You are on page 1of 30

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/233699946

Foundation Design: A Comparison of Oil and Gas Platforms with Offshore Wind
Turbines

Article  in  Marine Technology Society Journal · January 2010


DOI: 10.4031/MTSJ.44.1.5

CITATIONS READS
17 7,399

2 authors:

James Schneider Marc Senders


Consulting Engineer Woodside Energy
52 PUBLICATIONS   1,294 CITATIONS    10 PUBLICATIONS   120 CITATIONS   

SEE PROFILE SEE PROFILE

All content following this page was uploaded by James Schneider on 07 November 2015.

The user has requested enhancement of the downloaded file.


Foundation design – a comparison of oil and gas platforms with
offshore wind turbines

James A. Schneider
The University of Wisconsin-Madison
1415 Engineering Drive
Madison, Wisconsin 53706 USA
jamess@cae.wisc.edu

Marc Senders
Centre for Offshore Foundations Systems
The University of Western Australia
Crawley, WA 6009 Australia
m.senders@bigpond.com

Abstract:
The offshore oil and gas industry (O&G) has over 70 years of experience developing innovative
structures and foundation concepts for engineering in the marine environment. The evolution of
these structures has strongly been influenced by water depth as well as soil conditions in the area
of initial developments. As the offshore wind industry expands from the glacial soil deposits of
the North and Baltic Seas, experience from the O&G industry can be used to aid with a smooth
transition to new areas. This paper presents an introduction to issues that influence how design
and construction experience from the O&G industry can be used to aid foundation design for
offshore wind energy converters. A history of the evolution of foundation and substructure
concepts in the Gulf of Mexico and North Sea is presented, followed by a discussion of soil
behavior and the influence of regional geology on these developments. Mechanisms that
influence the resistance of shallow and deep foundations for fixed and floating offshore
structures is outlined so that areas of empiricism within offshore design codes can be identified
and properly modified for application to offshore wind turbine foundations. It is concluded that
there are distinct differences between offshore O&G and offshore wind turbine foundations, and
application of continued research into foundation behavior is necessary for rational, reliable, and
cost effective design.

1 Introduction used to drill 11 wells in the Creole field,


The offshore oil and gas (O&G) eventually producing nearly four million
industry is rooted 1.5km off the Louisiana barrels of oil (Wasson 1948, Leffler et al.
coast in the Gulf of Mexico (GOM). Figure 2003) or an equivalent 6800 GWh of
1 illustrates the 2700m2 platform in 4m of electricity. Since that time, approximately
water supported on timber pilings that was 4000 platforms have been constructed in the

Schneider, J.A., and Senders, M. 2010. Foundation design – a comparison of oil and gas
platforms with offshore wind turbines, Journal of the Marine Technology Society, 44(1), 32-51.
Figure 1a. Location map for the Creole Field
(Wasson 1948) Figure 1b. Photograph (looking south) of Creole
platform in the Gulf of Mexico (Wasson 1948)

piled foundations used for the platform in


the Creole field, the 11 Vindeby turbines are
GOM, and more than 2500 additional supported on shallow gravity base
platforms are located around the world. foundations. The use of gravity base shallow
Fixed platforms have been constructed in foundations is an extension of offshore oil
water depths up to 530m, and floating and gas experience in the North Sea. Like
platforms to 2400m (Wilhoit & Supan the O&G industry, successful early projects
2009). The experience gained in the offshore tend to act as prototypes resulting in future
O&G industry has resulted in development use of similar foundation solutions for
of innovative structures and foundation similar projects in similar geologies. The 10
concepts to handle complex loading turbines at Tunø Knob, also in shallow
conditions and soil behavior that is waters of the Danish Baltic Sea, used similar
associated with engineering in the marine gravity base foundations to those used at
environment. Vindeby. While the offshore wind industry
Like the offshore O&G industry, the has an 18 year track-record of design and
offshore wind industry started development construction of various substructure and
relatively close to shore in relatively shallow foundation types, there is still much that can
water. The first offshore commercial wind be learned from the experiences of the
farm is located 2.5 km off the cost of offshore O&G industry, particularly for
Denmark, near Vindeby, in 3 to 5m water deeper waters. Although, due to differences
depths of the Baltic Sea. Vindeby began in soil and loading conditions, among other
operation in 1991, nearly 55 years after the factors, experience from the O&G industry
Creole field, and is producing approximately must be assessed within a rational
11.2 GWh of electricity per year. Unlike the engineering framework for application to

Schneider, J.A., and Senders, M. 2010. Foundation design – a comparison of oil and gas
platforms with offshore wind turbines, Journal of the Marine Technology Society, 44(1), 32-51.
these new situations. Empiricism within occurs due to horizontal loading from a
offshore O&G design method formulations selected (i.e., 1 year, 10 year, 100 year,
must be identified, and modified for the 10000 year, etc) design storm. For a rigid
differing conditions of offshore wind turbine structure, such as a template jacket in Figure
development. 2a-i, moment induced by horizontal loads is
This paper first presents an overview resisted by compression in the leeward leg
of substructure types and loading conditions, and tension in the windward leg of the
focusing on evolution of practice in the platform. The horizontal loads must also be
GOM and North Sea. Development of resisted by lateral loading of foundation
substructures and foundation types in the elements. Floating offshore platforms with
GOM and North Sea has been quite vertical mooring lines have foundations that
different, which resulted from the regional are permanently in tension due to the
geology and associated soil behavior. buoyancy of the structure, and horizontal
Aspects of soil behavior that control storm loads lead to increased tension that
offshore foundation design are highlighted, must be resisted by the foundation.
and the methods for calculating the The magnitude of vertical and
resistance of (i) shallow foundations; (ii) horizontal forces for fixed offshore
skirted foundations / suction caisson; (iii) structures varies significantly with platform
monopiles; (iv) piled jackets / tripods; and size, and is also influenced to a large degree
(v) anchors, are discussed. This paper by water depth. Vertical force due to
presents an introduction to issues that platform weight increases from about
influence how design and construction 100MN in relatively shallow waters to about
experience from the oil and gas industry can 500 MN for water depths of about 60m (e.g.,
be used to aid in offshore wind turbine Poulos 1988). When water depth exceeds
design, and the reader is referred to 120m, platform weight may exceed
additional publications on offshore 2000MN. Horizontal storm loads on large
foundation design for a more thorough offshore O&G jacket structures in water
discussion of methods (e.g., George & depths of about 100m have been recorded at
Wood 1976, Poulos 1988, Randolph et al. up to 235 MN (Marshall & Bea 1976),
2005, Gerwick 2007, Randolph & although a majority of platforms are smaller
Gourvenec 2010). and in shallower water such that horizontal
design loads are typically on the order of 50
2 Loading conditions and substructure to 100MN. On the contrary, horizontal wave
types loads on offshore wind turbines are a
Loads transferred from a structure to the fraction of those loads, and on the order of 1
soil and soil engineering properties govern to 3MN (e.g., Senders 2005, Houlsby et al
the size and type of foundation that is 2005a, Manwell et al. 2007). This large
required to support a structure. A difference results from the much smaller
comparison of foundation loading for fixed size of offshore wind structures as compared
and floating offshore O&G structures, as to offshore O&G jacket structures. Wind
well as offshore wind turbines is illustrated loads for a 3MW wind turbine with a hub
in Figure 2a. The design loading condition height of 75 to 100m under normal
for an offshore O&G platform typically operational conditions are on the order on

Schneider, J.A., and Senders, M. 2010. Foundation design – a comparison of oil and gas
platforms with offshore wind turbines, Journal of the Marine Technology Society, 44(1), 32-51.
Figure 2a.
a. Comparison of loading conditions for fixed, floating, and monopile offshore structures

Figure 2b.
b. Influence of number of cycles on allowable load to prevent fatigue failure for a
variety of structures (after Spera 1994)

