You are on page 1of 27

Geo-Congress 2014 Keynote Lectures, GSP 235 ASCE 2014

Sustainability in an era of increasing energy demand: challenges for offshore


geotechnics
Mark J. Cassidy1, Conleth OLoughlin2, Christophe Gaudin3, Melissa Landon
Maynard4
1
Professor, Centre for Offshore Foundation Systems and ARC CoE for Geotechnical Science and
Engineering, Uni. of Western Australia, Perth, Australia; mark.cassidy@uwa.edu.au
2
Associate Professor, Centre for Offshore Foundation Systems and ARC CoE for Geotechnical
Science and Engineering, Uni. of Western Australia, Perth, Australia; conleth.oloughlin@uwa.edu.au
3
Professor, Centre for Offshore Foundation Systems and ARC CoE for Geotechnical Science and
Engineering, Uni. of Western Australia, Perth, Australia; christophe.gaudin@uwa.edu.au
4
Assistant Professor, Civil & Environmental Engineering, University of Maine, Orono, Maine, USA,
melissa.L.maynard@maine.edu

ABSTRACT: The worlds escalating demand for energy, combined with the
depletion of oil reserves in shallow waters and traditional regions, is resulting in the
move of offshore developments into deeper waters, new development regions and
transformation to cleaner natural gas and renewable energy sources. Summarized in
this paper are the geotechnical challenges facing the offshore industry as it attempts to
sustain the worlds expanding energy demands. Representative examples of new
methodologies being used in engineering design are provided, including deep water
anchoring and mudmat systems, installation of mobile jack-up platforms in the stratified
soils that are often encountered in new development regions around Australasia, and the
potential use of caissons for floating wind farms.
INTRODUCTION
In an era of escalating energy demand, securing long-term resources is one of the
major challenges of our generation. Together with a need to mitigate increasing CO2
emissions and climate change, engineers are faced with the additional challenge of
conversion to low-emission energy sources. The worlds oceans hold significant
potential for solutions. However, discovery of these reserves of cleaner natural gas
and offshore renewables requires new approaches. Geotechnical engineering has
significantly contributed to the development of offshore energy reserves in the past.
However, significant challenges remain, with safe and efficient technologies required
to unlock future energy resources in our extensive marine environment.
This paper provides an analysis of current energy trends and the potential role of
geotechnical engineering in providing novel solutions in the offshore environment.
The paper also discusses examples of applications in which practical analytical and
calculation methods have been developed to move offshore energy recovery into
deep waters, new regions, and for technology transfer to cleaner sources. The focus
of the paper, rather than a comprehensive review of all areas, is on critical examples
of possible solutions and challenges to offshore geotechnics applications.

Page 1

Geo-Congress 2014 Keynote Lectures, GSP 235 ASCE 2014

AN ERA OF ESCALATING ENERGY DEMAND


The worlds demand for energy continues to increase unabated, with the total
supply doubling from ~5 600 MTOE in 1971 to over 12 700 MTOE in 2010 (Figure
1, after IEA, 2012a). This energy demand is causing an infrastructure and resource
boom in many developing offshore regions. For instance, US$120 billion worth of
infrastructure projects are currently under construction off the coast of Australia
(RBA 2011). Although low-emission renewable energy is highly desirable, it still
only accounts for a little over 3% of the world energy supply, with the majority of
that being a 2.3% contribution from hydroelectricity (2010 IEA data, 2012a).
14000

Energy Supply (Mtoe)

12000
10000
8000

Proportions in 2010
Oil
32.4%
Coal/Peat
27.3%
Natural gas
21.4%
Biofuel/waste 10.0%
Nuclear
5.7%
Hydro
2.3%
Geothermal/solar/wind
0.9%

6000
4000
2000

0
1970 1973 1976 1979 1982 1985 1988 1991 1994 1997 2000 2003 2006 2009 2012

Year

Figure 1. Total primary energy supply (from IEA, 2012a)


Governments worldwide are instituting carbon emission reduction strategies, and
several have set targets for a mix of renewably generated electricity. For instance, the
UK and China have set targets of 15% renewably generated electricity by 2015,
whereas Australia and Europe (averaged across the union) have a target of 20% by
2020. Thirty states of the USA have ambitious targets, such as 30% by 2015 (New
York), 30 and 33% by 2020 (Colorado and California) and 40% by 2030 (Hawaii).
Building offshore developments in deep water, remote locations, and for
renewables necessitates paradigm shifts in geotechnical design. The characteristics of
these frontiers and the geotechnical challenges and examples of solutions are
summarized in Table 1. These challenges are expanded in the following sections,
with examples of solutions and challenges for each provided.
ULTRA-DEEP WATER
Challenges off the continental shelf
The depletion of known hydrocarbon reserves in traditional regions and in shallow
waters is resulting in the movement of exploration and development to deeper waters
and often into untested environments. Internationally, in the Gulf of Mexico, West
Africa and offshore Brazil, developments have proceeded off the continental slope
into water depths approaching 3000 m, with prospects beyond 4000 m currently
considered by Total in West Africa. These deep water environments are typically
characterized by soft, lightly overconsolidated, fine-grained sediments.
Page 2

Geo-Congress 2014 Keynote Lectures, GSP 235 ASCE 2014

Table 1. Challenges for offshore geotechnics due to increasing energy demand


Deep water
New regions
Characteristic > 1000 m; Typically soft, lightly
Australasia: often highly stratified
environment overconsolidated fine-grained sediments seabeds, carbonates and sensitive silts
Example
challenges

Example
geotechnical
solutions

In-situ characterization (strength, upper


1-2 m layer, wide spatial area,
sensitivity)
Mobility of infrastructure:
- disturbance/healing of soil
- changes in seabed topography
Transformation in soil properties
- under episodic cyclic loading
Fatigue based design (such as risers,
SCRs, pipelines)
Uplift capacity
Geohazard risk assessment
Larger reservoir (50x50 km), requiring
more extensive soil characterization
Improved in-situ technology (largediameter piston coring, continuous Tbar and ball penetrometers)
Failure envelopes written directly in
VHM load space
Improved large deformation analysis
capabilities
New anchor configurations (such as
torpedo and OMNI-max)
SCR-soil interaction models that
account for soil stiffness and erosion

Renewables
Waters < 200 m, close to shore for
connection to the grid, high energy sites,
higher prevalence of sandy soils
Less offshore experience
Large site area to characterize (often
Can be remote with long distances (for
with variable geological features)
example, > 300 km pipelines in Aust.) Light vertical loads with large horizontal
Prediction of drainage and rate effects
and moment loads
Partial drainage conditions
Stringent serviceability requirements
High compressibility
Cyclic loading:
Occasional cementation
- non-coherent wind/wave
Large changes of strength with cyclic
- accumulative rotations and stiffness
loading
degradation
Long pipeline tied back to shore:
Cost pressures
crossing of high scarp
High monitoring and servicing costs
Foundations represent ~15-30% of total
cost (increase with depth).
Needs cheaper site-investigation
Improved numerical methods that
Utilizing previous experience from the
account for rate effects, soil softening
oil and gas industry
through shearing, high sensitivity and Statistical methods for site
strength gain through reconsolidation
characterization with integrated
Improved pipeline stability models
geophysical/geological data
Increased emphasis on direct prediction Testing under large numbers of cycles
methods from CPT-penetrometers
and multidirectional loading
Punch-through models that account for Purpose-built installation vessels
soil dilatancy
Deployable for maintenance at shore
(for example, floating wind)
Shared anchor points

Page 3

Geo-Congress 2014 Keynote Lectures, GSP 235 ASCE 2014

The nature of geotechnical design has been radically changed by the unique
conditions in deep water environments. Developments usually consist of a moored
floating facility and pipes called risers for transmitting oil or gas to/from the seabed,
often to a flowline or pipeline system (Figure 2). There is a reduced emphasis in the
traditional areas of piles and large gravity-base foundations, in which designs are
focused on ultimate capacity and lack of foundation movement. New infrastructure is
being designed for mobility, whether during installation or service, and the challenge
for geotechnical engineering is to evolve to incorporate transformations in geometry
and soil material properties. There has been a rapid evolution of new anchoring
systems, such as suction caissons, plate anchors (either suction embedded or dragged
in) and dynamically embedded torpedo anchors. These systems can require long
installation paths and face challenges of cyclic and sustained loading capacity over
many years of service. There also has been a parallel demand for efficiently designed
shallow foundations for pipeline terminations and manifolds.

