Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

Offshore wind turbine

foundations e analysis and 20


design
B.C. O’Kelly
Trinity College Dublin, Dublin, Ireland

M. Arshad
University of Engineering & Technology, Lahore, Pakistan

20.1 Foundation options for offshore wind-turbine


structures
The economically viable development of wind-farms depends on efficient solutions
being available for a number of technical issues, one aspect being the foundations.
The foundation choice largely depends on water depth, seabed characteristics, applied
loading, available construction technologies and, importantly, economic costs
(Malhotra, 2010). Offshore wind turbine (OWT) structures may be found on gravity
base, suction caisson, monopile, tripod or jacket/lattice foundations (collectively
categorized as bottom-mounted structures) or using floating platforms tethered to
the seabed (Fig. 20.1). The most widely adopted foundation choice in terms of its
ease of installation, economy and logistics is the monopile, a single large-diameter
hollow (pipe) pile; with an estimated 75% of all installed OWTs employing this
solution (Blanco, 2009; Fischer, 2011; Malhotra, 2010). Hence, the focus of this
chapter is on the geotechnical aspects of monopile foundations, with some discussion
also provided on the other main foundation options.
Monopiles are typically used in shallow water depths (ie, 20e40 m), but may
become too flexible for water depths of between w30 and 40 m, in which case
monopiles fitted with guy wires, or alternatively tripod and jacket/lattice structures,
are considered as economical alternatives. For greater depths, time-consuming
installation and the effect of soil degradation, which occurs around the pile shaft
at seabed level, make monopile foundation solutions prohibitive (Irvine et al.,
2003). Tripods consist of a large-diameter central, steel tubular section that is sup-
ported over its lower length by three braces (Fig. 20.1(d)), which are connected to
the seabed using different foundation types, including gravity base, suction bucket
or piles. In this manner, the loads applied to the OWT and its support structure are
mostly transferred axially (via the braces) to the seabed foundation. Complete
installation of a tripod foundation system having, for example, a water surface to
seabed length of up to 50 m typically takes two to three working days, often

Offshore Wind Farms. http://dx.doi.org/10.1016/B978-0-08-100779-2.00020-9


Copyright © 2016 Elsevier Ltd. All rights reserved.
590 Offshore Wind Farms

(a) (b) (c) (d) (e) (f) (g)

Sea level
0

10
Seabed
20
Water depth (m)

30

40

50 a) Gravity
b) Monopile
60 c) Monopile with guy wires
d) Tripod
70 e) Jacket/lattice structure
f) Tension leg with suction buckets
80 (ballast stabilized)
g) Buoy with suction anchor

Figure 20.1 Support structure options showing typical ranges of water-depth application
(adapted from Malhotra, 2010).

requiring special equipment for driving/drilling and working underwater (Esteban


et al., 2011; Fischer, 2011). A jacket structure (see Fig. 20.1(e)) is a lattice frame
comprising small-diameter steel struts that, similar to the tripod, is anchored to the
seabed using different foundation types. Complete installation of the jacket struc-
ture generally takes up to 3 days. These structures are particularly suited for severe
maritime weather conditions because of the additional structural stiffness and larger
moment arm (across the seabed) for reacting against the bending loads, compared
with monopile foundations. Jacket/lattice structures are also more adaptable to the
conditions encountered onsite, increasing their application range, with geometrical
variations of the substructure achieved relatively simply, but without altering the
stiffness of the overall structure (De Vries, 2007). It is estimated that by 2020,
50e60% of OWTs will be supported by monopiles, with a further 35e40% sup-
ported by jacket and tripod systems (Babcock and Brown Company, 2012). The
main reason for this shift is the attraction of jacket and tripod systems for deeper
sea locations that provide consistently higher wind speeds and hence greater
wind-energy production, which is a cube function of the wind speed (Tempel
and Molenaar, 2002).
Offshore wind turbine foundations e analysis and design 591

In the future, it is anticipated that floating structures, which are currently at the
research and development stage, will be commercially used, particularly for water
depths greater than 50 m (Saleem, 2011). Such floating platforms for wind turbines
will impose many new design challenges. Among these, tension-leg platform concepts
(Fig. 20.1(f)) are currently considered as more economical (Fischer, 2011) since the rigid
body modes of the floater are limited to horizontal translation (surge and sway) and rota-
tion around the vertical axis (yaw). For spare floater systems (Fig. 20.1(g)), buoyancy is
provided to the wind turbine structure by a long slender cylinder/capsule that protrudes
well below the water line (De Vries, 2007; Esteban et al., 2011; Fischer, 2011). For bar-
rage floater systems, the wind turbine structure is placed on a barrage and attached, via
anchor lines, to the seabed. The design of floating offshore wind turbines is discussed in
chapter ‘Energy storage for offshore wind farms’.

20.2 System of loading on offshore foundations


Offshore foundations are subjected to a combination of loading, namely: axial (self
weight) forces of the structure/machinery; repeating horizontal/lateral loads; bending
and torsional moments. Apart from the self-weight forces, this system of loading is
generated by environmental conditions and the operation of the machinery
(Fig. 20.2). The foundations must be designed to resist large numbers of wind and

Ambient turbulence

Wake turbulence

Operational and accidental loads

Icing

Marine growth

Waves

Currents

Seabed Scour

Figure 20.2 Environmental impacts on offshore wind turbines.