Schneider, J.A., and Senders,, M. 2010. Foundation design – a comparison of oil and gas
platforms with offshore wind turbines, Journal of the Marine Technology Society, 44(1), 32-51.
32
0.1MN, although gust loads of up to 1MN 2007). Fatigue of structural elements is a
could occur (e.g., Senders 2005, Houlsby et very important consideration for both
al 2005a). For the case of the offshore wind offshore platforms and offshore wind
turbine in Figure 2a under wave and typical turbines, while the soil and foundation
wind loading conditions, foundation loads response to a large number of relatively low
due to wind and wave forces are similar for amplitude cycles is still being researched
water depths of 10 to 15m (for a hub height (e.g., LeBlanc et al. 2010, Senders 2008).
of 75 to 100m), and wave loads tend to Soil dynamics for foundation vibrations
dominate for water depths greater than 20 to (e.g., Richart 1960) and earthquake loading
25m. In the US and Canadian Great Lakes, (e.g., Seed & Idriss 1970), have a long
the relatively small wave environment and history of application, although the strict
greater flexural strength of freshwater ice (as vibration tolerances of offshore wind
compared to saltwater ice) may result in ice turbines as compared to offshore O&G
loads that control design. As water depth platforms requires careful consideration.
increases, loads on the foundation of a fixed The primary role of dynamics for the design
O&G or offshore wind structure will of offshore O&G platforms relates to the
increase due to the increased moment arm of predominant period of storm loading
the structure height above the seabed. This conditions in relation to that of a platform.
design difficulty can be overcome by the use Storm loading conditions typically have
of floating offshore structures where greater natural periods between 5 and 30 seconds,
water depth does not directly lead to and therefore structures need to be designed
increased storm loading, and, therefore, such that they are either stiff with a natural
essentially no limits on structure location period less than 5 seconds or compliant with
due to water depth. a natural period greater than 30 seconds (e.g.
It is noted that wind, waves, current, Ellers 1982). For the case of offshore wind
and ice loads in Figure 2a are cyclic in turbines, the natural period of the system
nature. Cyclic resistance is typically must differ from that of the environmental
considered to be a function of the static loading as well as both the natural period of
resistance and reduces with the number of the rotor (1P) and periods related to higher
cycles (e.g., Basquin 1910, Spera 1994). order harmonics (e.g., Kühn 1997, Liingaard
Figure 2b compares the number of cycles 2006). For foundations on softer sediments,
from an offshore design storm to those for dynamic response will deviate from the
various cases, including wind turbines. fixed condition that can be calculated using
Large storm events, with 100 to 1000 the Rayleigh method (e.g., Liingaard 2006),
relatively high amplitude cycles is typically and the influence of soil stiffness
the focus for foundation design of offshore nonlinearity on natural period of the
O&G structures, whereas the small structure becomes an important
amplitude cycles, with over 1 billion cycles, consideration (e.g., Houlsby et al. 2005a,
typically govern the wind turbines design. 2005b).
When wave loading conditions are relatively When looking at the evolution of
weak, the potential for wind induced fatigue substructure and foundation design in light
has controlled design at some European of water depth and soil conditions, the first
offshore wind turbine locations (Saigal et al. 25 years of offshore development occurred

Schneider, J.A., and Senders, M. 2010. Foundation design – a comparison of oil and gas
platforms with offshore wind turbines, Journal of the Marine Technology Society, 44(1), 32-51.
in relatively shallow waters and was evolved with time (Figure 3b, after Wilhoit
dominated by the use of template jacket & Supan 2009):
structures (Figures 2a-i, 3a). These • Tension Leg Platforms (typically 150 to
structures could be constructed on land, 1450m water depths)
floated to location, and installed in relatively • Semi-submersibles (typically 100 to
short periods of time to minimize the 2400m water depth)
potential for weather affecting construction • Spar (typically 600 to 2400m water
schedule and costs. By 1970, jacket depth); and
structures had been constructed in water • FPSO (Floating Production Storage and
depths of up to 120m, and developments in Offloading, up to 2500m water depth).
substructure technology were being explored FPSOs are ship shaped structures which
to extend these water depths to greater than dominate the number of floating structures
200m (McClelland 1974). for the O&G industry, but are not included
While it is possible to build increasingly in Figure 3b since they have little
larger jacket structures to accommodate application to the offshore wind turbine
greater water depths, increased substructure industry. Except for Spars, floating
weight leads to higher material costs as well structures can be used in a large range of
as increased vertical loads that require larger water depths. Spars for the O&G industry
and more expensive foundation. To reduce typically have a buoyancy can length on the
these additional costs a variety of order of 150 to 225m to resists horizontal
substructure types have been developed to loading through added ballast, therefore,
support offshore exploration and production preventing their use in shallower waters.
facilities. Structures that are rigidly attached Since the offshore wind industry
to the seabed (fixed structures) are shown in started development some 55 years after the
Figure 3a and include (after Poulos 1988; first offshore O&G platform, it is logical
Wilhoit & Supan 2009): that previously developed substructure and
• Steel Jacket (or Template) Structures foundation technology from the O&G
(typically less than 450m water depth) industry was used for initial wind farm
• Concrete Gravity Base Structures (GBS) development. Figure 4 illustrates foundation
(typically less than 350m water depth) concepts for the offshore wind industry as a
• Guyed Towers (typically greater than function of water depth. To date, most
300m water depth); and offshore wind turbines have been
• Compliant Towers (typically 300 to constructed in shallow waters of less than
600m water depth). 30m depth. Monopile foundations support
Developments in the GOM focused on Fixed over 70% of the installed capacity, followed
Structures until the early 1990s. by Gravity Base Structures (GBS). The
To minimize additional material costs Beatrice pair of 5MW offshore wind
related to the increasing size of fixed turbines, installed in 2007 in 45m of water,
structures in deep water, the use of floating were the first jacket structures constructed
structures began for the O&G industry in the for offshore wind turbines. The 5MW
early 1970s. Like fixed structures, there are Hywind Spar, installed in 2009 in 220m
a variety of floating structures that have water depth, was the first floating offshore
wind turbine with commercial scale

Schneider, J.A., and Senders, M. 2010. Foundation design – a comparison of oil and gas
platforms with offshore wind turbines, Journal of the Marine Technology Society, 44(1), 32-51.
Figure 3a. Typical fixed offshore O&G Figure 3b. Typical floating offshore O&G
structures with typical ranges of water depth structures with typical ranges of water depth

Figure 4. Substructure types for offshore wind turbine foundations (after Musial et al. 2006)

Schneider, J.A., and Senders,, M. 2010. Foundation design – a comparison of oil and gas
platforms with offshore wind turbines, Journal of the Marine Technology Society, 44(1), 32-51.
32
capacity. TLPs and semi submersibles are Seas. Figure 6 expands on those points,
still in the research phase for floating giving a brief description of each event.
offshore wind turbines. Compliant and Since offshore O&G development in the
guyed towers have not received much GOM started about 20 years prior to
interest from the offshore wind turbine offshore O&G development in the North
industry since the large lateral movements Sea, and since the GOM has much deeper
that can be accommodated by offshore oil maximum water depths, key points in GOM
platforms in deep waters have been development tend to track record water
considered to be detrimental to wind turbine depths. A notable foundation option in the
performance. Wind turbine manufacturers GOM, known as a ‘caisson’ or tapered
are currently developing systems which can monopile, has been used to support
accommodate larger rotations, and it is individual wells since the early 1960’s.
noted that the Hywind Spar substructure These are broadly similar to monopiles used
installed offshore of Norway in 2009 has for offshore wind turbines today except that
been designed to rotate 2-4 degrees under diameters are on the order of 0.7m (30”) to
normal operating conditions and up to 11 1.2m (48”) as compared to 3m (120”) to 5m
degrees for extreme storm loading (198”) diameter monopiles common for
conditions (Rølland 2009). offshore wind turbines. While few
Figure 5 shows the increase in foundation failures have been reported for
maximum water depth of offshore O&G O&G structures in the GOM, it is interesting
construction with time (line), as well as the to note that presented failure case histories
year and water depth (points) for key O&G typically include ‘caisson’ tapered monopole
developments in the GOM, O&G foundations (e.g., McClelland & Cox 1976,
developments in the North Sea, and offshore PMB 1996).
wind developments in the North and Baltic

Figure 5. Maximum water depth for offshore Oil & Gas (O&G) and offshore wind turbine
construction with key dates identified (see Figure 6 for description of key dates)

Schneider, J.A., and Senders, M. 2010. Foundation design – a comparison of oil and gas
platforms with offshore wind turbines, Journal of the Marine Technology Society, 44(1), 32-51.
Figure 6. Timelines for evolution of substructure and foundation technology for (a) Oil & Gas
development in the Gulf of Mexico; (b) Oil & Gas development in the North Sea; (c) Offshore
wind development in the North and Baltic Seas