Figure 2. Geotechnical infrastructure development in deep water


Example geotechnical solution: suction embedded plate anchors
An important deep water solution that responds to the need for anchors to
withstand significant vertical loading is the Suction Embedded PLate Anchor
(SEPLA). The concept combines the advantage of quick installation at a known
penetration depth and location through use of a suction caisson with the efficiency
and low cost of a plate anchor (Dove et al., 1998). SEPLAs can resist a high vertical
load component in a taut or semi-taut leg mooring configuration; therefore, they can
avoid the large offshore installation spread associated with alternate suction piles.
The plate anchor is housed initially within a slot at the tip of a suction caisson
(Figure 3). The anchor is embedded by pumping out the water inside the suction
caisson. When the targeted depth is reached, the pump flow is reversed, and the
caisson is retrieved, leaving the SEPLA in place in a vertical orientation. The slack
mooring chain attached to the anchor padeye is then tensioned, causing the SEPLA
to rotate to an inclination that is, for a symmetric anchor, approximately normal to
the local chain orientation. This process is known as keying and is conducted
Page 4

Geo-Congress 2014 Keynote Lectures, GSP 235 ASCE 2014

offshore with minimal delay after installation. During the keying process, the anchor
moves vertically upwards in the soil. As most deep water seabeds have a profile of
increasing strength with depth, keying results in a detrimental loss of capacity.
Although SEPLAs are established technology in the temporary mooring of floating
offshore structures, their application to permanent mooring has only been considered
recently because of uncertainties related to their performance:
(i) during keying - how to predict movement and loss of embedment;
(ii) under sustained loading - how to determine the loss or gain in capacity
associated with the combined effects of consolidation and strain softening;
(iii) under cyclic loading - how to predict degradation under different cyclic levels.
These issues have been recently addressed through an integrated program of
physical model tests in a geotechnical centrifuge, large deformation finite element
modeling and analytical developments (Randolph et al., 2010; Cassidy et al., 2012;
Wong et al., 2012).
suction installation
caisson retrieval
anchor keying
mobilized (sustained/cyclic)

Figure 3. Typical SEPLA used in deep water (after Gaudin et al., 2006)
The bearing capacity of a deeply embedded rectangular anchor is well established
(see Randolph et al, 2010 for example). The challenge of estimating the final anchor
capacity is prediction of the loss of embedment during the keying behavior (the final
depth to consider the undrained shear strength).
Loss of embedment has been investigated both experimentally (OLoughlin et al.,
2006; Gaudin et al., 2009, 2010; Song et al., 2009) and numerically with large
deformation finite element analysis (Wang et al., 2010, 2011; Tian et al., 2013). The
results show the importance of load eccentricity (defined by padeye location, see ep
in Figure 4), anchor thickness and weight. Loss of embedment has been summarized
for relevant non-dimensional groups in a non-dimensional equation by Wang et al.
(2011). However, more recently, a macroelement force-resultant approach has been
used to provide predictive capabilities of the keying process (Lowmass, 2006;
Cassidy et al., 2012; Yang et al., 2012). The entire anchor and surrounding soil is
considered as one element, with behavior expressed directly in terms of the anchors
loads and displacements within a plasticity framework. A yield surface represents the
capacity of the anchor at a certain embedment. Small strain finite element analyses
have been used to define the surface (square and rectangular plates: ONeill et al.,
2003; Elkhatib and Randolph, 2005; an anchor with shank: Wei et al. 2013). By
assuming perfect plasticity and associated flow, the load-displacement path is
predicted through simple differentiation of the yield surface by a conventional
plasticity approach (see Cassidy et al. 2012 for numerical implementation). Example
Page 5

Geo-Congress 2014 Keynote Lectures, GSP 235 ASCE 2014

trajectories calculated for a 4.64 m high SEPLA anchor are shown in Figure 4 for
various padeye offsets. The anchor was predicted to eventually dive for padeye
locations approximately 0.25 m below the center of the anchor.
These experimental, numerical and analytical developments also highlighted that
the keying flap found on most SEPLA anchors does not open during the initial stages
of keying, as SEPLAs slide backwards with soil resisting the hinge (Figure 4). The
flap does not perform the function for which it was intended: discouraging vertical
motion and loss of embedment. Furthermore, the flap eventually opens during the
subsequent translational phase as the anchor approaches its ultimate capacity, thus
reducing the final capacity with a dog-leg in the anchor. An improved design that
initially opens (reducing sliding and loss of embedment) and then closes (maximum
holding capacity) can be achieved by simply swapping the hinge location to the other
side of the anchor (the back), as shown by Tian et al. (2013).
0.8

Loss of embedmentz/B

0.7

direction of movement
post-keying

ep = 0.246 m

0.6

ep = 0 m

0.5

ep = -0.246 m
0.4

ep = -0.492 m

0.3
0.2

ep = -1.1 m

0.1
0
-0.2

-0.1

0.1

0.2

0.3

0.4

0.5

Horizontal displacementx/B

Figure 4. Prediction of anchor trajectory for a plate anchor loaded at 40 to the


horizontal with various padeye offsets (after Cassidy et al., 2012)
Centrifuge experiments have also evaluate the capacity of a plate anchor under
sustained and cyclic loading during operation (details in Randolph et al., 2010 and
Wong et al., 2012). In a series of six sustained load tests, the reduction in capacity
with uplifting displacements and slow displacement rate were balanced by the
increase in shear strength due to consolidation of the soil in front of the anchor. For
sustained loads at less than ~80% of the peak monotonic load, the increase in
capacity resulting from consolidation compensated for any reduction in capacity due
to low strain rates within the soil (Wong et al., 2012), as shown in the summary
results presented in Figure 5. At the point of 100% consolidation, the maximum
anchor displacement (measured in the experiments as the chain displacement) was
4.5 mm (or 15% of the total model anchor height of 30.9 mm). However, for the two
tests where the sustained load was above 80% of the monotonic capacity, the anchor
showed accelerating displacements, with the anchor no longer able to hold the load
after 5 to 18 days of prototype operations (25 to 70 s in the model centrifuge tests at
150 g). In summary, the two opposing phenomena of (i) strength gain due to
consolidation and dissipation of pore pressures ahead of the anchor and (ii) reduction
in operative shear strength due to slow displacement rates are balanced, for a
sustained loading of approximately 85% of the monotonic capacity (Wong et al.,
2012). Any sustained loading above this level led to an anchor failure.
Page 6

Geo-Congress 2014 Keynote Lectures, GSP 235 ASCE 2014

It was argued by Wong et al. (2012) that similar mechanisms occur, though with
additional damage, under cyclic loading. For any loading where the cyclic peak
loaded exceeded 75% of the monotonic capacity, anchor failure was measured. This
limit was suggested as a design basis and corresponds to an equivalent monotonic
bearing capacity factor of 7.

Peak load ratio, T m ax-sus /T m ax (-)

Peak load ratio, T max-sus/T max (-)

2
1.8
1.6
1.4
1.2
1
0.8

2
1.8
1.6
1.4
1.2
1
0.8

10

20

30

40

50

60

70

80

90

100

Level of sustained loading, T sus/T max (%)

10

20

30

40

50

60

70

80

90

100

Consolidation degree, U (%)

Figure 5. Increase in anchor capacity, followed by sustained loading (after


Wong et al., 2012)
Example geotechnical solution: torpedo anchors
Dynamically installed anchors (Figure 6) are torpedo-shaped and are designed so
that, after release from a designated height above the seafloor, they will penetrate to
a target depth in the seabed by the kinetic energy gained through free-fall and
through the self-weight of the anchor. The ease of installation makes them an
attractive anchoring solution in deep water, particularly for mobile drilling units. The
majority of installations have been in Brazilian waters with anchors that are 12 m
long and 422 kN in weight for temporary facilities and 17 m long and 961 kN in
weight for permanent installations (Araujo et al., 2004). Anchor impact velocities
depend on anchor release height but are typically between 20 and 30 m/s, with
anchor tip embedment depths of up to 3 times the anchor length.

