of Naval Architects and Marine Engineers.

Stanier, S.A. & White, D.J. (2014). Shallow penetrometer penetration resistance.

Journal of Geotechnical and Geoenvironmental Engineering, ASCE, 141(3),

04014117.

Stark, N. & Wever, T. F., 2009. Unraveling subtle details of expendable bottom

penetrometer (XBP) deceleration profiles. Geo-Marine Letters, 29(1), pp. 39–

45.

Stark, N., Hanff, H. & Kopf, A., 2009a. Nimrod: A tool for rapid geotechnical

characterization of surface sediments. Sea Technology, 50(4), pp. 10–14.

Stark, N., Kopf, A., Hanff, H., Stegmann, S. & Wilkens, R., 2009b. Geotechnical

investigations of sandy seafloors using dynamic penetrometers. Proceedings of

OCEANS 2009: Marine Technology for Our Future: Global and Local

Challenges, MTS/IEEE Biloxi, Biloxi, MS., pp. 1–10.

- 266 -

Stark, N., Wilkens, R., Ernstsen, V. B., Stegmann, S. & Kopf, A., 2012. Geotechnical

properties of sandy seafloors and the consequences for dynamic penetrometer

interpretations : quartz sand versus carbonate sand. Geotechnical and

Geological Engineering, 30, pp. 1–14.

Stegmann, S., Morz, T. & 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, 86(3), pp.199–208.

Stegmann, S., 2007. Design of a free-fall penetrometer for geotechnical

characterisation of saturated sediments and its geological application. PhD

thesis, Bremen University.

Steiner, A., L'Heureux, J., Kopf, A., Vanneste, M., Longwa, O., Lange, M. & 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. ed.

Submarine Mass Movements and Their Consequences. Springer

Science+Business, pp. 99–109.

Steiner, A., Kopf, A. J., L‟Heureux, J. S., Kreieter, S., Stegmann, S., Haflidason, H., &

Moerz, T., 2014, In situ dynamic piezocone penetrometer tests in natural clayey

soils – a reappraisal of strain-rate corrections, Canadian Geotechnical Journal,

51(3), pp. 272–288.

Stephan, S., Kaul N. & 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, 33(3),

pp. 209–221.

- 267 -

Stewart, D. P. & Randolph, M. F., 1991. A new site investigation tool for the centrifuge.

Proceedings of the International Conference on Centrifuge Modelling, Boulder,

Colorado, USA, pp. 531–538.

Stokes, G., 1880. Mathematical and physical papers, Vol. 1, Cambridge University

Press, Cambridge, U.K.

Stoll, R. D. & Akal, T., 1999. XBP-Tool for rapid assessment of seabed sediment

properties. Sea Technology,40(2), pp. 47–51.

Stoll, R. D., 2004. Measuring sea bed properties using static and dynamic

penetrometers. In Proc. ASCE International Conference Civil Engineering

Oceans VI. pp. 386–395.

Stoll, R. D., Sun, Y. F. & Bitte, I., 2007. Seafloor properties from penetrometer tests.

Journal of Oceanic Engineering, IEEE, 32(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, United States.

Stringer, M. E., Heron, C. M. & Madabhushi, S. P. G., 2010. Experience using MEMS-

based accelerometers in dynamic testing. Proceedings of the 7th International

Conference on Physical Modelling in Geotechnics, Zurich, Switzerland, Vol. 1,

pp. 389–394.

Strum, H., Lieng, J. T. & 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, OTC 22396.

- 268 -

Strum, H. & 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, pp. 255–260.

Sumer, B.M. & Fredsoe, J.,1997. Hydrodynamics around cylindrical structures. World

Scientific. Singapore.

Suryanarayana, G. K. & Prabhu, A., 2000. Effect of natural ventilation on the boundary-

layer separation and near-wake vortex shedding characteristics of a sphere. Exp.

Fluids 29, pp. 582–591.

Tanaka, H. & Locat, J., 1999, A microstructural investigation of Osaka Bay Clay: the

impact of microfossils on its mechanical behaviour, Canadian Geotechnical

Journal, 36(3), pp. 493–508.