Schneider, J.A., and Senders, M. 2010. Foundation design – a comparison of oil and gas
platforms with offshore wind turbines, Journal of the Marine Technology Society, 44(1), 32-51.
Differences in substructure and construction in rock underlying a significant
foundation developments in the GOM and thickness of softer soils. As with offshore
North Sea were significantly affected by O&G development in the North Sea,
predominant soil conditions, as discussed in development in offshore wind turbine
more detail in the next section. Despite practice was largely influenced by regional
having shallower waters, developments in geology and soil conditions.
the North Sea included the first floating
production platform and the first tension leg 3 Soil behavior and regional geology
platform. GOM foundations are dominated 3.1 Soil Behavior
by driven piles, with essentially all TLPs to The ease of visual identification and
date anchored by driven piles, and simple laboratory characterization tests has
approximately 55% of Spars to date led to classification of soils being typically
anchored by driven piles (the remaining based on grain size (i.e., gravel, sand, silt,
45% tend to use suction caissons). A range clay), although engineering behavior is more
of foundation and substructure types were strongly controlled by how fast water can
developed for the North Sea, including the flow through soil pores in relation to how
previously mentioned Gravity Base fast the soil is loaded. If water cannot flow
Structure (GBS), skirted shallow fast enough to escape from the pores, excess
foundations and suction caissons, as well as pore water pressure (∆u) will change during
concrete substructures and floating loading. This phenomenon can broadly be
platforms. The development of concrete identified as “water flow characteristics,”
substructures was brought about by the and quantified using the hydraulic
ability to construct and float out these tall conductivity (k), or more appropriately the
structures in the deep fjords of Norway. vertical (or horizontal) coefficient of
Evolution of offshore wind turbine consolidation, cv = k v (1 + e )σ ' v 0 (γ w λ ) ,
substructures was broadly similar to that of where the subscript v indicates the vertical
the offshore O&G industry in the North Sea.
component, e is the void ratio, σ'v0 is the
Initial construction focused on relatively
vertical effective stress, γw is the unit weight
shallow water depths, and development of
of water, and λ is the slope of the normal
different types of foundations that led to cost
compression line, (Terzaghi & Peck 1948).
effective construction. While monopile
Since water is essentially incompressible (in
foundations dominate the construction
relation to a soil matrix), increases in total
practice for offshore wind turbines, there are
bulk stresses will initially result in increased
actually a range of monopile solutions.
excess pore pressures such that effective
These include (i) driven monopiles; (ii)
stresses don’t change, therefore resisting
bored monopiles; as well as (iii) driven-
volume change. Water flows from areas of
drilled monopiles. The various construction
high excess pore pressure, leading to an
techniques were utilized to account for
equalization of pore pressures, local
subsurface conditions, with driven
increases in effective stress, and settlement
monopiles working best in relatively thick
of the soil. Water flow characteristics are
soil deposits; bored monopiles applicable to
near surface rock or very hard soil deposits, strongly related to the effective grain size
(d10, the grain diameter corresponding to
and the driven-drilled option required for
10% finer by weight) and the uniformity of

10

Schneider, J.A., and Senders, M. 2010. Foundation design – a comparison of oil and gas
platforms with offshore wind turbines, Journal of the Marine Technology Society, 44(1), 32-51.
the particle sizes, therefore resulting in sheared. Separating the state of sands as
engineering behavior also being strongly ‘loose’ or ‘dense’ is in quotes since it is
controlled by grain size and adequate use of actually density and effective stress that
existing grain size based classification control volume change of sands (e.g., Bolton
systems. 1986). Simply speaking ‘dense’ sands will
In addition to water flow characteristics, tend to dilate while ‘loose’ sands will tend to
engineering behavior of soil can be contract when sheared, and dilation will lead
quantified using three primary properties to higher peak friction angles (and higher
(e.g., Atkinson 2007): strength) for ‘dense’ materials.
• Shear strength (φ' and ψ, or su); Shear strength in uncemented soils
• Shear stiffness (G); and may be characterized for drained conditions,
• Compressibility (λ and κ, or mv, or 1/K). where water can escape pores faster than the
Effective stress (σ' = σ – u, were σ is the loading rate and no excess pore pressure
stress and u is the pore water pressure) (∆u) is generated, using the friction (φ') and
controls shear strength, shear stiffness, and dilation (tanψ=−dεv/dγs) angles within the
compressibility. Soil ‘state’ will also affect Coulomb failure criterion (τf=σ'tanφ', where
strength, stiffness, and compressibility τf is the shear stress at failure). During
through void ratio, ratio of horizontal to drained loading shear strain (γs) will cause
vertical effective stress, and dilation angle. volumetric strain (εv), and these changes in
The dilation angle controls changes in volume will affect foundation resistance. For
volume during drained loading as well as the soils with slow water flow characteristics,
tendency for positive or negative shear such as clays, or for rapid loading, such as
induced pore pressures during undrained storm waves, analysis can be considered as
loading. undrained, and excess water pressure (∆u)
Table 1 summarizes relative values will increase with increases in mean stress
of soil strength, stiffness, compressibility, (assuming that water is incompressible in
and water flow characteristics for typical relation to soil). Shearing of the soil may
soil and rock types. Clays and sands are cause increases or decreases in excess water
separated by ‘state’, or normally pressure depending upon soil state. Under
consolidated and over consolidated clays as undrained conditions there is no volume
compared to ‘loose’ and ‘dense’ sands. A change (ψ=0) and strength can be
normally consolidated (NC) clay is a characterized using the undrained shear
‘recently’ (within a geologic time scale) strength, su, within the Tresca failure
deposited soil that tends to contract or criterion (τf=su). After undrained loading,
generate positive pore pressure when excess pore pressures will tend to reduce
sheared. An over consolidated (OC) clay has with time, and after drained loading (to a
been previously compressed and unloaded, significant fraction of the failure stress)
such as the current condition of glacially volume change may result in a soil with a
affected deposits, and has much higher looser structure. For saturated subsea soils it
strength and stiffness than a normally must be recognized that effective stress and
consolidated clay. An OC clay will tend to void ratio change with time after foundation
dilate or generate negative pore pressures construction and loading, and therefore
(and increase the effective stress) when

11

Schneider, J.A., and Senders, M. 2010. Foundation design – a comparison of oil and gas
platforms with offshore wind turbines, Journal of the Marine Technology Society, 44(1), 32-51.
Table 1. Relative engineering characteristics of typical soils and rock

Material Type Water Flow Strength Stiffness Compressibility


Characteristics
Normally consolidated CLAY Very low Low Low High
Over consolidated CLAY Low to very low Med. to high Med. to high Low to med.
‘Loose’ SAND High Med. to high Med. Low
‘Dense’ SAND High High High Low
Calcareous SAND High High Med. to high Med.
SILTS and mixtures Low to med. Low to high Low to high Low to high
ROCK Very low to high Very high Very high Very low

Figure 7a. Cyclic degradation of two Figure 7b. Nonlinear stiffness of suction caisson
calcareous sands foundation (after Houlsby et al. 2005b)
(data from Sharma & Fahey 2003)

strength, stiffness, and compressibility will ‘endurance limit’, with the number of cycles
also change. raised to a power. The rate of degradation in
An extreme case of changes in Figure 7a is not unique for both of the
effective stress occurs during cyclic loading. calcareous sands shown, and this rate is
Under sustained cyclic loading, the effective likely to be soil type and state dependent.
stress will tend to decrease and therefore the When extrapolating to a number of cycles
soil strength will decrease, as illustrated for representative of a 20 year wind turbine
isotropic loading of two different calcareous design life, fatigue strength for these
sands in Figure 7a. Like fatigue of steel, the calcareous sands appears to be less than 1/10
soils data broadly follow ‘the exponential of the static strength. This extrapolation is
law of endurance tests’ (Basquin 1910), highly uncertain, particularly when
where strength decreases, up to an considering the potential for an ‘endurance

12

Schneider, J.A., and Senders, M. 2010. Foundation design – a comparison of oil and gas
platforms with offshore wind turbines, Journal of the Marine Technology Society, 44(1), 32-51.
limit,’ and a more rational theory is required depth, but also by soil type. Figures 8a and
to assess soil behavior. Cyclic degradation 8b illustrate typical cross sections through
can be quantified using four different types the GOM and North Sea, respectively. The
of mechanical behavior that depend on GOM and North Sea have distinctly
anisotropy of confining stress and different geological origins, where the GOM
magnitude of cyclic shear stresses relative to is predominantly young, recently deposited
the failure stress; (i) elastic; (ii) shakedown; fine grained sediments, and the soils of the
(iii) alternating plasticity; or (iv) incremental North Sea have been compressed and
collapse (e.g., Collins & Boulbibane 2000). reformed as a result of the advance and then
Elastic behavior mentioned above would retreat of glaciers up to about 10,000 years
imply the presence of an ‘endurance limit’ before present. The cross sections show
for soils, although, no endurance limit has large laterally continuous layers of
been observed in lab tests for stress ratios predominantly clayey soils, both having
(Q/Qult) as low as 0.05 and a number of relatively slow water flow characteristics
cycles up to 3x105 (Clukey et al. 1995). that result in ‘undrained loading’ during
Research into fatigue resistance of soils and foundation construction or rapid loading
foundation appears warranted, although, this from storm events. Differences arise in that
has not been a significant design issue in the the GOM clays are ‘normally consolidated’
O&G industry. (NC) and the North Sea clays are ‘over
It should be noted that soils behave consolidated’ (OC). As laid out in Table 1,
as a nonlinear elastic material from very the state of the clay soils will lead to vastly
small shear strain levels. Therefore, even different engineering behavior, with NC
under small cyclic shear loads the clays being soft and highly compressible,
accumulation of plastic displacements may and the OC clays having a higher strength
lead to a progressive failure of a foundation. and relatively low compressibility.
Figure 7b illustrates the nonlinear shear Additionally, the variability of the North Sea
modulus measured during rotational loading glacial geology is much more significant
of a suction caisson foundation (Houlsby et than that of the GOM. While the GOM
al. 2005b). Modulus reduction is soil type contains a few sand ‘fingers,’ North Sea
dependent and will affect not only the static geology includes a number of near surface
and cyclic response, but also the dynamic channels of sands and soft clays, as well as
response of wind turbine foundation sand ‘fingers’ buried below precompressed
systems. When calibrating springs and stiff to very stiff clays.
dashpots for simplified dynamic analysis, The main differences in evolution of
compatibility between shear strain level foundation types and substructure geometry
shear modulus must be checked. It is also resulted from the thick, continuous soft
noted that material damping will increase deposits of the GOM as compared to the
with shear strain (e.g., Seed & Idriss 1970). much stronger clays and larger amounts of
near surface sands evident in the North Sea.
3.2 Regional Geology For example, the strength of clays in the
The evolution of substructure and GOM tend to increase with depth at a rate of
foundation types for the GOM and North 1 to 2 kPa/m, while stiff overconsoldiated
Sea was influenced not solely by water clays of the North Sea often have a strength