Figure 6. Dynamically installed anchor (after Araujo et al., 2004)


Anchor capacity is governed by the vertical capacity for load inclinations greater
than ~30 relative to the horizontal (Randolph et al., 2011). Vertical capacity may be
predicted adequately using approaches for piles in clay (e.g. Lieng et al., 2000;
Page 7

Geo-Congress 2014 Keynote Lectures, GSP 235 ASCE 2014

OLoughlin et al., 2004; Richardson et al., 2009), although consideration must be


given to the reduced strength available just after installation. Richardson et al. (2009)
showed that this value was approximately 6% of the long term capacity, compared
with 25-45% for piles and suction caissons in normally consolidated clay (Esrig et al.,
1977; Bogard and Matlock, 1990; Chen and Randolph, 2007) and that the regain in
strength at the anchor interface may be quantified using a cavity expansion model.
It follows that, as for SEPLAs, the key challenge in prediction of anchor capacity is
calculation of the anchor embedment depth and hence the available shear strength in
the vicinity of the embedded anchor. This calculation is complicated by (i) 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 and Randolph, 2006) and
seven orders of magnitude higher that the standard laboratory testing rate of 1%/h
(2.8 10-6 s-1), and (ii) hydrodynamic effects associated with the very soft viscous
clay at shallow penetration and possible entrainment of a boundary layer of water
adjacent to the anchor (OLoughlin et al., 2013).
Approaches to predicting anchor embedment depth (examples in True, 1976;
OLoughlin et al., 2004, 2009, 2013; Audibert et al., 2006) consider Newtons
second law of motion and the forces acting on the anchor during penetration.
OLoughlin et al. (2013) consider such an approach and use a large database of
centrifuge tests on freefall projectiles to calibrate their model:

d 2z
Ws Fb R f (Fbear Ffrict ) Fd
dt 2

(1)

where m is the anchor mass, z is depth, t is time, Ws is the submerged anchor weight
(in water), Fb is the buoyant weight of the displaced soil, Rf is a shear strain rate
function, Fbear is bearing resistance, Ffrict is frictional resistance and Fd is inertial drag
resistance. The important dependence of shear strength on strain rate is generally
quantified using either a power or logarithmic function (Biscontin and Pestana,
2001):


R f 1 log
ref


or R f

ref

(2)

where and are strain rate parameters in the respective formulations, is the
strain rate, which may be approximated as the ratio of the anchor penetration velocity,
v, to the anchor diameter, d, and ref is the reference strain rate associated with the
reference value of undrained shear strength.
There are limited in situ anchor data in the public domain to calibrate Eq. 1.
Calibration has mainly been achieved using centrifuge tests in kaolin clay
(OLoughlin et al., 2004, 2009, 2013; Richardson et al., 2006), as shown by Figure 7.
Figure 7 justifies the inclusion of drag resistance (Fd in Eq. 1), as the predictions that
include Fd are in slightly better agreement with the tests. Whilst Figure 7 indicates
that the magnitude of the strain rate parameter, , used in the power formulation
Page 8

Geo-Congress 2014 Keynote Lectures, GSP 235 ASCE 2014

increases with anchor impact velocity, the strain rates in the centrifuge tests are
approximately 200 times that in the field, as the velocities are equivalent whilst the
diameters are scaled by 1:200. Hence, the lower bound strain rate parameter, ~0.07,
would be more appropriate for field cases and would therefore be similar to
parameters deduced from variable rate penetrometer tests: = 0.06-0.08 (Lehane et
al., 2009). Despite the very limited field data for dynamically installed anchors in the
public domain, evaluation of a deep ocean nuclear waste disposal option in the 1980s
included a number of field trials in which various projectiles were dynamically
embedded in the seabed in a manner similar to that for dynamically installed anchors.
The most successful and well documented trials took place in the Atlantic Ocean at
Great Meteor East, which is an area at the western extremity of the Madeira Abyssal
Plain (~800 km west of the Canary Islands) in water depths of ~5000 m (Freeman
and Burdett, 1986). The merit of Eq. 1 is demonstrated by Figure 8, which compares
predicted and measured velocity profiles. The parameters adopted for the field
predictions were consistent with the centrifuge study, with the interface friction ratio
taken as the inverse of the soil sensitivity and the rate parameter taken as the lower
bound, = 0.07, as discussed previously. The agreement in Figure 8 is encouraging,
with final embedment depths that are consistently within 4% of the measurements.
Velocity (m/s)
0

10

15

20

25

30

= 0.074
= 0.068

Anchor embedment (mm)

50
= 0.117
= 0.104

100

= 0.090
= 0.079
= 0.120

150

= 0.128
= 0.118
= 0.125

200

Prediction A (drag included)


Prediction B (drag omitted)
A4: m = 5.4 g, 0 mm drop height (drum)
A4: m = 5.4 g, 50 mm drop height (drum)
A4: m = 5.4 g, 200 mm drop height (drum)
A1: m = 14.8 g, 100 mm drop height (beam)
A1: m = 14.3 g, 150 mm drop height (beam)
A1: m = 14.3 g, 200 mm drop height (beam)

= 0.113
= 0.120

250

Figure 7. Predicted and measured velocity profiles from centrifuge tests


(after OLoughlin et al., 2013)
A first-order approximation of anchor embedment depth can be obtained by
considering the total energy of the anchor as it impacts the seabed (OLoughlin et al.
2013). Total energy is defined as the sum of the kinetic and potential energies
(relative to the final embedment depth) of the anchor at the moment of impact:
E total

1
2

mv i2 m' gz e

(3)

where m' is the effective mass of the anchor, vi is the impact velocity, g is Earths
gravitational acceleration and ze is the final embedment depth. Centrifuge and field
data are shown in non-dimensional form (removing the influences of diameter and
soil strength gradient, k) in Figure 9. The collective data form a relatively tight band,
considering the assumptions included shear strength gradients and anchor geometries
for the field data. A conservative fit to the data may be expressed as:

Page 9

Geo-Congress 2014 Keynote Lectures, GSP 235 ASCE 2014

E
ze
0.1 total
kd 4
d eff
eff

10

1/ 3

(4)

where deff is effective diameter. This relationship harmonizes a very large data set
that encompasses a wide range of anchor masses, geometries and impact velocities.

Figure 8. Predicted and measured velocity profiles from free-fall


penetrometer tests at Great Meteor East (after OLoughlin et al., 2013)

Figure 9. The use of total energy as a means of comparing dynamically


installed anchor embedment data (after OLoughlin et al., 2013).
Example geotechnical solution: mudmats
Subsea mats are used in deep waters as foundations on soft and normally
consolidated (or lightly overconsolidated) clay to support facilities such as pipeline
terminals, jumpers, riser bases and manifolds. They are subjected to combinations of
loading in all six degrees of freedom. In most cases, the mat may not provide
sufficient resistance against sliding or overturning failure. Pinned piles may then be
used in each corner of the mat to increase the capacity of the foundation (Figure 10).
Although mudmats with pinned piles have already been used in-situ, notably by BP
in the Gulf of Mexico, the behavior of the foundation is still uncertain, and their

Page 10

Geo-Congress 2014 Keynote Lectures, GSP 235 ASCE 2014

11

design remains challenging. The effect of the corner piles on foundation behavior
and capacity have been quantified only recently via a program of centrifuge model
tests (Gaudin et al., 2012), finite element analyses and analytical developments
(Dimmock et al., 2013). The results demonstrated that an increase in capacity is
obtained by supplementing the mat foundation with short piles placed at the corners.
A smaller foundation footprint can then resist the subsea foundation design loads.
The tests showed that the piles significantly increased the size of the yield envelope
in M-H space, notably with respect to the ultimate H, changing the mode of yielding
compared to a conventional mat. Consequently, sliding resistance is increased in a
greater proportion, and overturning moment becomes a more significant mode of
yielding. Based on the experimental and numerical outcomes, a simplified lower
bound design approach was developed, whereby the mat is assumed to carry the
vertical load (as it would prior to installation of the piles); the piles are assumed to
carry the horizontal and torsional loads, and the combined mat and piles resist the
applied moments (Dimmock et al., 2013). The design approach was shown to
provide a conservative estimate of the capacity obtained from numerical analysis.
An alternative to the corner piles, which may prove challenging to install, is the use
of suction caissons installed across the mat. In addition to its self-installation ability,
a mat equipped with a single suction caisson exhibits a horizontal and moment
capacity increase of factors of 2 and 1.2, respectively (Bienen et al., 2012). The
moment capacity can be further enhanced by a caisson of sufficient length to
intercept the rotational mechanism developed by the mudmat. Although it has yet to
be used in the field, such a foundation shows great promise, and research
development is progressing to increase torsion capacity with two suction caissons.

Figure 10. Mudmat reinforced with corner piles (after Gaudin et al., 2012).
NEW REGIONS
Introduction and challenges
Installation of platforms has become challenging in a number of new(er) oil and
gas regions, such as South East Asia, India and Australia, because of the prevalence
of highly stratified seabeds, carbonate soils with high compressibility and occasional
cementation, and sensitive silts that create partial drainage conditions. These
conditions surprised engineers during the installation of the first platform on the
Page 11

Geo-Congress 2014 Keynote Lectures, GSP 235 ASCE 2014

12

North West Shelf of Australia, the North Rankin A. Installation challenges have
been particularly acute in the operations of mobile jack-up platforms, as discussed by
Osborne et al. (2006) and others, cumulating in the establishment of the InSafe Joint
Industry Project (Osborne et al., 2009, 2010). Although other infrastructure elements,
such as on-bottom pipelines, have specific issues in these soils, this paper will
concentrate on a discussion of jack-ups.
Jack-ups consist of a buoyant triangular hull, three independent truss-work legs
approximately 170 m in length and inverted conical spudcan footings approximately
20 m in diameter. Jack-ups are self-installed by lifting the hull from the water and
pushing the large spudcans into the seabed. This procedure is dangerous, with as
many as five jack-up incidents annually attributable to geotechnical failures (Hunt and
Marsh, 2004; Jack et al., 2007). These events result in rig damage and lost drilling
time, with even temporary loss of serviceability of a rig having major financial
implications. The specific geotechnical challenges to jack-up operations emerging in
these new regions are summarized in Table 2 and illustrated in Figure 11.