592 Offshore Wind Farms

hydrodynamic (ocean waves, current flow and tidal/swell action) load cycles of vary-
ing direction, amplitude and frequency occurring at the proposed site over the project’s
lifetime of typically 25 years or more (Sahin, 2004). Another variable/cyclic load
acting on OWT structures, depending on the geographic setting, is from ice sheets.
Seismic loading is considered as a special type of dynamic loading. However, the
focus of this chapter is on wind and wave loading since these are considered more
important in the assessment of the OWT structure’s fatigue life.
The wave loading (forces) acting on OWT structures is often greater than the wind
loading. However, in terms of the overturning (bending) moments generated, the
wave loading generally has only a minor role, compared with the rotorethrust reac-
tion to the wind loads. This occurs due to the smaller lever arm for the bending
moment generated by the wave loading, as compared with the overall tower length,
and longer lever arm of the rotor thrust in considering the overall overturning
moment acting on the foundation system (LeBlance, 2009). For instance, for OWT
monopiles installed in the North Sea, Byrne and Houlsby (2003) reported that the
rotor thrust reaction contributed to approximately 25% of the total horizontal load,
but generated approximately 75% of the total overturning moment. The density of
the medium must also be considered when comparing wind and wave loading,
with the density of sea water significantly greater than that of air. Hydrodynamic
loads generally only become significant for greater water depths and (or) wave
heights, which cause the lever arm of the wave load to increase, along with the in-
tensity of the lateral load generated by the water (Fischer, 2011). Fluctuations in
wind speed about a mean value impose repeated aerodynamic loads, although
when considering the dynamic behaviour of offshore structures, this cyclic nature
is generally insignificant compared with the repeated wave loads (Journée and Mas-
sie, 2001). Dynamic analysis of offshore structures that takes into account the fluc-
tuating wind load is necessary in cases where the wind field contains energy at
frequencies near the offshore structure’s natural frequency, although for monopile
foundations, the difference between these frequencies is usually high. When the
loading frequency gets closer to the structure’s natural frequency of vibration
(API, 2010), the repeating load can be termed dynamic load. This tends to excite
the structure dynamically, leading to resonance and the development of higher
stresses in the support structure and foundation, but more significantly to a higher
range of stresses; an unfavourable situation in considering fatigue life. A correct
evaluation of the total hydrodynamic load acting on an offshore structure must
consider the combined current flow and wave particle velocities. Morison et al.
(1950) formulated an equation to predict the wave loads acting on a vertical pile
exposed to horizontal sinusoidal oscillatory flow. In their equation, the linear inertial
force and adapted quadratic drag force (from real flows and constant currents) are
superimposed to obtain the resultant force acting on the projected area of the pile.
Morison’s formula is strictly limited for use with slender structural elements, charac-
terized by D/l < 0.2, where D is the diameter of the structural element between
seabed level and the transition piece, and l is the impinging wavelength of the ocean
wave. For larger offshore structures (eg, gravity foundations and OWT monopiles),
the wave field is significantly influenced and a diffraction regime emerges. Potential
Offshore wind turbine foundations e analysis and design 593

flow theory is more suitable for the calculation of wave loads on such structures
(Batchelor, 1967). However, a significant number of existing offshore structure de-
signs have employed Morison’s equation even though the criterion of D/l < 0.2 may
not have been fully satisfied (Haritos, 2007).
For geotechnical design, the relative proportions and importance of the different types
of loads applied essentially depend upon the kind of foundation system being considered.
For gravity base foundations, potential failure modes are in bearing capacity or excessive
settlement; hence the vertical (self-weight) loads are generally the major design consid-
eration (Malhotra, 2010). For monopile foundations, the lateral deflection (rotation)
response of the monopile largely controls the serviceability limit state of the whole struc-
ture. Hence, lateral loads and resulting moments are more critical compared with the ver-
tical loads for monopile foundations. In other words, the monopile’s response under
repeated lateral loading is the major design consideration, with the monopile design
dominated by considerations of its dynamic and fatigue responses under working loads,
rather than its ultimate load-carrying capacity. For instance, existing OWTs have rotor
diameters ranging between w90 and 120 m (power-generation capacities of 3e6 MW
(Tong, 2010)) and produce gravitational loading in the range of w2e8 MN. For
instance, Byrne and Houlsby (2003) reported vertical loading of 6 MN acting on the
monopile foundation for an anticipated 3.5 MW OWT located in the North Sea, with
lateral loading from wind and wave factors accounting for up to 66% of the vertical
loading. The precise magnitude of these loads will vary with the size of the installation,
the detailed design, and local environmental conditions. This scenario is more onerous
when the repeating lateral loading occurs at varying frequency, load amplitude and direc-
tion (Arshad and O’Kelly, 2013). At some critical level of load amplitude and (or) fre-
quency, the repeating lateral loads can cause significant reductions in the lateral soil
resistance for a monopile foundation (Ramakrishna and Rao, 1999).

20.3 General aspects of OWT monopile


foundation system
Monopiles with outer diameters of 4e6 m have been successfully installed, with
embedment (penetration) depths of between 20 and 40 m, depending upon the
wind-power generation capacity of the OWT (Peng et al., 2006; Achmus et al.,
2009; LeBlanc et al., 2010; Peng et al., 2010; Cuéllar et al., 2012; Pappusetty and
Pando, 2013). These monopiles are manufactured from steel tubular sections with
wall thicknesses (depending on installation and loading conditions) ranging from 55
to 150 mm (Haiderali et al., 2013), but more frequently 60e80 mm, and have an over-
all mass of up to 1000 tonnes. Depending on soil conditions, monopiles are typically
installed in the seabed using largeimpact hammers, by vibratory pile driving (vibropil-
ing), prebored piling, or drill/driving techniques (Malhotra, 2010; Igoe et al., 2013a).
Noise-mitigation measures, including bubble curtains, isolation casings (pile sleeve),
dewatered cofferdams, and hydro sound dampers, are employed during pile driving
in order to limit emissions to environmentally friendly levels (Saleem, 2011; OSPAR
594 Offshore Wind Farms