13

Schneider, J.A., and Senders, M. 2010. Foundation design – a comparison of oil and gas
platforms with offshore wind turbines, Journal of the Marine Technology Society, 44(1), 32-51.
Figure 8. Cross sections through northern Gulf of Mexico and central North Sea

of 200kPa or greater starting close to the with deep skirts could be used for
seabed. To reach soils with a strength large construction on soft deposits. Experience
enough to support the large loads from includes (Randolph et al. 2005):
offshore O&G platforms, driven piles were • Gullfaks C – 1989: Soft NC clays and
required in the GOM, while Gravity Base silty clays (su ≈ 30 kPa)
shallow foundations could be utilized in the • Snorre A – 1991: TLP anchored to very
North Sea. soft NC clays
The first GBS in the North Sea was • Troll A – 1995: Soft NC clays with
constructed in 1973 for the Ekofisk tank, strength increasing at about 3kPa/m
and founded on dense sand to 26m underlain from zero strength at the mudline
by hard clay (su ≈ 300kPa) (e.g., Randolph These successes with Gravity Based shallow
et al. 2005). Based on those experiences foundations carried on to the early
Condeep GBSs were built for Beryl A and developments for the offshore wind industry
Brent B in 1975, with Brent B being in the North and Baltic seas, although
constructed predominantly on hard clay expenses related to surface preparation and
(e.g., Claussen 1976). Success with shallow effects of scour have led to monopiles
foundations on stiff deposits led to research dominating the foundations for installed
and analysis such that shallow foundations capacity to date.

14

Schneider, J.A., and Senders, M. 2010. Foundation design – a comparison of oil and gas
platforms with offshore wind turbines, Journal of the Marine Technology Society, 44(1), 32-51.
not employed, highlighting the over reliance
3.3 Calcareous Soils on empiricism that is common in offshore
When transferring knowledge and foundation design. Application of a
experience from the O&G industry to the recommended practice is often not sufficient
offshore wind industry, it must be noted that for successful design in new situations, and
design methods often rely on empiricism the following section aims to outline
rather than solid fundamentals of features of a systematic framework based on
mechanical behavior. The use of offshore O&G experience and also
predominantly empirical methods applicable to offshore wind turbine design.
complicates the engineering process in that
it is often difficult to extrapolate the results 4 Foundation Design
from small scale element tests to full scale Ultimately, successful foundation
design in new situations. A good example of design will produce a structure constructed
the costs that may be associated with over at acceptable costs that resists ultimate
reliance on empiricism comes from collapse as well as excessive deformations
experiences in carbonate deposits of the that would restrict use or result in excessive
North West Shelf of Australia, particularly maintenance costs over the design lifetime
the North Rankin A platform. of a structure. Financial risks arise for
The potential difficulties of design and investigation & design, materials &
construction of driven pile foundations in construction, as well as long term
calcareous sediments was well known, as maintenance. Engineering and design can be
discussed by McClelland (1974), and bored considered the reduction of financial risk
piles were commonly used to support through application of scientific
offshore structures in these deposits. Despite measurements and principals.
a thorough investigation of soil conditions Risks of failure are mitigated in
and evaluation of engineering parameters, engineering by the use of individual load
installation of foundation piles had starkly and resistance factors (or lumped safety
different behavior than anticipated. factors), which ensure that the estimated
Installation resistance when driving with a capacity is greater than the estimated
MHU 1700 hammer was originally demand. Risk is associated with uncertainty,
calculated to be 40 to 100 blows/m after and probabilistically based load and
initial pile self weight penetration (Khorshid resistance factors increase the safety margin
et al. 1988). Typical piles free fell on the as uncertainty increases for a constant
order of 15m or more during driving as the consequence of failure (e.g., Harr 1987). It
result of single hammer blows, and in one should be noted at this point that there are
case a pile free fell nearly 70m due to a large differences in consequence of failure
single blow (Khorshid et al. 1988). for offshore O&G facilities and offshore
While human intervention prevented wind turbines. Namely, these greater risks
this foundation failure from causing a result from the order of magnitude greater
platform failure, foundation retrofit costs expense of each individual O&G structure,
were on the order of AU$340 million. In this as well as the potential loss of life associated
case, an adequate systematic framework with failure of offshore O&G platforms that
with which to extend current practice was is not a factor for offshore wind turbines.

15

Schneider, J.A., and Senders, M. 2010. Foundation design – a comparison of oil and gas
platforms with offshore wind turbines, Journal of the Marine Technology Society, 44(1), 32-51.
This paper will focus on uncertainty (and (v), the foundation diameter (d) and the
bias) in design, although it is recognized that horizontal coefficient of consolidation (ch)
understanding of failure consequences is of (e.g., Finnie & Randolph 1994, Randolph
equal importance for managing risk. 2004, Erbrich 2005):
To minimize costs and produce a v⋅d
‘successful design,’ an engineer would V=
ch
likely aim to reduce the safety margin by
For V less than 0.05 foundation loading is
reducing uncertainty. Due to the empirical
often considered as drained, and for V
nature of many offshore design standards,
greater than 15 foundation loading is
techniques to reduce uncertainty may be
considered essentially undrained (e.g.,
counter intuitive. Successful design is often
Schneider et al. 2008a, Lehane et al. 2009).
a result of conservative bias rather than low
Due to large shallow foundation diameters
uncertainty (e.g., Bea et al. 1999), and
for offshore O&G structures, undrained
therefore a better understanding of design
loading is often the dominant design case,
method formulation is required to minimize
even in sandy soils. The following
bias and allow for better assessment of
discussion of shallow foundation behavior
uncertainty in design (e.g. Schneider 2009).
focuses on the undrained case (ψ=0, τf=su).
The following sections aim to provide
When subjected to a horizontal load,
insight into mechanisms controlling
such as a storm or wind load for an offshore
foundation resistance such that the potential
O&G platform or wind turbine, the leeward
for bias in offshore foundation design is
side of a (hypothetical) weightless
minimized. This will allow for successful
foundation will be in compression and the
extrapolation from foundations for offshore
windward side of the foundation will be in
O&G structures to the case of offshore wind
tension. Shallow foundations provide little
turbines.
resistance to tension (uplift), and the weight
of the foundation must be increased to
4.1 Shallow Foundations
overcome the uplift forces. The strength of
Shallow foundations are those resting
the soil (su) must therefore resist the weight
on the surface of the soil or slightly
of the foundation as well as the additional
embedded such that the embedment depth is
loads induced by the storm loading. Sliding
less than the equivalent diameter of the
resistance, in this case, is usually estimated
foundation. Equivalent diameters of offshore
as the undrained strength for a rough
O&G shallow foundations are typically on
interface.
the order of 50 to 150m, while embedment
Figure 9a illustrates a range of
depths vary from less than 1m to
failure mechanisms for undrained loading of
approximately 35m (e.g., Randolph et al.
shallow foundations (after Bransby &
2005). As previously mentioned, soil
Randolph 1998, Gourvenec & Randolph
behavior is controlled by the rate of water
2003, Randolph et al. 2005, Gourvenec
flow out of the pores in relation to the rate of
2007). Foundation resistance is related to the
loading of the foundation. The rate of
controlling failure mechanism, and the
loading relative to rate of water flow can be
failure mechanism is affected by the relative
assessed using a normalized velocity (V)
magnitudes of vertical (V), horizontal (H)
that is a function of the foundation velocity
and moment (M) loading. Vertical and

16

Schneider, J.A., and Senders, M. 2010. Foundation design – a comparison of oil and gas
platforms with offshore wind turbines, Journal of the Marine Technology Society, 44(1), 32-51.
Figure 9. Failure mechanisms for foundations attached to fixed offshore structures