Figure 11. Geotechnical challenges for jack-ups encountered in new regions


Table 2. Geotechnical challenges installing jack-ups in new Australasian regions
Installation Issue

Specific Challenge

References for Solutions

Prediction of spudcan
penetration
Risk of punch-through
failures
Offshore operations to
mitigate punch-through
Reinstallation into
seabed footprints

Partial drainage in
sensitive silts
Prevalence of buried sand
layers, multi-layered soils
Use of perforation drilling

Erbrich, 2005; Cassidy, 2012; Osborne et al.,


2010
Edwards and Potts, 2004; Lee et al., 2013a,b; Teh
et al., 2009; Hossain and Randolph, 2009
Brennan et al., 2006; Chan et al., 2008; Hossain et
al., 2010
Dean and Serra, 2004; Leung et al., 2007; Cassidy
et al., 2009; Gan et al., 2012; Kong et al., 2013

Eccentric loading on
spudcans and legs

Example geotechnical solution: prediction of spudcan punch-through failure


The potential for unexpected punch-through failure of a jack-up exists during
installation and pre-loading in layered soils. This failure occurs when the spudcan
uncontrollably pushes a locally strong zone of soil into underlying softer material.
Page 12

Geo-Congress 2014 Keynote Lectures, GSP 235 ASCE 2014

13

Such failures can lead to buckling of the jack-up leg, effectively temporarily
decommissioning the platform and even toppling the unit.
A thin layer of sand (less than a spudcan diameter) overlaying a weaker stratum of
clay is particularly hazardous, and a new calculation method is discussed below that,
for the first time, takes into account the stress level and dilatant response of sand.
Other soil conditions that can cause rapid leg penetration include stiff-over-soft clay
layers (see Edwards and Potts, 2004 and the new calculation methodology of Hossain
and Randolph, 2009), a thick clay layer whose strength decreases with depth, and a
very soft clay in which the rate of increase of bearing capacity does not match the
loading rate and firm clay with sand or silt pockets (Osborne et al., 2009).
Capturing spudcan failure mechanisms by recording digital images of a halfspudcan installed against a transparent window (analyzed using particle image
velocimetry (PIV) coupled with close-range photogrammetry corrections) has
provided the basis of new analytical models. As shown in Figure 12, Teh et al.
(2008) provided visual evidence that the peak punch-through resistance in sand-overclay cases occurs at a relatively shallow embedment between ~0.12 and 0.15D (also
confirmed by Lee et al., 2013a and Hu et al., 2013) and was caused by a failure
mechanism that consists of a truncated cone of sand being forced vertically into the
underlying clay layer. The truncated cone changes from pure vertical movement
(directly under the spudcan) to radial and small amounts of heave further from the
center line. The outer angle of the sand frustum is forced into the clay and reflects the
dilation of the sand.
qpeak
q0
Sand

z
Hs

' ' s
qclay

Clay

Figure 12. PIV image at punch-though recorded in a centrifuge (after Teh et al.,
2008) and a conceptual mechanism of the recently proposed model
The assumed mechanism at the peak resistance of the conceptual model proposed
by Lee et al. (2013b) is also shown in Figure 12. In this mechanism, a sand frustum
with a dispersion angle equal to the dilation angle is assumed to be pushed into the
underlying clay. The footing pressure and weight of the sand frustum are resisted by
the frictional resistance along the sides of the sand block and in the bearing capacity
of the underlying clay (Lee et al., 2013b).
A design equation for the peak resistance qpeak has been derived by treating the
conceptual sand frustum as a series of infinitesimal horizontal discs and equating
vertical force equilibrium as in a silo analysis. The detailed derivation and final
design equation can be found in Lee et al. (2013b). In the method the stress level and
Page 13

Geo-Congress 2014 Keynote Lectures, GSP 235 ASCE 2014

14

dilatant response of the sand is calculated using an iterative approach. The


capabilities of this new method to predict results from a database of 47 centrifuge
tests, such as those shown in Figure 13, is provided in Lee et al. (2013b) and Hu et al.
(2013). These provide evidence that the existing punching shear and load spreading
models in current jack-up industry guidelines (such as SNAME, 2008) significantly
and consistently under-predict qpeak. Recent advances reported by Hu et al. (2013)
have extended Eq. 5 to account for the depth of penetration under medium to loose
sand conditions and updated the expression of the distribution factor to account for
different spudcan shapes.
0

200

Depth (m)

Sand

6m diameter
flat foundation
8m dia.
10m dia.
12m dia.
6.2m sand thickness
14m dia.

1000

Clay

8
10
12

Punching shear - SNAME (2002)

800

qpeak,calculated (kPa)

16m dia.

4
6

1200

Equation 1

0
2

Penetration resistance (kPa)


400
600
800
1000

Load spread 1h:5v - SNAME (2002)

600

400

200

14
16

18

0
Series of D1F30a to D1F80a

20

200

400

600

800

1000

qpeak,centrifuge (kPa)

Figure 13. Example results of sand-over-clay penetration of a flat-based footing


and comparison of numerical models (modified after Lee et al., 2013a,b)

MARINE RENEWABLE ENERGY


The challenge of lowering the cost of low-emission, sustainable energy
Regions with significant offshore wind, wave and tidal resources, such as countries
along the western coasts of Europe, Australia, Japan and the USA, have an
opportunity to derive much of their renewably generated electricity from the ocean.
The most rapid development to date has been the installation of offshore wind
turbines, though it remains a fraction of the amount of onshore wind capacity (Figure
14). The International Energy Association (IEA) forecasts that offshore wind will
continue to scale significantly over the next few years to 25.9 GW in 2017, up from
6.1 GW in 2012 (IEA, 2012b). Equivalent projections for offshore wave and tidal
energies are much lower, as these technologies are much less mature.
Most of the installed offshore wind capacity is in European waters, particularly that
of the UK (forecast to grow to 7.4 GW by 2017, Crown Estate, 2013, IEA, 2012b).
China is expected to be the second-largest offshore wind producer by 2017, with a
forecasted 7.0 GW installed capacity by 2017 (IEA, 2012b). Much of the current
growth in offshore wind is driven by government incentives, which are required to
scale the technology but unsustainable in the long term. Future growth beyond the
current IEA projected periods must come from market demand, which will

Page 14

Geo-Congress 2014 Keynote Lectures, GSP 235 ASCE 2014

15

30

Onshore wind

450

25

400

Offshore wind

350

20

300
15

250
200

10

150
100

50
0

1996
1997
1998
1999
2000
2001
2002
2003
2004
2005
2006
2007
2008
2009
2010
2011
2012
2013
2014
2015
2016
2017

Cummulative onshore capacity (GW)

500

Cummulative offshore capacity (GW)

necessitate technical innovation to drive down the installation and operation costs so
that offshore wind can be competitive with alternative technologies.

Figure 14. Cumulative installed wind capacity (data from IEA and EWEA)
In these offshore wind developments, monopile and gravity-based foundations are
favored due to shallow waters (<30 m), from direct knowledge transferred from the
offshore oil and gas industry. Tripod structures and lattice frames are being
considered within depths of 30-50 m, and floating turbines moored to the seabed are
considered beyond 60 m. There are floating wind turbine concepts at varying stages
of development. The most advanced amongst these include the Hywind spar concept,
which has been in operation in 198 m water depth off the southwest coast of Norway
since 2009 (with advanced plans for small integrated arrays of four of five floating
turbines off the coast of Scotland and the US), and the WindFloat semi-submersible
concept that has been tested in 40 - 50 m water depths off the coast of Portugal since
2011. Various other consortia have planned installations of one or more floating
turbines in Japanese, European and US waters over the coming two to three years.
This increased activity is expected to demonstrate concept feasibility, allowing
attention to turn towards reduction of installation costs for fully commercial farms.
Floating wind turbines are a paradigm shift in technology for the future, even for
shallow water, as they (i) reduce the cost of foundations though anchor arrays, (ii)
significantly reduce the cost for maintenance as turbines can be towed to shore, and
(iii) limit the high cost of scour protection commonly encountered in shallow water.
If floating wind turbines can become economical, access to the worlds best wind
resources can be achieved. For instance, 61% of the USAs offshore wind resources
are in water depths greater than 100 m (estimated resources in Figure 15), almost all
of Japans offshore wind resources are in deep water and the recent Round 3
development sites released by the UK Crown Estate include water depths up to 63 m.
Geotechnical challenges
Current mooring and anchoring technology has been shaped by the offshore oil and
gas industry (O&G), with floating facilities in water depths approaching 3,000 m.