Commision, 2014). Complete installation of the monopile is usually achieved within


24 h (Junginger et al., 2004; Fischer, 2011; Saleem, 2011). The transition piece of the
OWT structure provides a means of correcting vertical imprecisions of the monopile
that may have arisen during its installation. The transition piece, which has a slightly
larger (or smaller) diameter than the monopile, is grouted to the monopile, with an
overlap. Mechanical shear connectors increase the reliability of this connection and
alleviate the effect of long-term grout shrinkage on the connection capacity. New
design features include grouted conical-shaped connections and shear keys (DNV,
2011). In some instances, the installed monopile can extend above the water surface,
connecting directly to the tower.
The monopile diameter and embedment length are primarily dependent on the
OWT’s power-generation capacity (an indirect measure of the applied loading), seabed
characteristics/soil properties and severity of environmental loading. OWT monopiles
invariably have slenderness (in terms of embedment length to diameter) ratio values of
less than 10 (typically ranging from 5 to 6), and the embedded portion of the monopile
is therefore considered to behave as a ‘rigid’ structure, for which rotation is prominent
over bending (Tomlinson, 2001; Peng et al., 2011). In other words, the surrounding
soil would fail in bearing capacity rather than the monopile failing over its embedded
length by plastic hinges; ie, its structural behaviour under lateral loading is defined by
its rotation as a rigid body. The embedment length must be sufficient for the monopile
to meet design criteria, including vertical stability and limiting horizontal deflection/
rotation over its design life. Limiting the rotation is more important for ‘rigid’ monop-
iles. As a general rule, under field loading, rotation of the monopile by up to 0.5 degree
from its vertical alignment (Achmus et al., 2009; LeBlanc, 2009; LeBlanc et al., 2010;
DNV, 2011) or lateral deflections occurring at seabed level of up to 120 mm (based on
practical experience) (Malhotra, 2010) are considered as limiting values for the proper
operation of the wind turbine. Transport of sediment from beneath the scour protection
(typically rock/stone layers) zone around OWT monopiles may cause sinking of the
scour protection, particularly for piles founded in sandy soil deposits. This alters the
natural frequency of the dynamic response in an unfavourable manner (van der Tempel
et al., 2004).
For a typical 5-MW OWT installed in the North Sea, a hub height of w95 m above
mean sea level and rotor diameter of 125 m would produce the approximate quasi-
static loading scenario (acting at the seabed level) given in Table 20.1. This example
scenario demonstrates that for the monopile foundation, horizontal loading from wind
and wave causes extremely high bending moments (resisted in lateral compression of
Table 20.1 Loading at seabed level for
monopile supporting 5-MW OWT
(Lesny and Wiemann, 2005)
Axial load 35 MN
Horizontal load 16 MN
Bending moment 562 MN m
Torsional moment 4 MN m
Offshore wind turbine foundations e analysis and design 595

the surrounding soil) and principally controls the foundation design. To ensure the
monopile is torsionally stable, sufficient circumferential shear resistance must be
mobilized at the pileesoil interface, although the torsional moment to be resisted is
usually small (Table 20.1). Further, the connections between the tower and transition
piece and between the transition piece and monopile foundation must be capable of
transferring these bending and torsional moments, with adequate factors of safety.

20.4 Offshore design codes and methods


At present, current practice for the analysis, design and installation of OWT monopiles
usually relies on general geotechnical standards which are complemented by more spe-
cific guidelines and semiempirical formulae developed mainly by the offshore oil/gas
industries (American Petroleum Institute (API), 2010; DIN, 2005; Det Norske Veritas
(DNV), 2011). However, large-diameter monopile foundations for current and future
OWT structures are well outside the scope of current experience and analysis/design
methods, including the API (2010) and DNV (2011) recommended practices, which
are largely based on limited field data obtained for relatively small-diameter (ie, flexible)
piles under low numbers of load cycles. For these standards, wave loading is of primary
concern when extrapolating to predict extreme events. However, designers of OWT
structures must consider wave and wind load spectra simultaneously (IEC, 2005).
Hence, careful consideration of these differences in applied loading, as well as other
inherent limitations underlying semi-empirical offshore oil/gas industry formulae, is
required in extrapolation of these formulations for the design of OWT monopile foun-
dations. Often these formulations cannot be applied with confidence by the offshore
wind-power industry to achieve optimum results and economy (Dobry et al., 1982).
Current design standards and guidelines on the serviceability of piles under cyclic
lateral loading are limited. Over its service life, a typical 2-MW OWT structure is sub-
jected to w100 cycles of 2.0-MN magnitude and 107 cycles of 1.4-MN magnitude in
lateral loading, which correspond to the serviceability limit state and fatigue limit state,
respectively (GL, 2005). Factors affecting the cyclic response include the diameter and
wall thickness of the monopile, its free spanning and embedded lengths, the soil prop-
erties, soilepile relative stiffness, the characteristics of the applied loading and the pile
installation method (Malhotra, 2010).