17

Schneider, J.A., and Senders, M. 2010. Foundation design – a comparison of oil and gas
platforms with offshore wind turbines, Journal of the Marine Technology Society, 44(1), 32-51.
horizontal loads are normalized by the area vertical load (e.g., Gottardi et al. 1999). As
multiplied by the undrained strength, and the rotation of a wind turbine foundation is
moment is normalized by the area multiplied often limited to 1o by manufacturers,
by the diameter and undrained strength. For deformation analysis may be a more
the limiting cases of pure vertical loading, stringent requirement than bearing capacity
pure horizontal loading, and maximum during drained loading. Deformations are
moment (at V/Vult ≈ 0.5), failure surfaces also controlled by the location of loading
(for a circular foundation in Figure 9a-vi) within the three dimensional yield surface,
simplifies to: and it is noted that shear modulus will
• V/Asu ≈ 6; H/Asu = 0; M/ADsu = 0 reduce with increasing shear strain (i.e.,
• V/Asu = 0; H/Asu = 1; M/ADsu = 0 Figure 7b) or mobilized stress ratio.
• V/Asu ≈ 0.5Vult ≈ 3; H/Asu ≈ 0; M/ADsu
≈ 0.65 4.2 Skirted Foundations / Suction
These correspond to the analytical solution Caisson
of a vertically loaded circular footing (or All early O&G GBSs in the North Sea
Prandtl failure mechanism with shape had some sort of ‘skirts’ below the base of
correction factor), sliding on a ‘rough’ the foundation (e.g., Randolph et al. 2005).
interface (τf = αsu, α=1) for the horizontal These skirts provided additional resistance
case, and analytical solution for a scoop to horizontal loading (sliding) as well as
mechanism at maximum moment uplift capacity. Figure 9b illustrates three
(Gourvenec & Randolph 2003). Interaction possible failure mechanisms for skirted
between V, H, and M loads will change the foundations. The addition of skirts moves
shape of the failure surface, and therefore the shallow foundation failure mechanism to
the resistance of the foundation system. deeper soil layers, and possible stronger
The above discussion is applicable soils (9b-i). Additional resistance comes
for undrained loading conditions (i.e., V > from extension of the failure surface through
15). Shallow foundations for offshore wind the soils above the foundation (e.g.,
turbines are up to an order of magnitude Skempton 1951, Gourvenec et al. 2006).
smaller in diameter than those for offshore For deep skirts (Figure 9b-ii), a flow
O&G structures, 15 to 25m as compared to round mechanism may generate at the base
50 to 150m. Undrained analyses may not be of the skirted foundation, and resistance is
applicable for design in rapidly drainage mobilized at the base (Qb=qb·Ab) as well as
soils, such as sands. Drained analysis of the along the sides (shaft) of the skirted
ultimate capacity of shallow foundations is foundation (Qs=τfπDL, where τf is the unit
less developed than that for undrained soils, shaft friction along the side of the skirted
yet will still depend on the interaction foundation). If there is a gap between the top
between V, H, and M loads. Typical plate of the skirted foundation and the
normalized parameters for drained loading internal soil such that plug movement can
of sands include (i) v=V/V0; (ii) h=H/V0; occur, and the plug resistance (internal shaft
and (iii) m=M/DV0; where V0 is the initial friction, Qs,in) is less than the end bearing

18

Schneider, J.A., and Senders, M. 2010. Foundation design – a comparison of oil and gas
platforms with offshore wind turbines, Journal of the Marine Technology Society, 44(1), 32-51.
resistance (Qb), the failure mechanism mechanism 9b-iii is typically used for
shown in Figure 9b-iii may occur. In this evaluation of installation resistance, as
case, end bearing resistance (qb) only occurs discussed by Senders et al. (2007), Senders
on the annulus of the skirted foundation and & Randolph (2009), among others. Shaft
internal and external shaft friction contribute friction and end bearing of suction caissons
to the ultimate vertical resistance. and skirted foundations with a deep failure
Within the O&G industry, mechanism are typically analyzed using
environmental loading of skirted similar methods as those used to evaluate τf
foundations are typically analyzed as and qb for low displacement (open ended)
shallow foundations discussed in the piles (e.g., Randolph 2003, Senders &
previous section (e.g., Bransby & Randolph Randolph 2009).
1998, Gourvenec & Randolph 2003,
Randolph et al. 2005, Gourvenec 2007). 4.3 Piled Jackets / Tripods
Failure occurs due to combined V, H, and M Piled jackets and tripods tend to resist
loading. This will be a similar case for the moment resulting from horizontal storm
offshore wind turbines on single, large loading through a push-pull system, as
diameter skirted foundations, such as that at illustrated in Figure 2a-i. Horizontal load
Fredrikshaven discussed by Ibsen (2008). must also be overcome by lateral resistance
For deeper water fixed offshore wind turbine of the piles. Pile jacket structures are
foundations supported by tripods or jacket common in the O&G industry and are
structures, moment induced by horizontal typically founded on driven piles for ground
loading is resisted by push-pull (e.g., Figure conditions consisting of soils (i.e., sand, silt,
2a-i), and the failure mechanisms in Figure clay, etc.). Bored piles for rock or calcareous
9b-ii or 9b-iii typically control behavior. deposits will not be addressed in this paper.
Horizontal load must also be less than the Failure mechanisms for driven piles are
lateral resistance of the caissons. This will illustrated in Figure 9c, and are similar to
be similar to the suction caissons for the those for the ‘deep’ failure mechanisms of
Draupner E and Sleipner SLT offshore O&G skirted foundations / suction caissons
jacket structures discussed by Bye et al. (Figure 9b-ii, 9b-iii). Like suction caissons,
(1995). the installation resistance of a driven pile
In this section we have combined will typically be controlled by the coring
discussion of skirted shallow foundations failure mechanism illustrated in 9c-ii (e.g.,
with suction caissons. Suction caissons are a Stevens 1988) while resistance to static
special case of shallow foundations or deep loading is controlled by the plugged failure
foundations (depending upon geometry), but mechanisms illustrated in 9c-i (e.g.,
are distinguished by the installation method. Randolph et al. 1991). Due to the slender
Suction is used to install the foundation, but geometry, driven piles in clay soils gain a
pumps are typically removed after majority of their resistance from shaft
installation and no additional long term friction (τf). When piles are founded in
resistance results from applied ‘suction.’ dense sands, 50 percent of the ultimate
Resistance during installation must also be resistance may come from end bearing (Qb),
evaluated, and often differs from that which depending upon soil layering. Cone
controls resistance to storm loading. Failure penetration test tip resistance can provide

19

Schneider, J.A., and Senders, M. 2010. Foundation design – a comparison of oil and gas
platforms with offshore wind turbines, Journal of the Marine Technology Society, 44(1), 32-51.
Bias towards (a) pile tip depth and (b) soil relative density for API (2000/2007) main text
method for axial shaft friction of piles in siliceous sand (after Schneider et al. 2008)

(c) Database calibration of API (2000/2007) (d) Potential bias in API (2000/2007) main
main text method for axial shaft friction of text method for axial shaft friction of piles in
piles in clay clay (after Karlsrud et al. 2005)

Figure 10. Bias in offshore methods for calculating shaft friction on pile foundations

the most reliable estimate of end bearing for friction involves larger degrees of
both open and closed ended piles (e.g., Xu et empiricism. When evaluating pile end
al. 2008), although evaluation of shaft bearing resistance, differences in the size of

20

Schneider, J.A., and Senders, M. 2010. Foundation design – a comparison of oil and gas
platforms with offshore wind turbines, Journal of the Marine Technology Society, 44(1), 32-51.
a small diameter cone penetrometer and a piles (or suction caisson) in siliceous sands
large diameter pile must be accounted for. published in API (2007) have not yet been
Cone penetration tip resistance should be adopted in the offshore wind turbine
averaged up to 4 equivalent pile diameters industry in the DNV (2007) standard.
below the tip of the pile, and approximately Historically, shaft friction of piles in
8 equivalent pile diameters above the pile tip clays has been estimated from undrained
(e.g., Schmertmann 1978, Xu et al. 2008). shear strength and a modification factor, τf =
The American Petroleum Institute α·su. For evaluation of offshore piles in the
(API) Recommended Practice for Fixed insensitive high plasticity NC deposits
Offshore Structures (RP 2A) has relied on (su/σ'v0 = 0.2 to 0.3) of the GOM, α values
bearing capacity theory for tip resistance and are often taken as close to unity (e.g.,
a lateral earth pressure coefficient based Randolph & Murphy 1985, API 2000,
approach for evaluation of pile shaft Figure 10c). Values of α tend to decrease
capacity (e.g., API 1969, 2000). When with increases in soil state, quantified using
designing in areas outside of the GOM that su/σ'v0, and are often taken as approximately
contain significant amounts of sandy soils, 0.4 for heavily overconsolidated clays of the
such as the North Sea, API RP2A has been North Sea. The successful use of α values
deemed to potentially be overly conservative does not occur because the failure at the
for some states of sandy soils and soil-pile interface is an undrained failure in
unconservative for others (e.g., Toolan et al. the clay, but rather that the undrained
1990). When comparing the design method strength correlates to the increase in radial
to a database of case histories, as shown in effective stress that results from installation
Figure 10a and 10b, this bias is clear. The of a displacement pile (e.g., Lehane &
API (2000) method overpredicts short term Jardine 1994, Randolph 2003). The shaft
capacity of long piles in ‘loose’ sands and friction values from Figure 10c are for the
underpredicts capacity of short piles in fully equalized strength that occurs a
‘dense’ sands. Since the bias is a function of significant time after foundation installation.
both soil density and pile geometry, Over time, additional tests have been
alternative method formulations were performed on foundations in a variety of soil
developed. The API (2007) update now conditions that were outside of the database
includes four CPT based methods for analyzed by Randolph & Murphy (1985).
evaluation of the shaft friction of piles in Additional data have highlighted much
sands. These methods evaluate changes in larger ‘uncertainties’ than have been
radial effective stress due to installation of a presumed for evaluation of the shaft friction
pile with a correlation to cone tip resistance of piles in clay. These additional data points
that is modified by pile geometry. In certain are illustrated in Figure 10d, and the bias
cases the different methods deviate between new data and previous trends have
significantly from each other, and a been described as a function of pile length
discussion of this bias is provided by Lehane (Kolk & Van der Velde 1996), plasticity
et al. (2005) and Schneider (2007, 2009). It index (Karlsrud et al. 2005), and/or
is noted that modifications to minimize sensitivity (Jardine et al. 2005). These
unconservative bias in the API (2000) differences require additional study and
design method for shaft friction of driven caution should be used when extrapolating