Page 15

Geo-Congress 2014 Keynote Lectures, GSP 235 ASCE 2014

16

The mooring/anchoring requirements for wind, wave and tidal/current facility will be
different in several ways and pose the following challenges:
1. Economies of Scale: Economic returns for installed floating renewable energy
technologies are much lower than for an O&G installation. The level of platform,
mooring, and anchoring infrastructure required for renewable energy facilities
(especially wave and tidal/current) is prohibitive when using current mooring and
anchoring systems employed by the offshore O&G industry. Anchors and moorings
are a very small proportion of the cost of energy production for O&G platforms. For
instance, 18% of installed wave energy costs are due to moorings and anchors
compared with 2% for an installed floating oil production facility (Fitzgerald, 2009).
There is therefore potential for geotechnical cost savings in the renewable sector and
development of new technologies must be cost driven.

Figure 15. Offshore wind resources in the United States by region and depth
for annual average wind speed sites above 7.0 m/s (after Musial and Ram, 2010)
2. Design Requirements: Currently, many of the design guidance documents being
developed for offshore renewable energy technologies are heavily based on O&G
industry experience and guidance documents. To become economically viable,
however, the offshore renewable energy industry will be motivated to take a much
less conservative approach. This is justified as renewable facilities have lower
consequences of failure due to being unmanned, of lower capital investment and as
they do not carry environmentally hazardous oil and gas.
3. Optimized
Environmental
Interaction:
Floating
renewable
energy
platforms/devices are designed to interact with the wind, wave and currents for
optimized energy extraction. Simultaneously, these devices are designed to maximize
design life under millions of operation environmental loading cycles with minimal
loss of infrastructure during extreme storm events. For instance, a wave energy
converter must be permitted to move relative to the waves so that power can be
extracted, whilst also keeping on station during an extreme storm. Because wave and
tidal/current energy converters will be sited in high energy sea states, cyclic loading

Page 16

Geo-Congress 2014 Keynote Lectures, GSP 235 ASCE 2014

17

variable in magnitude and spatial orientation will be more prevalent for these
facilities during their operational life than for a floating O&G facility that may only
experience relatively significant cyclic loading during extreme storm events.
Optimized device interaction within its high energy environment requires
investigation of the coupling of cyclic and multi-directional loading. This is a rarely
considered and will present geotechnical challenges, particularly for sands.
4. Mooring System: The choice of a mooring system represents a cost trade-off
between anchoring complexity and device operation. Taut-leg mooring systems have
the advantages of smaller footprints and shorter moorings lines. However, the high
capacity mooring system required for stability under cyclic environmental loading
presents a complex foundation-substructure interaction problem. Taut leg systems
require anchors that can withstand significant components of vertical load and few
anchors that are suitable for sand meet this requirement. The rational against mooring
stabilized designs likely relates to the greater complexity of design required for taut
and tension mooring systems that rely on seafloor sediments for regular device
operation over the full design life as well as during extreme events. This is in direct
contrast to platform stabilized design that rely on anchors mainly for extreme events.
Musial et al. (2006) maintain that floating wind will have lower costs if new drag
embedment and vertically loaded anchors (VLAs) are utilized and mooring line
stabilized designs are avoided. For this, catenary mooring systems typically
comprised of simple, low cost anchors (albeit with long mooring lines) are used.
However, with this system, substructures are subject to greater wave loading and
motion, which increases the complexity and cost of wind turbine design (Butterfield
et al. 2005). However, parametric studies of various floating wind concepts
conducted by Sclavounos et al. (2008) using fully coupled dynamic numerical
simulations found that wind turbine nacelle accelerations (housing covering gearbox
etc) varied as a function of both substructure geometry and mooring configuration.
They found that high nacelle accelerations hasten deterioration of turbine gearbox
and blades, and results showed accelerations were least for tension leg platforms,
greater for slack catenaries, and were greatest for taught catenaries.
While a catenary mooring arrangement appears to be more suitable for floating
renewable energy devices/platforms in order to allow necessary environmental
interaction (e.g., wave, current) and reduce the required contribution from seafloor
sediments to extreme or infrequent loading events, they are not optimal with regard
to long term design and performance of wind and possibly tidal turbines or
minimizing the seabed footprint (which will be discussed).
5. Platform Excursion Limitations: The seabed footprint area covered by the
mooring arrangement for each floating device must be small to avoid the interaction
and abrasion of mooring lines against each other and minimize excursion and
seafloor abrasion of power umbilicals associated with each individual device. This
arrangement is in contrast to typical offshore O&G platforms, where the footprint
area is mostly unrestricted and the maximum allowable excursion is defined by the
allowable movement of the risers.
6. Number of Anchors: A floating O&G facility is a single unit that is typically
positioned using a 12- or 16-point mooring arrangement. A renewable energy facility

Page 17

Geo-Congress 2014 Keynote Lectures, GSP 235 ASCE 2014

18

could include up to 200 devices, each of which is likely to be positioned using either
three- or four-point mooring systems. Thus, the number of anchor installations is
significantly higher than for an O&G facility, necessitating the need to develop more
efficient anchoring technologies that can also share anchors between multiple
devices/platforms.
The large number of mooring lines will have significant implications for permitting.
Multi-use areas that include fisheries and marine wildlife habitat and migration paths
will be more affected than for seabed fixed renewable energy or O&G facilities.
While anchor sharing will help to reduce cost, development of floating systems
where multiple devices are linked closer to the sea surface with fewer mooring lines
going to the seafloor is a valid option for reducing the impact of the facilities. This is
an approach often taken in aquaculture.
7. Anchor Innovation: Currently, drag embedment anchors are preferred for catenary
moored floating offshore wind platforms (e.g. Hywind and WindFloat) because of
their low cost and simplicity of design and installation. These anchors, however,
cannot be used for anchor sharing or multi-directional loading that will be required
for more cost effective renewable energy facilities. Innovative anchoring systems
have been employed for O&G facilities, e.g., dynamically installed anchors and
SEPLAs. However, these are specifically designed for clay deposits that dominant
deep water O&G fields and there has not been equivalent innovation in anchor
designs for sand. This poses a challenge for both wave and floating wind facilities, as
these are likely to be initially sited in shallower water where the depositional
environment is energetic and coarse-grained seabed deposits are expected.
8. Seafloor Area: The seabed footprint occupied by a typical renewable energy
facility is extremely large compared with an O&G installation, which has significant
implications for site investigations.
9. Subsurface Variability: Spatial variability of seafloor sediments is prevalent in
glacially formed sediments. For instance, glacially affected sediments offshore of the
North Atlantic can include some or all of the following sediments types, in depth
descending sequence (or may be interlayered): 1) recent, soft clays and loose sands,
2) varied marine and fluvial marine sediments of significant and variable thickness,
layering and lateral extent, 3) variable glacial deposits (till, lacustrine, moraine and
outwash) with varied strength, thickness, layering and lateral extent, and 4) bedrock.
An IEA analysis of the experiences from five operational European wind farms (IEA
2005) emphasizes: a) the importance of seabed soils and their relation to design of
foundations, installation method and therefore cost for installation; and b) that site
selection and evaluation of ground conditions and morphology should be considered
from project conception as being critical to technical completion. Additionally,
incomplete and untimely definition of seabed conditions and variability adversely
affected eight offshore wind projects was an important lesson from developments of
offshore wind in northern Europe (Gerdes et al. 2006).
Due to the large footprint areas and inherent variability of seafloor geologic
conditions and engineering properties, there is greater uncertainty in encountering
unanticipated or unknown seabed conditions. This uncertainty directly affects the