20.5 Investigation of monopileesoil behaviour


20.5.1 Soil behaviour and testing
Apart from very small strain levels of <103 strain (Atkinson and Sallfors, 1991;
O’Kelly and Naughton, 2008), the stressestrain relationship for soil is invariably highly
non-linear (inelastic), with the strength and stiffness properties often strongly dependent
on the stress history and stress path followed under loading. Seabed/sedimentary
deposits are typically cross-anisotropic (O’Kelly, 2006). Due to the initial and
stress-induced anisotropy (Naughton and O’Kelly, 2004), soil deformations can occur
596 Offshore Wind Farms

within the influence zone of the embedded monopile whenever changes occur in the
magnitudes of the three principal stresses acting on a soil element and (or) the orienta-
tions of the principal stress axes on account of the applied cyclic lateral loading. In this
regard, an analogy may be drawn between the soil behaviour in the immediate vicinity
of the OWT monopile’s shaft and that beneath a highway pavement under repeated
wheel loading of varying intensity. For both scenarios, changes in the magnitudes
and the directions of the principal stresses acting on a soil element occur simultaneously
during each load cycle.
The values of pertinent parameters used to describe the soil response under cyclic
loading are often determined using cyclic triaxial tests (Das, 2008), although the
axisymmetric system of cyclic axial loading and all around confinement pressure
acting on the test-specimen is usually not compatible with the generalized stress con-
ditions encountered in situ. An advancement on cyclic triaxial testing is provided by
the hollow cylinder apparatus (HCA) (O’Kelly and Naughton, 2005, 2009), which al-
lows independent control of the magnitudes of the three principal stresses, as well as
the orientation of the majoreminor principal stress axis. The HCA is ideally suited for
simulating cyclic multidirectional loading conditions on cross-anisotropic test speci-
mens. Generalized stress-path testing can be performed to simulate the stress history
and in-service loading conditions at specific locations in the soil foundation (Naughton
and O’Kelly, 2005; O’Kelly and Naughton, 2009). In many practical situations, labo-
ratory testing may become too laborious, expensive and time-consuming. In situ
testing techniques, including the Cone Penetration Test method (Doherty et al.,
2012; Igoe et al., 2013b), can be used for offshore site investigations and afford
another approach in the determination of pertinent design parameter values.

20.5.2 Physical testing of monopile installations


Compared with the single one-off structure typically used by the offshore oil and gas
industries, the strategy of predictions and in situ measurement of the behaviour for
every OWT foundation in a wind farm project is often technically laborious, econom-
ically unrealisable and hence not practicable. Full-scale pile tests are expensive and
time-consuming to perform. In this case, an economical design procedure can be
efficiently and confidently applied provided experimental verifications are achievable
to understand the monopileesoil behaviour through model studies. For instance,
El Naggar and Wei (1999), D€ uhrkop (2009), LeBlanc et al. (2010), Peng et al.
(2011) and Arshad and O’Kelly (2014) have reported investigations of scaled model
monopiles subjected to many thousands of lateral load cycles in order to validate nu-
merical predictions and design rules. For example, the apparatus developed by Arshad
and O’Kelly (2014) allows the investigation of the effects of lateral loading direction,
amplitude, frequency and waveform shape on the response of a model monopile at 1 g.
Scaling laws (Lai, 1989; Muir Wood et al., 2002; LeBlanc et al., 2010; Bhattacharya
et al., 2011; Cuéllar et al., 2012) preserve constitutive and kinematic similarities
between the prototype and model such that the test results can be directly applied,
once the soil at site is the same as that used for the scaled tests. The best way is to
link the model tests with element test parameters (Lombardi et al., 2013).
Offshore wind turbine foundations e analysis and design 597

20.5.3 Strain accumulation models


20.5.3.1 Models for long-term cyclic lateral loading
The strain accumulation in a soil element under repeated loading is dependent on its
engineering properties, the stress path/level and the number of load cycles (Niemunis
et al., 2005; Karg, 2007). Many methods with different complexity and levels of
acceptability have been developed to predict the strain accumulation in a soil element
subjected to large numbers of load cycles. However, a major limitation to their appli-
cation for offshore foundation design calculations is that only a few (listed in Ta-
ble 20.2) consider simulation of the monopileesoil foundation interaction under
cyclic lateral loading. Considerable differences of opinion exist regarding the rate of
cyclic strain accumulation; for instance, power function (Little and Briaud, 1988)
and logarithmic trend (Lin and Liao, 1999) relationships have been proposed. The
most popular of these, and the method recommended in offshore design codes,
including the API (2010) and DNV (2011), is based on the Winkler model (ie, pile
acts as a beam supported by a series of uncoupled non-linear elastic springs which
represent the soil reaction) and is commonly referred to as the pey method.

20.5.3.2 The pey method


In general terms, a pey curve is typically obtained by plotting the lateral soil resistance
(p) response against the lateral deflection (y) of the pile, arising from the action of a
horizontal load (H) applied at the pile head (Fig. 20.3). Fig. 20.3(b) shows the lateral
soil resistance generated around the pile circumference at a particular depth (xt) and the
corresponding pile deflection (yt) response. The effects of soil stratification,
non-linearity and other soil properties are automatically taken into account by deter-
mining pey curves specific to different depth ranges along the pile embedment length

Strain accumulation models that consider pileesoil


Table 20.2
interaction
Little and Briaud (1988). See also Kuo et al. (2012), numerical studies; Peralta and Achmus
(2010), model studies in medium-dense sand
Long and Vanneste (1994)
Lin and Liao (1999). See also Verdure et al. (2003) and Li et al. (2010), centrifuge tests on
miniature piles in sand
High cycle accumulation method. See Niemunis et al. (2005), finite-element calculations
using ABAQUS; Wichtmann et al. (2010)
Achmus et al. (2009)
LeBlanc (2009) and LeBlanc et al. (2010)
Bienen et al. (2012)
Klinkvort and Hededal (2013)
598 Offshore Wind Farms

(Fig. 20.3(c)). For instance, soil stiffness typically increases with depth, which is re-
flected by increasing values of the spring stiffness (Epy), defined as the secant modulus
of the pey curve. The pile deflection that develops under given loading conditions and
constraints can be predicted by implementing the related pey curve in a simple
non-dimensional framework (eg, using the approach after Randolph (1981)), or in nu-
merical methods using computer software such as COM624P (Wang and Reese,
1993).