21

Schneider, J.A., and Senders, M. 2010. Foundation design – a comparison of oil and gas
platforms with offshore wind turbines, Journal of the Marine Technology Society, 44(1), 32-51.
to design situations in soft sensitive clays of have a diameter on the order of 3 to 5m, and
low plasticity. L/D between 4 and 20 with an average of 8.
In addition to vertical loads, piled Piles for offshore O&G structures will
foundations for jacket structures and tripods therefore typically fail with a flexible mode,
must also resist lateral loads. Resistance to illustrated in Figure 9d-ii, while monopile
lateral loads are discussed in more detail for for offshore wind turbines will typically fail
the next section on monopiles. in a rigid mode illustrated in Figure 9b-i
(e.g., Poulos 1989).
4.4 Monopiles Analytical solutions for ultimate
Multi-well structures for the offshore capacity and deformation of rigid and
O&G industry are typically large enough flexible laterally loaded piles typically
that designing a foundation with a single perform well (e.g., Poulos 1989, Mayne et
pile, or a monopile, would not provide al. 1995), although evaluation of cyclic
sufficient support. However, monopiles are degradation is highly empirical. The
the dominant foundation type for offshore empirical degradation factors have been
wind turbines constructed to date. developed for flexible rather than rigid pile
Evaluation of resistance for installation of behavior, and are likely not directly
monopiles follows the failure mechanism of applicable to the case of offshore wind
axially loaded piles in Figure 9c-ii. When turbine foundation design.
subject to large horizontal wind and wave Recent work has indicated that the
loads, monopile foundations for offshore excess progressive rotation as compared to
wind turbines will need to resist large lateral the static case (∆θ/θs) for an offshore wind
loads. There is significant experience for turbine founded on a rigid monopile in sand
evaluation of lateral loads for offshore is related to the maximum and minimum
foundations (e.g., Matlock 1970, Reese et al. moment of the load cycle (Mmax, Mmin), the
1974, Murff & Hamilton 1993, Long & static moment capacity of the pile (MR), the
Vanneste 1994, Dunnavant & O’Neil 1989), relative density of the soil (Rd), and the
although, this experience and these design number of cycles (N) (LeBlanc et al. 2010):
methods may not be directly applicable to ∆θ M  M 
lateral loading of monopile foundations. = Tb  max , R d  ⋅ Tc  max  ⋅ N 0.31
Typical offshore piles in the GOM have a θs  MR   M min 
ratio of length (L) to diameter (D), where Tb and Tc are functions of the
slenderness ratio (L/D), on the order of 45 to bracketed properties. It should be noted that
105 for piles installed prior to 1980 and 20 these model tests require some larger scale
to 70 for piles installed after 1980. Increase field validation and the power function on N
in pile driving hammer sizes has led to the may vary with sand type, as illustrated for
use of larger diameter higher capacity piles laboratory element tests in Figure 7a.
(e.g., McClelland 1974). The stronger Whether laboratory model tests are
glacial deposits of the North Sea typically recommended for each site or cyclic
used shorter piles, with L/D on the order of laboratory element tests can be used to
20 to 60, although slender piles (L/D ≈ 130) modify the updated empirical expressions
have also been used in some cases. In requires further validation.
contrast, monopile foundations typically

22

Schneider, J.A., and Senders, M. 2010. Foundation design – a comparison of oil and gas
platforms with offshore wind turbines, Journal of the Marine Technology Society, 44(1), 32-51.
4.5 Anchors al. 2005, Senders & Kay 2002, Randolph &
A wide range of floating structures for Gourvenec 2010):
offshore wind turbines is currently being • Gravity anchors (box, grillage and berm)
researched. Substructures adopted from the • Anchor piles
O&G industry include: • Suction caissons
• TLP (e.g., Blue H) • Dynamically penetrating (torpedo)
• Mini TLP / Sea Star (e.g., MIT / NREL) anchors
• Spar (Hywind, Sway) • Vertically loaded drag anchor (VLA)
• Semi-sub (e.g., Gusto, Principal Power) • Plate anchors (e.g., suction embedded,
Floating structures may be anchored using SEPLA)
near vertical mooring lines or catenary • Drag anchors (fixed fluke)
mooring lines, depending upon the Failure mechanisms are broadly similar to
substructure, and a wide range of anchoring those discussed for fixed offshore structures
solutions currently exist (e.g., Randolph et and illustrated in Figure 11, although

Figure 11. Drained and undrained failure mechanisms for vertically and horizontally loaded
anchors

23

Schneider, J.A., and Senders, M. 2010. Foundation design – a comparison of oil and gas
platforms with offshore wind turbines, Journal of the Marine Technology Society, 44(1), 32-51.
foundations are predominantly loaded in This paper has provided an overview of
tension for vertically moored structure and considerations for substructure selection and
loaded laterally (depending upon padeye foundation design for offshore structures.
location, Randolph et al. 1998) for catenary Specifically, evolution of concepts in the
moored structures. Capacity of anchor piles, O&G industry have been reviewed for GOM
suction caissons, and dynamically and North Sea developments highlighting
penetrating (torpedo) anchors are controlled parallels and differences to the offshore
by the same mechanisms as those for axially wind turbine industry. It is concluded that
and laterally loaded piles and caissons offshore O&G substructure and foundation
(Figure 9c, 9d), although there is no end concepts evolved differently in the GOM
bearing component during uplift and North Sea due the differences in:
(Qb=0=Qann). Plate and drag anchor (i) water depth; as well as
resistance in clay soils is predominantly (ii) soil conditions.
related to soil undrained shear strength Additionally, design of offshore wind
through a capacity factor. A reverse bearing turbine foundations is similar to offshore
capacity mechanisms initially forms and O&G foundations except that for offshore
resistance is controlled by a flow round wind turbines:
failure (Figures 11a, 11b). Failure (i) stricter specifications apply related to
mechanisms for plate and drag anchors dynamic behavior;
during drained penetration (sandy soils, typ.) (ii) tighter tolerances apply on rotation;
are more complex, and resistance is (iii) monopiles typically deform in a rigid
controlled by friction and dilation angles manner rather than a flexible manner,
(e.g., Figure 11c; Vermeer & Surjiadi 1985, such that empirical relationships for
White et al. 2001, White et al. 2008: Figure cyclic loading typical of piles for
11d; Neubecker & Randolph 1996). offshore O&G structures are not
While it is common to assume that directly applicable to the offshore wind
design of drag anchors only requires soil industry; and
type and anchor size through the ‘anchor (iv) failure typically does not lead to
efficiency’, this misconception will result in human casualties and costs associated
design with high levels of uncertainty. with failure are relative low.
Therefore the O&G industry adopts a policy Recommended practice / design
where every drag anchor needs to be proof- standards exist for both offshore O&G
loaded to a certain percentage of its ultimate foundations and foundations for offshore
capacity before it is approved. This is a wind turbines, and while these standards are
costly exercise and design using analysis useful for reproducing designs in similar
based on the failure mechanisms illustrated situation, they are not as useful for designing
in Figure 11 provides a more rational in new conditions. These standards often
framework that will lead to more reliable rely on empiricisms and may not reflect
extrapolation of offshore O&G experience appropriate failure mechanisms, such as
to offshore wind turbine foundation design. those discussed in this paper. Continued
research into offshore foundation behavior is
5 Summary, Conclusions & needed to update empiricism in design
Recommendations method formulations that better reflect