Tani, K. & Craig, W. H., 1995. Bearing capacity of circular foundations on soft clay of

strength increasing with depth. Soils and Foundations, 35(4), pp. 21–35.

Teh, C. I. & Houlsby, G. T. 1991. An analytical study of the cone penetration test in

clay. Géotechnique, 41(1), pp. 17–34.

Tho, K. K., Leung, C. F., Chow, Y. K. & Palmer, A. C., 2012. Deep cavity flow

mechanism of pipe penetration in clay. Canadian Geotechnical Journal, 49(1),

pp. 59–69.

Tho, K. K., Chen, Z., Leung, C. F. & Chow, Y. K., 2013. Pullout behaviour of plate

anchor in clay with linearly increasing strength. Canadian Geotechnical

Journal, 51(1), pp. 92–102.

Thompson, D., March, R. & Herrmann, H., 2002. Groundtruth results for dynamic

penetrometers in cohesive soils. Proceedings of OCEANS 2002: Marine

- 269 -

Frontiers: Reflection of the Past, Visions of the Future. MTS/IEEE, Biloxi, MS,

4, pp. 2117–2123.

True, D. G., 1976. Undrained vertical penetration into ocean bottom soils. PhD thesis,

University of California, Berkeley.

Vaid, Y. P. & Campanella, R. G., 1977. Time-dependent behaviour of undisturbed clay.

Journal of the Geotechnical Engineering Division, 103(7), pp. 693–709.

Watson, P. G., Newson, T. A. & Randolph, M. F., 1998. Strength profiling in soft

offshore soils. Proceedings of the 1st International Conference on Site

Characterization – ISC ‟98, Atlanta, pp. 1389–1394.

Westgate, Z. J., White, D. J. & Randolph, M. F., 2010. Pipeline laying and embedment

in softgrained soils: field observations and numerical simulations. Proceedings

of the Offshore Technology Conference. 2010. OTC 20407.

White, D. J., 2015. Personal communication.

White, D. J. & Randolph, M. F., 2007. Seabed characterisation and models for pipeline-

soil interaction. The International Journal of Offshore and Polar Engineering,

17(3), pp. 193–204.

White, D. J., Gaudin, C., Boylan, N. & Zhou, H., 2010. Interpretation of T-bar

penetrometer tests at shallow embedment and in very soft soils. Canadian

Geotechnical Journal, 47(2), pp. 218–229.

Wieselsberger, C., 1923. Der widerstand von kugeln, ellipsoiden und scheiben. Ergebn.

der Aerodynamischen Versuchsanstalt, 2(28), pp. 28–32.

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

- 270 -

Yafrate, N. J. & Dejong, J. T., 2006. Interpretation of sensitivity and remolded

undrained shear strength with full flow penetrometers. Proceedings of 16th

International Conference Offshore and Polar Engineering Conf. San Francisco,

USA. pp. 572–577.

Yafrate, N.J. & Dejong, J.T., 2007. Influence of penetration rate on measured resistance

with full flow penetrometers in soft clay. In Geotechnical Special Publication, n

173, Proceedings of Sessions of Geo-Denver 2007 Congress: Advances in

Measurement and Modeling of Soil Behavior. pp. 1–10.

Yan, Y., White, D.J. & Randolph, M.F., 2011. Penetration resistance and stiffness

factors for hemispherical and toroidal penetrometers in uniform clay.

International Journal of Geomechanics, 11(4), pp. 263–275.

Young, A. G., Bernard, B. B., Remmes, B. D. Babb, L. V. & 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.

Zakeri, A., Hoeg, K. & Nadim, F., 2008. Submarine debris flow impact on pipelines –

Part I: Experimental investigation. Coastal Engineering, 55(12): 1209–1218.

Zakeri, A., Hoeg, K. & Nadim, F., 2009. Submarine debris flow impact on pipelines –

Part II: Numerical analysis. Coastal Engineering, 56(1) pp. 1–10.