H
(a)

xt

(b) yt
yt

pt

V
V
M
M
(c) H
H

Epy 1 Epy 1
y
p

Epy 2 Epy 2
y
p

Epy 3
Epy 3
y
Figure 20.3 pey method of analysis for laterally loaded pile. Note: Epy, spring stiffness; H,
horizontal/lateral load; M, bending moment; V, axial load. (a) Deflected shape of monopile
(vertical section). (b) Soil resistance pt exerted due to pile deflection yt for a specific depth xt
(horizontal section). (c) Winkler model approach with changing shape of pey curves against
depth.
Offshore wind turbine foundations e analysis and design 599

20.5.3.3 Limitations of design approaches based on pey method


Current pile design methodology based on pey curves, as described in API (2010) and
DNV (2011) recommended practice, has gained broad recognition on account of the
low failure rate for piles in service over many decades. However, some caution is neces-
sary in applying this methodology to the design of OWT monopiles since the approach
may then be applied outside of its verified range. Further, several important design issues
may not be properly taken into account. Firstly, these standards (codes) rely on methods
built upon limited empirical data obtained for long (flexible) piles, for which bending is
significant. In contrast, existing and planned OWT monopile foundations, having slender-
ness ratio values of less than 10, exhibit rigid pile behaviour (Achmus et al., 2009; Peng
et al., 2011). Secondly, pile rotation (rather than deflection) is generally more prominent
for OWT foundations, with the rotation occurring about a point located approximately one
diameter above the pile base. Thirdly, the pey curves for cyclic loading presented in API
(2010) and DNV (2011) were primarily formulated for the evaluation of the ultimate
lateral load-carrying capacity mobilized under relatively few load cycles (<200). In
contrast, OWT structures experience many millions of low-amplitude cycles over their
service life. These points cast doubt on the appropriateness of API (2010) and DNV
(2011) recommended practice (based on pey curves) in predicting the in-service behav-
iour of OWT monopiles. Fourthly, despite the pey stiffness parameter Epy being a pilee
soil interaction parameter, these standards only consider the soil properties in formulating
the pey curves. The influence of the pile properties on the mobilized pey curves remains
an open question. Fifthly, these standards always predict a degradation of the absolute
secant stiffness for monopiles in sandy soil, irrespective of the density state or number
of lateral load cycles. Model studies on monopiles in loose (LeBlanc et al., 2010;
Bhattacharya et al., 2011) and dense (Cuéllar et al., 2012; Rosquoet et al., 2007) sands
suggest that the foundation stiffness (cyclic) increases with the number of load cycles
due to densification of the soil next to the pile (LeBlanc, 2009; Bhattacharya and Adhikari,
2011; Cuéllar et al., 2012). Furthermore, the API (2010) and DNV (2011) approaches do
not provide a means of calculating the accumulated pile deflection (rotation) occurring
during cyclic loading, rather they only suggest an empirical factor for this purpose.

20.6 Design of OWT foundation


20.6.1 Resonance in structural system
The basic driving motive for the design is to avoid the occurrence of resonance in the dy-
namic behaviour of the structural system under in-service loading. The OWT structure is
considered too flexible if its natural frequency falls within the ‘softesoft’ zone (region
before the first excitation (‘1P’) frequency range of typically 0.17e0.33Hz) and too rigid,
heavy and expensive if its natural frequency falls within the ‘stiffestiff’ zone (region after
the blade-passing frequency range (‘3P’) of typically 0.5e1.0 Hz for three-bladed tur-
bines). Another important reason for avoiding the ‘softesoft’ frequency region is that
wind turbulence and wave excitation frequencies usually fall within this zone (LeBlance,
600 Offshore Wind Farms

2009). For OWT structures, this invariably leads to the development of higher stresses in
the support structure and its foundation, which is an unfavourable situation in considering
fatigue life. Hence, it is important to ensure that excitation frequencies with high energy
levels do not coincide with the natural frequency of the support structure and its founda-
tion. In this regard, DNV (2011) suggests that the natural frequency should not come close
to the 1P or 3P frequency ranges, with the ‘wanted frequency’ region (referred to as the
‘softestiff’ zone) remaining away from the 1P and 3P ranges by a margin of at least
10%. Note that scour alters the natural frequency of the dynamic response in an unfavour-
able manner, particularly for OWT monopiles installed in sandy soil. The maximum scour
depth depends on the monopile diameter, flow Froude number, and soil characteristics
(van der Tempel et al., 2004). A detailed insight into the dynamic of the soilestructure
behaviour for offshore applications is given by Adhikari and Bhattacharya (2011,
2012), Bhattacharya et al. (2011, 2013), Bhattacharya and Adhikari (2011) and Lombardi
et al. (2013).
For OWT monopiles, ‘softestiff’ design necessitates high structural and dynamic
stiffness, which can be achieved by increasing the monopile diameter, or less effi-
ciently by increasing the pile wall thickness. However, larger-diameter monopiles
introduce drawbacks (Schaumann and B€ oker, 2005), including greater wave loading
and larger equipment necessary for their installation, thereby leading to an increase
in the initial capital cost of the project. Compared with monopiles, the lattice frame
of jacket support structures (Fig. 20.1(e)) provides large bending stiffness and a
more favourable mass-to-stiffness ratio, resulting in relatively high bending
Eigen-frequencies and reduced hydrodynamic excitation (De Vries et al., 2011). How-
ever, the torsional stiffness is reduced, potentially leading to dynamic problems.
Further, the fabrication of a jacket structure is complex, difficult to automate and
resource-intensive, with its installation generally requiring up to 3 days, compared
to within typically 24 h for a monopile foundation. These factors tend to make jacket
structures costly. They are designed as either ‘softestiff’ or ‘stiffestiff’ systems, with
the requirement to avoid resonance with the 3P frequency range, especially for ‘softe
stiff’ systems (Fischer, 2011). For tripods, bracing along the lower length of the central
tubular section (reduces bending-moment loading) increases overall bending stiffness
(Schaumann and B€ oker, 2005; Saleem, 2011), with typical Eigen-frequency values
ranging between those for the monopile (at the lower end of this range) and jacket sup-
port structures, considering similar rotorenacelle configurations and environmental
conditions.