24

Schneider, J.A., and Senders, M. 2010. Foundation design – a comparison of oil and gas
platforms with offshore wind turbines, Journal of the Marine Technology Society, 44(1), 32-51.
failure mechanisms of offshore wind turbine Bolton, M.D. 1986. The strength and
structures. This new research should include dilatancy of sands. Géotechnique.
small scale physical models, field scale 36(1): 65-78.
physical models, centrifuge tests, analytical Bransby, M.F. and M.F. Randolph 1998.
studies, as well as numerical simulations. Combined loading of skirted
foundations. Géotechnique. 48(5): 637-
6 Acknowledgements 655.
The authors would like to thank Prof. Peter Bye, A., C. Erbrich, B. Rognlien and T.I.
Marshall and Dr. Andrew D. Deeks for Tjelta 1995. Geotechnical design of
helpful review comments on the manuscript. bucket foundations. In: Proc. Annual
The first author would also like to Offshore Technology Conference,
acknowledge financial support from the OTC7793. Houston: OTC.
UW-Madison Graduate School. Claussen, C.J.F. 1976. The Condeep story.
In: Offshore Soil Mechanics, ed. P.
7 References George and D.M. Wood, 256-270.
Cambridge: Cambridge University
API 1969. Recommended Practice for Engineering Department.
Planning, Designing, and Constructing Clukey, E.C., Morrison, M.J., Garnier, J.,
Fixed Offshore Platforms, API RP2A, and Corté, J.F. (1995). The response of
1st Edition. Dallas, TX: American suction caissons in normally
Petroleum Institute. consolidated clays to cyclic TLP
API 2000. Recommended Practice for loading conditions, In: Proc. 27th
Planning, Designing, and Constructing Annual Offshore Technology
Fixed Offshore Platforms – Working Conference, pp. 909-918. OTC7796.
Stress Design, API RP2A, 21st Edition. Houston: OTC.
Washington D.C.: American Petroleum Collins. I.F. and M. Boulbibane 2000.
Institute. Geomechanical analysis of unbounded
API 2007. Recommended Practice for pavements based on shakedown theory.
Planning, Designing and Constructing J. Geotech. Geoenv. Eng., ASCE.
Fixed Offshore Platforms-Working 126(1): 50-59.
Stress Design, RP2A, Revisions to Det Norske Veritas (DNV) 2007. Design of
Edition: 21, Supplement 2. Washington offshore wind turbine structures,
D.C.: American Petroleum Institute. Offshore Standard DNV-OS-J101: 142
Atkinson, J. 2007. The Mechanics of Soils pp.
and Foundations, 2nd Edition, London: Dunnavant, T.W. and M.W. O’Neil 1989.
Taylor & Francis. 442 pp. Experimental p-y model for submerged,
Basquin, O.H. 1910. The exponential law of stiff clay, J. of Geotech. Eng., ASCE.
endurance tests. Proceedings ASTM, 115(1): 95-114.
11: 625-630. Ellers, F.S. 1982. Advanced offshore oil
Bea, R.G., Z. Jin, C. Valle and R. Ramos platforms, Scient. Am. 246(4): 31-41.
1999. Evaluation of reliability of Eribrich, C.T. 2005. Australian frontiers –
platform pile foundations. J. Geotech. spudcans on the edge. In: Proc. Int.
Geoenv. Eng., ASCE. 125(8): 696-704. Sym. Frontiers in Offshore

25

Schneider, J.A., and Senders, M. 2010. Foundation design – a comparison of oil and gas
platforms with offshore wind turbines, Journal of the Marine Technology Society, 44(1), 32-51.
Geotechnics. pp. 49-74. London: Taylor Harr, M.E. 1987. Reliability based design in
& Francis. civil engineering. New York: Dover
Finnie, I.M.S. and M.F. Randolph 1994. Publications. 291 pp.
Punch-through and liquefaction induced Houlsby, G.T., Ibsen, L.B. and Byrne B.W.
failure of shallow foundations on 2005a. Suction caissons for wind
calcareous sediments. In: Proc. Int. turbines, In: Proc. Int. Sym. Frontiers in
Conf. on Behavior of Offshore Offshore Geotechnics. pp. 75-93.
Structures, BOSS'94, pp. 217 – 230. London: Taylor & Francis.
Boston: Pergamon. Houlsby, G.T., R.B. Kelly, J. Huxtable and
Fisk, H.N. and B. McClelland 1959. B.W. Byrne 2005b. Field trials of
Geology of the continental shelf off suction caissons in clay for offshore
Louisiana: Its influence on offshore wind turbine foundations,
foundation design. Bulletin of the Géotechnique. 55(4): 287-296.
Geological Society of America. 70: Ibsen, L.B. 2008. Implementation of a new
1369-1394. foundations concept for Offshore Wind
George, P. and D.M. Wood, ed. 1976. farms. In: Proc. 15th Nordic Geotech.
Offshore Soil Mechanics. Cambridge: Meeting.
Cambridge University Engineering Jardine, R. J., F.C. Chow, R.F. Overy and
Department. 468 pp. J.R. Standing 2005. ICP design
Gerwick, B.C. 2007. Construction of Marine methods for driven piles in sands and
and Offshore Structures. Boca Raton: clays. London: Thomas Telford. 97 pp.
CRC Press. 840 pp. Karlsrud, K., C.J.F. Clausen, and P.W. Aas
Gottardi, G., G.T. Houlsby and R. 2005. Bearing capacity of driven piles
Butterfield 1999. The plastic response in clay, the NGI approach, In: Proc. Int.
of circular footings on sand under Sym. Frontiers in Offshore
general planar loading. Géotechnique. Geotechnics. pp. 775-782. London:
49(4): 453-470. Taylor & Francis.
Gourvenec, S. 2007. Failure envelopes for Khorshid, M.S., B.C. Haggerty and R. Male
offshore shallow foundations under 1988. Development of geotechnical
general loading. Géotechnique. 57(9): aspects of the investigation programme.
715-728. In: Engineering for Calcareous
Gourvenec, S. and M.F. Randolph 2003. Sediments: Proc. of the Int. Conf. on
Failure of shallow foundations under Calcareous Sediments, Vol. 2. pp. 377-
combined loading. In: Proc. 13th Eur. 386. Rotterdam: Balkema.
Conf. on Soil Mech. and Geotech. Eng. Kolk, H.J. and E. Van der Velde 1996. A
pp 583-588. Prague: CGtS. reliable method to determine friction
Gourvenec, S.,M.F. Randolph and O. capacity of piles driven into clays, In:
Kingsnorth 2006. Undrained bearing Proc. 28th Offshore Technol. Conf.,
capacity of square and rectangular OTC 7993, pp. 337-246. Houston:
footings. Int. J. Geomech., ASCE. 6(3): OTC.
147-157. Kühn, M. 1997. Soft or Stiff: A
Fundamental Question For Designers of
Offshore Wind Energy Converters. In:

26

Schneider, J.A., and Senders, M. 2010. Foundation design – a comparison of oil and gas
platforms with offshore wind turbines, Journal of the Marine Technology Society, 44(1), 32-51.
European Wind Energy Conference '97. of laterally-loaded drilled shafts in clay.
4 pp. EWEC. J. Geotech. Eng., ASCE. 121(12): 827-
LeBlanc, C., G.T. Houlsby and B.W. Byrne 835.
2010. Response of stiff piles in sand to McClelland, B. 1974. Design of deep
long term cyclic lateral loading. penetration piles for ocean structures. J.
Géotechnique. online. Geotech. Eng. Div., ASCE. 100(GT7):
Leffler, W.L., R. Pattarozzi and G. Sterling 709-747.
2003. Deepwater Petroleum McClelland, B., and W.R. Cox 1976.
Exploration & Production: A Performance of pile foundations for
Nontechnical Guide. Houston: fixed offshore structures, In: Behaviour
Pennwell Publishing. 166 pp. of Offshore Structures, BOSS ’76, pp.
Lehane, B.M. and R.J. Jardine 1994. 1-17. The Norwegian Institute of
Displacement-pile behavior in a soft Technology.
marine clay. Can. Geotech. J. 31(2): Murff, J.D. and J.M. Hamilton 1993. P-
181-191. ultimate for undrained analysis of
Lehane, B.M., J.A. Schneider and X. Xu laterally loaded piles. J. Geotech. Eng.
2005. A review of design methods for Div., ASCE. 119(1): 91-107.
offshore driven piles in siliceous sand. Manwell, J.F., Elkinton, C.N., Rogers, A.L.,
UWA Report GEO 05358. Perth: The and McGowan, J.G. 2007. Review of
University of Western Australia. 105 design conditions applicable to offshore
pp. wind energy systems in the United
Lehane, B.M., C.D. O'Loughlin, C. Gaudin States. Renewable and Sustainable
and M.F. Randolph 2009. Rate effects Energy Reviews. 11: 210-234.
on penetrometer resistance in kaolin. Musial, W., S. Butterfield, and B. Ram,
Géotechnique. 59(1): 41-52. 2006. Energy from offshore wind, In:
Liingaard, M. 2006. Dynamic behaviour of Proc. 2006 Offshore Technology
suction caissons, PhD Thesis, Aalborg Conference, OTC18355. 11 pp.
University, Denmark. Houston: OTC.
Long, J. and G. Vanneste 1994. Effects of Neubecker, S.R. and M.F. Randolph 1996.
cyclic lateral load on piles in sand. J. The static equilibrium of drag anchors
Geotech. Eng., ASCE. 120(1): 225-244. in sand. Can. Geotech. J. 33(4): 574-
Marshall, P.W., and R.G. Bea 1976. Failure 583.
modes of offshore platforms, In: PMB Engineering, Inc. 1996. Hurricane
Behaviour of Offshore Structures, Andrew Effects on Offshore Platforms:
BOSS ’76, pp. 579-635. The Phase II Joint Industry Project – Final
Norwegian Institute of Technology. Report, Report 229AB.
Matlock, H. 1970. Correlations for design of Poulos, H.G. 1988. Marine Geotechnics.
laterally loaded piles in soft clay, OTC London: Unwyn Hyman Ltd. 473 pp.
1204. In: Proc. 2nd Offshore Poulos, H.G. 1989. Pile behaviour – theory
Technology Conference, pp. 579-594. and application. Géotechnique. 39(3):
Houston: OTC. 365-415.
Mayne, P.W., F.H. Kulhawy and C.H.
Trautmann 1995. Laboratory modeling