Zakeri, A., Chi, K. & Hawlader, B., 2011. Centrifuge modelling of glide block and out-

runner block impact on submarine pipelines. In Proceedings of the Offshore

Technology Conference, Houston, Texas, USA, paper OTC 21256.

Zhou, H. & Randolph, M.F., 2009a. Resistance of full-flow penetrometers in rate-

dependent and strain-softening clay. Géotechnique, 59(2), pp. 79–86.

- 271 -

Zhou, H. J. & Randolph, M. F., 2007. Computational techniques and shear band

development for cylindrical and spherical penetrometers in strain-softening clay.

International Journal of Geomechanics, 7(4), pp. 287–295.

Zhou, H., Hossain, M. S., Hu, Y. & Liu, H., 2013. Behaviour of a ball penetrometer in

uniform single and double layer clays. Géotechnique 63(8), pp. 682–694.

Zhu, H. & Randolph, M.F., 2011. Numerical analysis of a cylinder moving through

rate-dependent soils. Ocean Eng. 38 (7), pp. 943–953.

Zhu, J. G. & Yin, J. H., 2000. Strain-rate-dependent stress-strain behavior of

overconsolidated Hong Kong marine clay. Canadian Geotechnical Journal,

37(6), pp. 1272–1282.

Zimmerman, E. H. & 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., 2007. New gravity installed anchor requires less soil penetration.

Offshore, Vol. 67(1), pp. 70–71.

Zimmerman, E.H., Smith, M.W. & Shelton, J.T., 2009. Efficient Gravity Installed

Anchor for Deepwater Mooring. Proceedings of the Offshore Technology

Conference, Houston, Texas, OTC 20117.

- 272 -

APPENDIX A. BUOYANCY FORMULATION

A.1. Analytical closed-form solution

- 273 -

This section describes an analytical closed form solution to calculate the soil buoyancy

for an advancing sphere. The method was introduced in Chapter 4, however, the

solution was not presented in the publication.

In order to capture the increased buoyancy force due to soil heave, a simple multiplier

can be applied on Fbuoy = Vballγ' which can be derived by considering the work required

to lift soil from the position of the soil mass centroid, shown on Figure 4.4 a and b. For

an incrementally penetrating sphere, the volume of soil (V) that is to be lifted to the

surface can be expressed as the volume of the (curved) blue cylinder shown in Figure

9.1 and Figure 9.2 where,

2cos12

2sincossin222

0

2

0

2

0

maxmaxmax

RdRdRRHdzV 9.1

The position of the mass centroid depends on the sphere penetration depth. For a sphere

penetrating soil with an invert depth zi ≤ 0.5D (Figure 9.1), all of the soil displaced by

the advancing spherical cap is lifted to the soil surface from the changing centroidal

height position, h . However, at an invert depth 0.5D < zi ≤ Hc + D, (where Hc is the

height of the conical cavity) only a portion of the soil displaced by the advancing sphere

is lifted to the surface because some of the soil is allowed to flow partially above the

sphere which is shown in Figure 4.4b. In this case, part, and not all, of the displaced soil

will be lifted to the surface from the centroidal height, h of the upper spherical cap to the

soil surface. The estimation of h for both cases is shown in Figure 9.1 and Figure 9.2.

- 274 -

Figure 9.1. The formulation of the centroidal height for an invert depth < 0.5D

Figure 9.2. The formulation of the centroidal height for an invert depth > 0.5D

- 275 -

If is infinitely small, the volume change for every d is equal to the volume of the

blue cylinder.

RdCosRV sin2 9.2

max

0

hdVVh 9.3

where,

dzRRdV cossin2 9.4

max

0

coscossin2

RdRRVh 9.5

max

0

23 sincos2

dRVh 9.6

max

0

3 sincos21

212

dRVh 9.7

max

0

3 sin213sin

21

21sin

212

dRVh 9.8

max

0

3 3sin41sin

412

dRVh 9.9

max

0

3

313cos

121cos

412

RVh 9.10

where V is the volume of the blue cylinder, described in Equation 9.1.

max

0

2

3

2cos12

313cos

121cos

412

R

Rh 9.11

max

0

2cos1313cos

121cos

414

Rh 9.12