20.6.2 Design procedure


An iterative procedure is usually adopted in the design, with the basic steps involved
for an OWT monopile foundation system set out in Fig. 20.4. Data on the environ-
mental conditions, wind turbine and soil stratigraphy are required. The environmental
data are used to determine the required platform and hub elevations for the OWT struc-
ture and to select initial/trial dimensions for the monopile foundation, leading to the
determination of the natural frequency of the whole structural system. Checks on reso-
nance frequency are applied, along with predictions of the anticipated rotation/
Offshore wind turbine foundations e analysis and design 601

Environmental
Determine platform Determine
data and soil
and hub elevations extreme loads
stratigraphy

Determine allowed Determine pile


natural frequency embedment length

Determine Perform buckling


Turbine
preliminary pile and fatigue
data
dimensions checks

OK?
Figure 20.4 Design of OWT support structure and its monopile foundation.

deflection and settlement responses of the proposed monopile over its design life. The
applied loads and moments are estimated for the initial/trial dimensions of the OWT
structure considering its associated wind turbine and environmental data. For instance,
the RECAL software, a MATLAB tool for offshore wind turbine modelling developed
by Cerda Salzmann (2004), can be used to estimate the horizontal (shear) force and
bending moment acting (at seabed level) on the monopile, as documented by De Vries
and van der Tempel (2007). RECAL can simulate both wind and wave time-series,
from which it calculates the loads acting on the support structure (including monopile),
as well as its dynamic behaviour. For monopile foundations, torsional moments are
usually only a small fraction of the bending moment (Lensy and Wiemann, 2005);
eg, see typical loading values listed in Table 20.1 for a 5-MW wind turbine. Hence,
provided the checks on lateral forces and bending moments are satisfied, torsional mo-
ments are generally not critical. Fatigue and buckling checks are performed at a more
advanced stage in the design process. A discussion on numerical modelling, consid-
ering dynamic soil constitutive models, torsional loading and damping-related issues,
is beyond the scope of this chapter. Details on these topics can be found in Basack and
Dey (2012), Basack and Sen (2014), Guo (2006, 2013) and Rani and Prashant (2015).

20.6.3 Monopile embedment length and foundation stability


The monopile embedment length must be sufficient to ensure vertical and lateral sta-
bility. Studying the interaction effects for piles under combined axial and lateral
loading would no doubt call for a systematic and sophisticated analysis, although
the pertinent literature is very limited. Numerical analysis by Karthigeyan et al.
(2007) indicates that for sandy soils, the presence of axial loads increases the pile’s
lateral load-carrying capacity by as much as 40% (depending on the magnitude of axial
loading), but causes marginal reductions for clayey soils. According to current practice
(Karthigeyan et al., 2006; Moayed et al., 2012; Rahim and Stevens, 2013), for
602 Offshore Wind Farms

monopile design, separate analyses are performed that consider (i) the axial loading
only, to determine the bearing capacity and settlement response, and (ii) the lateral
loading only, to determine its ultimate lateral load-carrying capacity and flexural
behaviour through cantilever action. The input data for the monopile design include
soil strength and stiffness profiles against depth, pile characteristics (its
cross-sectional dimensions and properties, material strength and stiffness), and the
design loads and moments (De Vries, 2007; Jaimes, 2010; Tong, 2010).

20.6.3.1 Axial pile stability


The ultimate axial load-carrying capacity (under static loading) (Qd) of the monopile
can be determined using Eq. [20.1], which is the recommended practice by API (2010),
and has been employed by many researchers; eg, De Vries (2007), Igoe et al. (2010),
Haiderali et al. (2013) and Bisoi and Haldar (2014).

Qd ¼ Qf þ Qb ¼ fA þ qAb [20.1]

where Qf is the shaft friction capacity, Qb is the end-bearing capacity, f is the unit skin
friction, A is the shaft area pertaining to the embedded monopile length, q is the unit
end-bearing capacity, and Ab is the base area of the monopile.
For monopiles with a diameter of 4 m or greater, the pile plug resistance is usually
not taken into account (van der Tempel, 2006). Further, it has been shown that degra-
dation of the shaft friction capacity due to cyclic axial loading (Gavin and O’Kelly,
2007; Igoe et al., 2011) can lead to accumulating displacements and may ultimately
cause a severe reduction in the axial load-carrying capacity. Consideration of this
degradation of shaft friction in the design process is still an open question. However,
no reduction in the axial load-carrying capacity is to be expected if a certain magnitude
of cyclic load amplitude is not exceeded; for further details, see Poulos (1988) and
Abdel-Rahman and Achmus (2011). Numerous design charts available in the literature
make it possible to distinguish between stable and unstable loading levels to ensure a
design solution on the safe side.