27

Schneider, J.A., and Senders, M. 2010. Foundation design – a comparison of oil and gas
platforms with offshore wind turbines, Journal of the Marine Technology Society, 44(1), 32-51.
Randolph, M. F. 2003. Science and Richart, F.E. 1960. Foundation vibrations. J.
empiricism in pile foundation design. Soil Mech. Found. Div., ASCE,
Géotechnique, 53(10): 847-875. 86(SM4): 34 pp.
Randolph, M.F. 2004. Characterisation of Rølland, T. 2009. Wind power projects in
soft sediments for offshore applications. StatoilHydro. Presentation to LOG
In: Proc. Second Int. Conf. on Site Meeting, Molde, 15 January 2009.
Characterization, ISC’2, pp. 209-232. Saigal, R.K., D. Dolan, A. Der Kiureghian,
Rotterdam: Millpress. T. Camp and C.E. Smith 2007.
Randolph, M.F. and B.S. Murphy 1985. Comparison of design guidelines for
Shaft capacity of driven piles in clay. offshore wind energy systems. In: Proc.
In: Proc. 17th Offshore Technol. Conf., 2007 Offshore Technology Conference,
OTC 4883, pp. 371-377, Houston: OTC 18984-PP, 7 pp. Houston: OTC.
OTC. Schmertmann, J. H. 1978. Guidelines for
Randolph, M.F., E.C. Leong and G.T. cone test, performance, and design (No.
Houlsby 1991. One-dimensional FHWATS-78209): U.S. Federal
analysis of soil plugs in pipe piles. Highway Administration.
Géotechnique. 41(4): 587-598. Schneider, J.A. 2007. Analysis of piezocone
Randolph, M.F., M.P. O’Neill, D.P. Stewart data for displacement pile design, PhD
and C. Erbrich 1998. Performance of Thesis, Crawley: The University of
suction anchors in fine-grained Western Australia.
calcareous soils, In: Proc. Annual Schneider, J.A. 2009. Uncertainty and bias
Offshore Technology Conference, in evaluation of LRFD ultimate limit
OTC8831. 9 pp. Houston: OTC. state for axial loading of driven piles,
Randolph, M.F., M. Cassidy, S. Gourvenec Journal of the Deep Foundations
and C. Erbrich 2005. Challenges of Institute, 3(2): 38-49.
offshore geotechnical engineering. In: Schneider, J.A., M.F. Randolph, P.W.
Proc. of the 16th Int. Conf. Soil Mech. Mayne and N. Ramsey 2008a. Analysis
and Geotech. Eng.. pp. 123-176. of factors influencing soil classification
Rotterdam: Millpress. using normalized piezocone tip
Randolph, M.F. and S. Gourvenec 2010. resistance and pore pressure
Offshore Geotechnical Engineering. parameters, J. Geotech.
London: Taylor & Francis. 560 pp. Geoenvironmental Eng., ASCE.
Redding, J. 1976. Glacial genesis of North 134(11): 1569-1586.
Sea soils. In: Offshore Soil Mechanics, Schneider, J.A., X. Xu and B.M. Lehane
ed. P. George and D.M. Wood, 80-100. 2008b. Database assessment of CPT
Cambridge: Cambridge University based design methods for axial capacity
Engineering Department. of driven piles in siliceous sands. J.
Reese, L.C., W.R. Cox, and F.D. Koop. Geotech. Geoenvironmental Eng.,
1974. Analysis of laterally loaded piles ASCE. 134(9): 1227-1244.
in sand. Proceedings, 6th Annual Seed, H.B. and I.M. Idriss 1970. Soil moduli
Offshore Technology Conference, OTC and damping factors for dynamic
2080: pp. 473-485, Houston: OTC. response analysis. Rpt. No.
UCB/EERC-70/10, U.C. Berkeley.

28

Schneider, J.A., and Senders, M. 2010. Foundation design – a comparison of oil and gas
platforms with offshore wind turbines, Journal of the Marine Technology Society, 44(1), 32-51.
Senders, M. 2005. Tripods with suction Conf. on App. of Stress Wave Theory
caissons as foundations for offshore to Piles, pp. 861-868.
wind turbines on sand. In: Proc. Int. Terzaghi, K., and R.B. Peck 1948. Soil
Sym. Frontiers in Offshore Mechanics in Engineering Practice.
Geotechnics. pp. 397-403. London: Wiley: New York
Taylor & Francis. Toolan, F.E., M.L. Lings and U.A. Mirza
Senders, M. 2008. Suction caissons in sand 1990. An appraisal of API RP2A
for offshore wind turbine foundations, recommendations for determining skin
PhD Thesis, Crawley: The University friction of piles in sand. In: Proc., 22nd
of Western Australia. Offshore Technol. Conf., OTC 6422,
Senders, M. and S. Kay 2002. Geotechnical pp. 33-42. Houston: OTC.
suction pile anchor design in deep water Vermeer, P.A. and W. Surjiadi 1985. The
soft clays. In: Proc.7th Annual uplift resistance of shallow embedded
Conference on Deepwater Risers, anchors. In: Proc. 11th Int. Conf. Soil
Moorings and Anchorings. 50 pp. Mech. Found. Eng. pp. 1635-1638.
London: IBC. Rotterdam: Balkema.
Senders, M., M.F. Randolph and C. Gaudin Wasson, T. 1948. Creole field, Gulf of
2007. Theory for the installation of Mexico, Coast of Louisiana. In: SP 14:
suction caissons in sand overlaid by Structure of Typical American Oil
clay. In: Proc. of the 6th Int. Conf. on Fields, Vol. III, 281-298. American
Site Investigation and Geotechnics, pp. Association of Petroleum Geologists
367-382. London: SUT. (AAPG).
Senders, M. and M.F. Randolph 2009. CPT- White, D.J., A.W. Take and M.D. Bolton
based method for the installation of 2001. Centrifuge modeling of upheaval
suction caissons in sand. J. Geotech. buckling in sand. Int. J. Phys.
Geoenvironmental Eng., ASCE. 135(1): Modelling Geomech. 2(1): 19-28.
14-25. White, D.J., C.Y. Cheuk and M.D. Bolton
Sharma, S.S., and M. Fahey 2003. 2008. The uplift resistance of pipes and
Evaluation of cyclic shear strength of plate anchors buried in sand.
two cemented calcareous soils. J. of Géotechnique. 58(10): 771-779.
Geotech. Geoenvironmental Eng., Wilhoit, L. and C. Supan 2009. 2009
ASCE. 129(7): 608-618. Deepwater solutions & records for
Skempton, A.W. 1951. The bearing capacity concept selection (poster). Houston:
of clays. In: Proc. Building and Mustang Engineering.
Research Congress. pp. 180-189. Xu, X., J.A. Schneider and B.M. Lehane
London. 2008. Cone penetration test (CPT)
Spera, D.A. 1994. Chapter 12: Fatigue methods for end-bearing assessment of
design of wind turbines, Wind Turbine open- and closed-ended driven piles in
Technology, ed. D.A. Spera, 547-588. siliceous sand, Canadian Geotech. J.,
New York: ASME Press. 45(8): 1130-1141.
Stevens, R.F. 1988. The effect of soil plug
on drivability in clay, In: Proc. 3rd Int.

29

Schneider, J.A., and Senders, M. 2010. Foundation design – a comparison of oil and gas
platforms with offshore wind turbines, Journal of the Marine Technology Society, 44(1), 32-51.

View publication stats

You might also like