Page 18

Geo-Congress 2014 Keynote Lectures, GSP 235 ASCE 2014

19

costs of the foundation design, fabrication, and installation, and has been shown to
play a pivotal role in the relative success of offshore wind projects.
Random, uniformly-spaced investigations are not an optimum layout for renewable
energy developments. The types, numbers, and depths of investigations depend on
the spatial extent of the planned development, and the properties, layering, and
lateral variability of seabed soils. The anticipated variability of the subsurface needs
to be initially assessed through a desktop study and mapped using geophysical
survey data such that boring and in situ penetration investigations can be optimally
located in typical and critical soil layers. Stuyts et al. (2011) provide an excellent
example of the use of borehole data with geostatistics and reliability methods to
quantify wind turbine risks and costs and to identify areas where additional site
investigation information is required. The ability to identify the presence and scale of
seabed features enables the optimal anchor design(s) for the project because i) it can
eliminate unnecessary conservatism; ii) allows for appropriate choice of methods and
equipment for installation; and iii) reduces operational inefficiencies and
maintenance costs resulting from unanticipated conditions.
10. Site Investigations: The cost and time associated with site investigations
required for the large areas and numbers of turbines associated with commercial
renewable energy projects is significant. Borehole drilling and in situ testing from the
deck of a large ship will no longer be time/cost effective. Additionally, the need to
site wave and tidal/current projects in high energy and possibly shallow water
locations will increase the complexity of site investigations, which will warrant the
use of specialized equipment such as vessels with dynamic positioning, seabed
frames, and high capacity jack-ups.
The use of seabed frames for deep water O&G borings and in situ testing have
been very successful. However, these devices are designed to ensure reliable remote
functionality under extreme conditions. Mini-CPTs deployed from vessels or seabed
frames are currently an option for shallow water site investigations, however, these
are limited to about 10 m of investigation depth in soft clays due to their limited
reaction force. The use of newer generation seabed frames for site investigation of
shallower water renewable energy facilities could represent significant savings.
Newer generation seabed frames would need to address the following challenges for
in situ testing and sample collection: i) penetration into sediments other than soft
clays (e.g., loose and stiff sands, stiff clays) that are likely to exist in geologically
complex nearshore environments; and ii) penetration to depths between 10 m and
50 m. Deeper penetration without the use of a drilling vessel are required for many
anchor types, but especially for drag embedment anchors and VLAs associated with
taut line and high tension mooring lines in order to gain adequate capacity while
keeping anchor size small and installation costs low.
A novel option for site investigations is to bypass the requirement for borings or in
situ tests at each location specified by offshore industry classification agencies, and
develop investigation tools that are more similar to the anchoring devices themselves.
This would allow for characterization of seafloor soil performance directly, instead
of using laboratory or in situ measured sediment parameters to correlate to
anticipated anchor performance.

Page 19

Geo-Congress 2014 Keynote Lectures, GSP 235 ASCE 2014

20

Example geotechnical solution: foundations for floating wind turbines


Floating wind turbines are classified in three categories that depend on their
method of achieving static stability with respect to pitch and roll. The three concepts
are: 1) the tension leg platform, which maintain stability essentially by the vertical
mooring line taut to the seabed and excess buoyancy; 2) the spar buoy, which
maintains stability by deep draft and a ballast and is anchored to the seabed by taut
inclined mooring lines; and 3) the barge buoy (or semi-submersible), which uses its
large water plane area and shallow draft for stability and is anchored to the seabed
via catenary mooring lines. Suction caissons are favored for all three options because
they are simple to install accurately with respect to location, orientation, and depth,
and they benefit from design and installation procedures that are well-established
after a couple of decades of use in the offshore oil and gas industry.
However, design methods for suction caissons have been established to determine
their ultimate capacity in soft homogeneous sediments. Therefore, their use for
anchoring floating wind turbines in shallow water requires the development of
renewed design methods that focus on serviceability (notably accounting for stiffness
degradation over a large number of cycles) and account for the complex soil
stratigraphy frequently encountered in shallow waters.
Research has been undertaken to address these issues (Foglia et al., 2012; Cermelli
et al., 2012). While such studies will certainly contribute to an optimization of
caisson design and promote the development of floating wind turbines, the potential
cost-savings remains marginal with respect to the total cost of the anchoring system.
Other strategies are to consider 1) a network of interconnected anchors for an array
of floating wind platforms or wave/tidal/current energy devices (Huang and Aggidis
2008; Burns et al. 2014), or 2) multi-turbine platforms that are anchored to the
seafloor with fewer anchors per turbine. These solutions have the potential to reduce
the number of caissons necessary for a floating wind facility, providing cost-savings.
However, the solution also introduces new geotechnical challenges, including more
complex multi-directional loading than encountered in deep water O&G facilities.
The multi-directional loading of suction caissons has recently been investigated by
Burns (2013) and Burns et al. (2014) in experimental and numerical feasibility
studies. Centrifuge model testing was conducted using a 1/150th scale suction caisson
(7.5 m length, 3 m diameter) in normally consolidated kaolin clay to investigate
caisson performance under loading from two orthogonal mooring lines. The caisson
was subject to (i) uni-directional 45 inclined monotonic and cyclic loading and (ii)
monotonic and cyclic loading from two 45 inclined orthogonal mooring lines under
a combination of sustained loading (54% of peak monotonic capacity) and loading to
failure. The experiments also focused on understanding lateral and rotational
stiffness degradation, accumulation of caisson displacement/rotation, and post-peak
ultimate capacity. 3D finite element numerical simulations of the monotonic
orthogonal line centrifuge test found similar trends in increased capacity, but
attributed this to non-optimal pad-eye location in the model caisson for the soil
conditions in the centrifuge tests. Burns (2013) found that numerical simulations of
monotonic orthogonal line loading over a range of mooring line inclination angles

Page 20

Geo-Congress 2014 Keynote Lectures, GSP 235 ASCE 2014

21

(0 to 90) and various sustained loading vs. loading to failure scenarios that the
resolved load at failure of the two lines was similar to the peak single line capacity at
the equivalent resolved load inclination angle. Cyclic loading also offers interesting
results. Figure 16 illustrates the load versus line displacement for both sustainedcyclic and cyclic-cyclic orthogonal line loading scenarios from Burns et al. (2014).
LMLA: SC S2T06

12

12

8
6
4
2
0

(b)

L1
L2

10
P / (su D L)

10
P / (su D L)

(a)

L1
L2

8
6
4
2

0.5

1.5

/D

0.5

1.5

/D
(c)

(d)
L2

L1

Figure 16. Caisson response under orthogonal loading: initial cyclic loading in
L1 at 35% (mean) and 15% (cyclic) of the peak monotonic capacity, followed by
(a) monotonic load to failure in L2 and (b) cyclic before monotonic load to
failure in L2. (c)/(d) illustration of soil zone of influence (after Burns et al., 2014).
As shown for SEPLA plate anchors (Figure 5 and Wong et al., 2012), the cyclic
response of caissons is a direct function of the peak cyclic load with respect to the
monotonic ultimate capacity. Below a critical threshold, cyclic loading results in an
increase in post-cyclic monotonic ultimate capacity. Interestingly, increased
monotonic capacity is also observed, though in a lesser magnitude, if cyclic loading
is undertaken first in the orthogonal direction. This result occurs because the
consolidation occurring during cyclic loading offsets the potential strength
degradation (Matsui et al. 1980), increasing the monotonic caisson capacity.
The response is different if loading is applied simultaneously in both orthogonal
directions. If monotonic loading is sustained in one direction, the post-cyclic
monotonic ultimate capacity is slightly reduced in the orthogonal direction, though
the overall caisson capacity increases when the load resultant is considered. This
result is explained by a larger zone of soil mobilized around the caisson, as illustrated

Page 21

Geo-Congress 2014 Keynote Lectures, GSP 235 ASCE 2014

22

in Figure 16. The size of the mobilized zone is shown to be a function of the caisson
geometry and soil properties but also of the mean and peak cyclic loads.
For the relatively low level of peak cyclic load considered, limited stiffness
degradation was observed during cycling loading on the subsequent monotonic to
failure load. Caisson rotations were also limited, although greater accumulated
rotation was observed for higher cyclic load amplitudes and when cycles are applied
on both orthogonal directions.
Further research is required to extrapolate these results to a wider range of caisson
geometries, soil conditions, and cyclic amplitudes, but these preliminary results are
promising and indicate that interconnected caissons are a viable solution as
anchoring systems for arrays of floating wind turbines.
CONCLUSIONS
In an era of increasing energy demands, geotechnical engineering is rapidly evolving to
meet challenges in energy extraction from water depths exceeding 1000 m and new and
remote regions with challenging soil conditions and translation of novel solutions to more
economically develop wind, wave and tidal renewable energies from the worlds oceans.
This paper endeavored to summarize the current geotechnical challenges and provide
relevant examples of geotechnical solutions. Though not an exhaustive list, these examples
included new design methods for suction installed plate anchors, torpedo anchors, subsea
mudmats and caissons subjected to orthogonal cyclic loading, and a new formula was
presented for predicting the peak capacity as a mobile jack-up spudcan foundation punches
a strong sand layer into a weaker underlying clay layer.

ACKNOWLEDGEMENTS
This work forms part of the activities of the Centre for Offshore Foundation
Systems, currently supported as a node of the Australian Research Council Centre of
Excellence for Geotechnical Science and Engineering and as a Centre of Excellence
by the Lloyd's Register Foundation. Lloyds Register Foundation invests in science,
engineering and technology for public benefit, worldwide. The authors acknowledge
and thank their colleagues Mark Randolph, Kok Kuen Lee, Yinghui Tian, Dong
Wang, Sam Stanier, and Pan Hu for their contributions to the research discussed.