20.6.3.2 Lateral pile stability


In current practice, the ultimate lateral load-carrying capacity of the monopile is
determined using the conventional pey curves (API, 2010; DNV, 2011), which
are dependent on the soil type and loading conditions. For serviceability, compared
with the axial loading case, the cyclic lateral loads are considered governing, as
described in several design codes (eg, API (2010); DNV (2011)) and documented
by many researchers (Achmus, 2010; Leblanc et al., 2010; Malhotra, 2010; Peng
et al., 2011; Bhattacharya et al., 2013; Kuo et al., 2012; Zhu et al., 2013; Haiderali
et al., 2013; Lombardi et al., 2013; Nicolai and Ibsen, 2014; Carswell et al., 2015).
Hence, the pile size and embedment length necessary to satisfy the lateral load re-
quirements are generally greater than those necessary for axial loading (Kopp,
2010). For OWT foundations, the monopile must mobilise sufficient soil resistance
Offshore wind turbine foundations e analysis and design 603

over its embedded length to transfer all the different types of applied loads to the sur-
rounding soil, with adequate safety margins provided, and also to prevent toe ‘kick’
(displacement at the pile base and excessive lateral deflection/rotation of the pile
from occurring. Checks on the values of lateral deflection (occurring at seabed
level) and rotation of the monopile are applied and its embedment length optimized
accordingly. The lateral deflection and rotation anticipated for the design loads
are calculated by considering the closed-form solutions proposed, for example, by
Randolph (1981), Broms (1964) or Matlock and Reese (1960). These methods are
related to monotonic (static) loading conditions and hence there is no consideration
of the number of lateral load cycles.

20.7 Future outlook and research needs


Reductions in the levelized cost of energy and projected increases in the design life
of OWT projects can be achieved, in part, by developing and implementing
improved design codes for the OWT support structures and foundations (as well
as for the wind turbines themselves), along with the use of innovative materials
in their fabrication. Sources, types and methods of analyses for the determination
of the magnitudes of the environmental loads and moments exerted on OWT foun-
dations, including for long-term and extreme conditions, are well documented in the
literature. However, the design of large-diameter (rigid) monopile foundations for
OWT structures is well outside the scope of current experience and analysis/design
methods (including API (2010) and DNV (2011) recommended practice), which are
mainly based on experimental data obtained for smaller-diameter (flexible) piles
that were subjected to low numbers of load cycles (<200). Further, the response
of OWT monopiles to in-service cyclic (dynamic) loading scenarios and their
long-term low-amplitude cyclic lateral-load behaviour are not well documented/un-
derstood. There are also conflicting views among researchers regarding the severity
of lateral strain accumulation for monopiles under different loading scenarios. The
scarcity of field data reported in the literature for cyclic lateral loading of
large-diameter monopiles makes the validation of current and improved design
methods, and also the calibration of numerical models for offshore monopiles, diffi-
cult. In-depth experimental and numerical studies are necessary to bridge the
knowledge gap between the existing design codes/guidelines (primarily developed
for the offshore oil/gas industries) and the more onerous loading conditions and
larger monopile foundations required for OWTs. Instrumented field testing of
full-scale monopiles (having comparable size (slenderness ratio) to that of current
and proposed OWT monopile foundations), with the application of high numbers
of lateral load cycles (>106), combined with more extensive programmes of testing
on reduced-scale piles for different soil conditions, is warranted. Such studies
would provide valuable information for the validation of current and improved
methods/theories for the analysis and design of OWT monopiles, as well as for
the calibration of numerical models.
604 Offshore Wind Farms

Nomenclature

Ab Base area of pile


A Shaft area pertaining to embedded pile length
D Diameter of structural element between seabed level and transition piece
Epy Soil stiffness (secant modulus of pey curve)
f Unit skin friction of pile
H Horizontal load applied at pile head
M Bending moment
p Lateral soil resistance
pt Lateral soil resistance exerted due to pile lateral deflection yt
q Unit end-bearing capacity of pile
Qd Ultimate static axial load-carrying capacity of pile
Qf Shaft friction capacity
Qb End-bearing capacity
V Axial load applied at pile head
y Lateral deflection of pile
yt Lateral deflection of pile for specific soil depth xt
l Impinging wavelength of ocean wave

Abbreviations

API American Petroleum Institute


DIN Deutsches Institut f€ur Normung
DNV Det Norske Veritas
HCA Hollow cylinder apparatus
OWT Offshore wind turbine
1P First excitation frequency range
3P Blade-passing frequency range for three-bladed turbine
Offshore wind turbine foundations e analysis and design 605