REFERENCES
Araujo, J.B., Machado, R.D., Medeiros, C.J. (2004). "High holding power torpedo
pile results from the first long term application." Proc. Int. Conf. on Offshore
Mechanics and Arctic Engng, Vancouver, Canada, OMAE2004-51201.
Audibert, J.M.E., Movant, M.N., Won, J-Y. Gilbert, R.B. (2006). "Torpedo piles:
laboratory and field research." Proc. Int. Polar and Offshore Engng Conf., San
Francisco, USA, 462-468.
Page 22

Geo-Congress 2014 Keynote Lectures, GSP 235 ASCE 2014

23

Bienen, B., Raush, L., Gaudin, C., Cassidy, M.J., Purwana, O.A. (2012). Numerical
modelling of the undrained capacity of a hybrid skirted foundation under
combined loading. Int. Journal of Offshore and Polar Engng. 22(3): 1-7.
Biscontin, G. and Pestana, J.M. (2001). "Influence of peripheral velocity on vane
shear strength of an artificial clay." ASTM Geotechnical Testing Journal,
24(4):423-429.
Bogard, D. and Matlock, H. (1990)."Application of model pile tests to axial pile
design." Offshore Technology Conf., Houston, USA, OTC 6376.
Bolton, M.D. (1986). "The strength and dilatancy of sands." Gotechnique 36(1):6578.
Brennan, R., Diana, H., Stonor, R.W.P., Hoyle, M.J.R., Cheng, C.-P., Martin, D.,
Roper, R. (2006). "Installing jackups in punch-through-sensitive clays." Offshore
Technology Conference, Houston, OTC 18268.
Burns, M, Landon-Meynard, M., Davids, W.G., Chung, J., Gaudin, C. (2014).
"Centrifuge modelling of suction caissons under orthogonal double line loading."
Proc. 8th Int. Conf. on Physical Modelling in Geotechnics, Perth, Australia.
Butterfield, S., Musial, W., Jonkman, J., Sclavounos, P., and Wayman, L. (2005).
"Engineering challenges for floating offshore wind turbines," Proc. Copenhagen
Offshore Wind Conference, Copenhagen, DK: 1-10.
Cassidy, M.J. (2012). "Experimental observations of the penetration of spudcan
footings in silt." Gotechnique, 62(8): 727-732.
Cassidy, M.J., Gaudin, C., Randolph, M.F., Wong, P.C., Wang, D., Tian, Y. (2012).
"A plasticity model to assess the keying behaviour and performance of plate
anchors." Gotechnique 62(9): 825-836.
Cassidy, M.J., Quah, C.K., Foo, K.S. (2009). "Experimental investigation of the
reinstallation of spudcan footings close to existing footprints." Journal of
Geotechnical and Geoenvironmental Engineering, ASCE 135(4): 474-486.
Cermelli C.A., Rodier D.G., Weinstein A. (2012). "Implementation of a 2MW
floating wind turbine prototype offshore Portugal." Offshore Technology
Conference, Houston, USA, paper OTC23492.
Chan, N. H. C., Paisley, J. M., Holloway, G. L. (2008). "Characterization of soils
affected by rig emplacement and Swiss cheese operations - Natuna Sea, Indonesia,
a case study. " Proc. Jack-up Asia Conf. and Exhibition, Singapore.
Chen, W. and Randolph, M. F. (2007). "External radial stress changes and axial
capacity for suction caissons in soft clay." Gotechnique, 57(6):499511.
Crown Estate (2013). "Offshore wind operational report." www.
thecrownestate.co.uk/media/418869/offshore-wind-operational-report-2013.pdf
Dean, E.T.R. and Serra, H. (2004). "Concepts for mitigation of spudcan-footprint
interaction in normally consolidated clay." Proc. Int. Symp. Offshore and Polar
Engng, Toulon, France.
Dimmock, P., Clukey, E., Randolph, M.F., Murff, D., Gaudin, C. (2013). "Hybrid
subsea foundations for subsea equipment." Journal of Geotechnical and
Geoenvironmental Engineering. In Press.
Dove, P., Treu, H., Wilde, B. (1998). "Suction embedded plate anchor (SEPLA): a
new anchoring solution for ultra-deepwater mooring." Proc. Deep Offshore
Technology Conference, New Orleans, USA.

Page 23

Geo-Congress 2014 Keynote Lectures, GSP 235 ASCE 2014

24

Edwards, D.H. and Potts, D.M. (2004). "The bearing capacity of a circular footing
under punch through failure." Proc. Int. Symp. Num. Models in Geomech.,
(NUMOG), Ottawa, 493-498.
Einav, I. and Randolph, M.F. (2006). "Effect of strain rate on mobilised strength and
thickness of curved shear bands." Gotechnique, 56(7):501-504.
Elkhatib, S. and Randolph, M.F. (2005). "The effect of interface friction on the
performance of drag-in plate anchors." Proc. Int. Symp. on Frontiers in Offshore
Geotechnics (ISFOG), Perth, Australia, 171-177.
Erbrich, C.T. (2005). "Australian frontiers spudcans on the edge." Proc. Int. Symp.
Frontiers in Offshore Geotechnics (ISFOG), Perth, Australia, 4974.
Esrig, M.I., Kirby, R.C., Bea, R.G., Murphy, B.S. (1977). "Initial development of a
general effective stress method for the prediction of axial capacity for driven piles
in clay." Offshore Technology Conf., Houston, USA, OTC 2943.
Fitzgerald, J. (2009). "Position mooring of wave energy converters." PhD Thesis,
Chalmers University of Technology.
Foglia A., Ibsen L.B., Andersen L.V., Roesen H.R. (2012). Physical modeling of
bucket foundation under long term cyclic lateral loading. Proc. 22nd Int. Offshore
and Polar Engng Conf., 667-673.
Freeman, T.J. and Burdett, J.R.F. (1986). "Deep ocean model penetrator
experiments." Final report to Commission of European Communities. Report No.
EUR 10502.
Gan, C.T., Leung, C.F., Cassidy, M.J., Gaudin, C., Chow, Y.K. (2012). "Effect of
time on spudcan-footprint interaction in clay." Gotechnique 62(5): 401-413.
Gaudin, C., O'Loughlin C.D., Randolph, M.F., Lowmass, A.C. (2006). "Influence of
the installation process on the performance of suction embedded plate anchors"
Gotechnique 56(6): 381-391.
Gaudin C., Randolph M.F., Clukey E., Dimmock P. (2012). "Centrifuge modelling of
a hybrid subsea foundation for subsea equipment." Proc 7th Int. Conf. of Offshore
Site Investigations and Geotechnics, London, England, 411-420.
Gaudin C., Simkin M., White D.J., OLoughlin, C.D. (2010). Experimental
investigation into the influence of a keying flap on the keying behaviour of plate
anchors. Proc. 20th Int. Offshore and Polar Engng Conf., Beijing, China: 533-540.
Gaudin C., Tham K.H., Ouahsine S. (2009). "Keying of plate anchors in NC clay
under inclined loading." Int. Journal of Offshore and Polar Engng. 19(2):1-8.
Gerdes, G., Tiedemann, A. and Zeelenberg, S. (2006). Case Study: European
Offshore Wind Farms: A Survey to analyse experiences and lessons learnt by
developers of offshore wind farms, Final Report, Deutsche WindGuard GmbH.
Hodder, M.S., White, D.J., Cassidy, M.J. (2009). "Effect of remoulding and
reconsolidation on the touchdown stiffness of a steel catenary riser: observations
from centrifuge modelling." Offshore Technology Conference, Houston, USA.
Paper OTC19871.
Hodder, M.S., White, D.J., Cassidy, M.J. (2013). "An effective stress framework for
the variation in penetration resistance due to episodes of remoulding and
reconsolidation." Gotechnique 63(1):30-43.

Page 24

Geo-Congress 2014 Keynote Lectures, GSP 235 ASCE 2014

25

Hossain, M.S. and Randolph, M.F. (2009). "New mechanism-based design approach
for spudcan foundations on stiff-over soft-clay." Offshore Technology Conference,
Houston, USA, OTC19907.
Hossain, M.S., Cassidy, M.J., Daley, D., Hannan, R. (2010). "Experimental
investigation of perforation drilling in stiff-over-soft clay." Applied Ocean
Research, 32(1):113-123.
Houlsby, G.T. and Martin, C.M. (2003). "Undrained bearing capacity factors for
conical footings on clay." Gotechnique, 53(5): 513-520.
Hu, P., Stanier, S., Cassidy, M.J., Wang, D. (2013). "Predicting the peak punchthrough penetration resistance of a spudcan penetrating sand overlying clay."
Journal of Geotechnical and Geoenvironmental Engineering, ASCE, In Press.
IEA (2005). "Offshore Wind Experiences," International Energy Agency, France.
IEA (2012a). "2012 Key world energy statistics." International Energy Agency. Paris,
France. www.iea.org
IEA (2012b). "Medium-term renewable energy market report 2012 Market trends
and projections to 2017." International Energy Agency. France.www.iea.org.
Kong, V.W., Cassidy, M.J., Gaudin, C. (2013). "Experimental study of the effect of
geometry on the reinstallation of a jack-up next to a footprint." Canadian
Geotechnical Journal 50(5):557-573.
Lee, K.K., Cassidy, M.J., Randolph, M.F. (2013a). "Bearing capacity on sand
overlying clay soils: Experimental and finite element investigation of potential
punch-through failure." Gotechnique, In press, (doi): 10.1680/geot.12.P.175.
Lee, K.K., Randolph, M.F., Cassidy, M.J. (2013b). "Bearing capacity on sand
overlying clay soils: A simplified conceptual model" Gotechnique, (doi):
10.1680/geot.12.P.176.
Lehane, B.M., OLoughlin, C.D., Gaudin, C., Randolph, M.F. (2009). "Rate effects
on penetrometer resistance in kaolin." Gotechnique, 59(1):41-52.
Leung, C.F., Gan, C.T. and Chow, Y.K. (2007). "Shear strength changes within jackup spudcan footprint." Proc. 7th Int. Offshore and Polar Engng Conf. (ISOPE),
Lisbon, Portugal, 1504-1509.
Lieng, J.T., Kavli, A., Hove, F., Tjelta, T.I. (2000). "Deep Penetrating Anchor:
further development, optimization and capacity clarification." Proc. Int. Offshore
and Polar Engng Conf., Seattle, USA, 410-416.
Lowmass, A.C. (2006). "Installation and keying of follower embedded plate
anchors." Masters of Engineering Thesis, UWA
Matsui, T., Ohara, H., Ito, T. (1980). "Cyclic stress-strain history and shear
characteristics of clay." Journal of the Geotechnical Engineering Division ASCE,
106(10): 1101-1120.
Musial, W., Butterfield, S., and Ram, B. (2006). "Energy from offshore wind," Proc.
2006 Offshore Technology Conference, OTC18355: 11.
Musial, W. and Ram, B. (2010). Large-Scale Offshore Wind Power in the United
States. Assessment of opportunities and barriers. NREL/TP-500-40745,
Retrieved from http://www.nrel.gov/wind/pdfs/40745.pdf
OLoughlin, C.D., Lowmass, A., Gaudin, C., Randolph, M.F. (2006). "Physical
modelling to assess keying characteristics of plate anchors." Proc. 6th Int. Conf.
on Physical Modelling in Geotechnics, Hong Kong, 659-665.

Page 25

Geo-Congress 2014 Keynote Lectures, GSP 235 ASCE 2014

26

O'Loughlin, C.D., Randolph, M.F., Richardson, M.D. (2004). "Experimental and


theoretical studies of Deep Penetrating Anchors." Offshore Technology
Conference, Houston, USA, OTC 16841.
OLoughlin, C.D., Richardson, M.D., Randolph, M.F. (2009). "Centrifuge tests on
dynamically installed anchors." Proc. 28th Int. Conf. on Ocean, Offshore and
Arctic Engng, Honolulu, Hawaii, OMAE2009-80238.
O'Loughlin, C.D., Richardson, M.D., Randolph, M.F., Gaudin, C. (2013).
"Penetration of dynamically installed anchors in clay." Gotechnique, 63, doi:
10.1680/geot.11.P.137.
ONeill, M.P., Bransby, M.F., Randolph, M.F. (2003). "Drag anchor fluke-soil
interaction in clay." Canadian Geotechnical Journal, 40(1):78-94.
Osborne, J.J., Houlsby, G.T., Teh, K.L., Bienen, B., Cassidy, M.J., Randolph, M.F.,
Leung, C.F. (2009). "Improved guidelines for the prediction of geotechnical
performance of spudcan foundations during installation and removal of jack-up
units." Offshore Technology Conference, Houston, OTC-20291.
Osborne, J.J., Pelley, D., Nelson, C., Hunt, R. (2006). Unpredicted jack-up
foundation performance. Proc. 1st Jack-up Asia Conference. Singapore.
Osborne, J.J., Teh, K.L., Houlsby, G.T., Cassidy, M.J., Bienen, B., Leung, C.F.
(2010). "Improved guidelines for the prediction of geotechnical performance of
spudcan foundations during installation and removal of jack-up units." RPS
Energy Report Number EOG0574-Rev1. Final Guidelines of the InSafe Joint
Industry Project, 124p. http://insafe.woking.rpsplc.co.uk/Default.asp
Randolph, M.F., Gaudin, C., Cassidy, M.J., Wang, D., Tian, Y. (2010).
"Qualification of SEPLA for permanent mooring: Interpretation of centrifuge,
finite element and analytical plasticity modelling." Geo: 10519, COFS, UWA.
Randolph, M.F., Gaudin, C., Gourvenec, S., White, D.J., Boylan, N., Cassidy, M.J.
(2011). "Recent advances in offshore geotechnics for deep water oil and gas
developments." Ocean Engineering, 38:818-834.
Randolph, M.F., Jamiolkowski, M.B., Zdravkovic, L. (2004). "Load carrying
capacity of foundations." Proc. Skempton Memorial Conf. London, 1:207-240.
RBA (2011). "The Global Market for Liquefied Natural Gas." Reserve Bank of
Australia, Bulletin, September Quarter 2011, Jacobs, D. 17-27.
Richardson, M., OLoughlin, C.D., Randolph, M.F. (2006). "Drum centrifuge
modelling of Dynamically Penetrating Anchors." Proc. 6th Int. Conf. on Physical
Modelling in Geotechnics, 1:673 678.
Richardson, M.D., OLoughlin, 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):487-496.
Sclavounos, P. D., Tracy, C. and Lee, S. (2008). "Floating Offshore Wind Turbines:
Responses in a Seastate, Pareto Optimal Designs and Economic Assessment."
Offshore Mechanics and Arctic Engineering Conference, Lisbon, Portugal.
SNAME (2008) "Site specific assessment of mobile jack-up units. Research Bulletin
5-5A." New Jersey: Society of Naval Architects and Marine Engineers.
Song, Z., Hu, Y., OLoughlin, C., Randolph, M.F. (2009). "Loss in anchor
embedment during plate anchor keying in clay." Journal of Geotechnical and
Geoenvironmental Engineering, ASCE 135(10):1475-1485.

Page 26

Geo-Congress 2014 Keynote Lectures, GSP 235 ASCE 2014

27

Stuyts, B., Vissers, V., Cathie, D.N., Jaeck. C., and Dorfeldt, S. (2011). "Optimizing
site investigations and pile design for wind farms using geostatistical methods: A
case study," Proc. 2nd Int. Symp. on Frontiers in Offshore Geotechnics, Perth,
Australia, 629-633.
Teh, K.L., Cassidy, M.J., Leung, C.F., Chow, Y.K., Randolph, M.F., Quah, C.K.
(2008). "Revealing the bearing failure mechanisms of a penetrating spudcan
through sand overlaying clay." Gotechnique 58(10):793-804.
Teh, K.L., Leung, C.F., Chow, Y.K., Handidjaja, P. (2009). "Prediction of punchthrough for spudcan penetration in sand overlying clay." Offshore Technology
Conference, Houston, USA.
Tian, Y., Gaudin, C., Cassidy, M.J., Randolph, M.F. (2013). Considerations on the
design of keying flap of plate anchors. Journal of Geotechnical and
Geoenvironmental Engineering, ASCE 139(7):1156-1164.
True, D.G. (1976). "Undrained vertical penetration into ocean bottom soils." PhD
thesis, University of California, Berkeley.
Wang, D., Hu, Y., Randolph, M.F. (2010). "Three-dimensional large deformation
finite-element analysis of plate anchors in uniform clay." Journal of Geotechnical
and Geoenvironmental Engineering, ASCE, 136(2):355-365.
Wang, D., Hu, Y., Randolph, M.F. (2011). "Keying of rectangular plate anchors in
normally consolidated clays." Journal of Geotechnical and Geoenvironmental
Engineering, ASCE, 137(12):1244-1253.
Wei, Q., Cassidy, M.J., Tian, Y., Gaudin, C. (2013). "Incorporating shank resistance
into prediction of the keying behaviour of SEPLAs." Journal of Geotechnical and
Geoenvironmental Engineering, ASCE, submitted 12 March 2013.
Wong, P.C., Gaudin, C., Randolph, M.F., Cassidy, M.J., Tian, Y. (2012).
"Performance of suction embedded plate anchors in permanent mooring
applications." Proc. 22nd Int. Offshore and Polar Engng Conf., Rhodes, Greece,
640-645.
Yang, M., Aubeny, C.P. and Murff, J.D. (2012). "Behaviour of suction embedded
plate anchors during the keying process." Journal of Geotechnical and
Geoenvironmental Engineering, ASCE, 138(2):174183.
Zimmerman, E.H., Smith, M.W. and Shelton, J.T. (2009). "Efficient gravity installed
anchor for deepwater mooring." Offshore Technology Conf., Houston, USA.

Page 27

You might also like