References
Abdel-Rahman, K., Achmus, M., 2011. Behavior of foundation piles for offshore wind energy
plants under axial cyclic loading. In: Proceedings of the SIMULIA Customer Conference,
17e19th May 2011, Barcelona, Spain. Dassault Systemes Press, Paris, France,
pp. 331e341.
Achmus, M., Kuo, Y.-S., Abdel-Rahman, K., 2009. Behavior of monopile foundations under
cyclic lateral load. Computers and Geotechnics 36 (5), 725e735.
Achmus, M., 2010. Design of axially and laterally loaded piles for the support of offshore wind
energy converters. In: Proceedings of the Indian Geotechnical Conference GEOtrendz,
16e18th December 2010, Mumbai, India, pp. 92e102.
Adhikari, S., Bhattacharya, S., 2011. Vibrations of wind-turbines considering soil‒structure
interaction. Wind and Structures 14 (2), 85e112.
Adhikari, S., Bhattacharya, S., 2012. Dynamic analysis of wind turbine towers on flexible
foundations. Shock and Vibration 19 (1), 37e56.
API (American Petroleum Institute), 2010. API RPA2: Recommended Practice for Planning,
Designing and Constructing Fixed Offshore Platforms e Working Stress Design,
twenty-second ed. API, Washington, DC.
Arshad, M., O’Kelly, B.C., 2013. Offshore wind-turbine structures: a review. Proceedings of the
Institution of Civil Engineers, Energy 166 (4), 139e152.
Arshad, M., O’Kelly, B.C., 2014. Development of a rig to study model pile behaviour under
repeating lateral loads. International Journal of Physical Modelling in Geotechnics 14 (3),
54e67.
Atkinson, J.H., Sallfors, G., 1991. Experimental determination of stressestrainetime charac-
teristics in laboratory and in-situ tests. In: Proceedings of the 10th European Conference on
Soil Mechanics and Foundation Engineering, 26e30th May 1991, vol. 3, pp. 915e956.
Florence, Italy. Balkema: Rotterdam, The Netherlands.
Babcock and Brown Company, 2012. The Future of Offshore Wind Energy. See. http://
bluewaterwind.com (accessed 20.08.12.).
Basack, S., Dey, S., 2012. Influence of relative pile-soil stiffness and load eccentricity on single
pile response in sand under lateral cyclic loading. Geotechnical and Geological Engineering
30 (4), 737e751.
Basack, S., Sen, S., 2014. Numerical solution of single piles subjected to pure torsion.
Geotechnical and Geoenvironmental Engineering 140 (1), 74e90.
Batchelor, G.K., 1967. An Introduction to Fluid Dynamics. Cambridge University Press,
Cambridge, UK.
Bhattacharya, S., Adhikari, S., 2011. Experimental validation of soil-structure interaction of
offshore wind turbines. Soil Dynamics and Earthquake Engineering 31 (5e6), 805e816.
Bhattacharya, S., Lombardi, D., Muir Wood, D., 2011. Similitude relationships for physical
modelling of monopile-supported offshore wind turbines. International Journal of Physical
Modelling in Geotechnics 11 (2), 58e68.
Bhattacharya, S., Cox, J., Lombardi, D., Muir Wood, D., 2013. Dynamics of offshore wind
turbines supported on two foundations. Proceedings of the Institution of Civil Engineers,
Geotechnical Engineering 166 (2), 159e169.
Bienen, B., D€uhrkop, J., Grabe, J., Randolph, M.F., White, D., 2012. Response of piles with
wings to monotonic and cyclic lateral loading in sand. Geotechnical and Geoenvironmental
Engineering 138 (3), 364e375.
606 Offshore Wind Farms

Bisoi, S., Haldar, S., 2014. Dynamic analysis of offshore wind turbine in clay considering
soilemonopileetower interaction. Soil Dynamics and Earthquake Engineering 63 (2014),
19e35.
Blanco, M.I., 2009. The economics of wind energy. Renewable and Sustainable Energy Re-
views 13 (6e7), 1372e1382.
Broms, B., 1964. Lateral resistance of piles in cohesionless soils. Soil Mechanics and Foun-
dation Engineering Division, ASCE 90 (3), 123e156.
Byrne, B.W., Houlsby, G.T., 2003. Foundations for offshore wind turbines. Philosophical
Transactions of the Royal Society of London 361 (1813), 2909e2930.
Carswell, W., Arwade, S.R., DeGroot, D.J., Lackner, M.A., 2015. Soilestructure reliability of
offshore wind turbine monopile foundations. Wind Energy 18 (3), 483e498.
Cerda Salzmann, D.J., 2004. Dynamic Response Calculations of Offshore Wind Turbine
Monopile Support Structures (M.Sc. thesis). Delft University of Technology, Delft, The
Netherlands.
Cuéllar, P., Georgi, S., Baeßler, M., Rucker, W., 2012. On the quasi-static granular convective
flow and sand densification around pile foundations under cyclic lateral loading. Granular
Matter 14 (1), 11e25.
Wang, S.T., Reese, L.C., 1993. COM624PdLaterally Loaded Pile Analysis Program for the
Microcomputer, Version 2.0. Final report FHWA-SA-91-048. Federal Highway Admin-
istration, Washington, DC.
Das, B.M., 2008. Advanced Soil Mechanics. CRC Press, New York, NY.
De Vries, W.E., 2007. Project UpWind WP4 Deliverable D4.2.1: Assessment of
Bottom-Mounted Support Structure Types with Conventional Design Stiffness and
Installation Techniques for Typical Deep Water Sites. See. http://www.upwind.eu/
Publications/w/media/UpWind/Documents/Publications/4%20-%20Offshore%20Foundations/
UpwindWP4D421%20Assessment%20of%20bottommounted%20support%20structure%
20types.ashx (accessed 15.03.15.).
De Vries, W.E., van der Tempel, J., 2007. Quick monopile design. In: Proceedings of the Eu-
ropean Offshore Wind Conference and Exhibition, Berlin, Germany, 4e6th December
2007.
De Vries, W.E., Vemula, N.K., Passon, P., Fischer, T., Kaufer, D., Matha, D., Schmidt, B.,
Vorpahl, F., 2011. Support Structure Concepts for Deep Water Sites. Final report WP 4.2.
UpWind Project: Offshore Foundations and Support Structures Deliverable D4.2.8. See.
http://repository.tudelft.nl/view/ir/uuid%3A4d009175-4cc4-4a90-881c-9ffb056b0806/
(accessed 30.05.15.).
DIN (Deutsches Institut f€ur Normung), 2005. DIN 1054: Baugrund Sicherheitsnachweise im Erd
und Grundbau. DIN, Berlin, Germany. See. http://www.baunormenlexikon.de/Normen/
DIN/DIN%201054/1b69b621-74a0-4c33-8b29-a82a24afd4d4 (accessed 15.03.15.).
DNV (Det Norske Veritas), 2011. DNV-OS-J101: Design of Offshore Wind Turbine Structures.
DNV, Oslo, Norway.
Dobry, R., Vicenti, E., O’Rourke, M.J., Roesset, J.M., 1982. Horizontal stiffness and damping
of single piles. Journal of Geotechnical Engineering Division, ASCE 108 (3), 439e459.
Doherty, P., Kirwan, L., Gavin, K., Igoe, D., Tyrrell, S., Ward, D., O’Kelly, B., 2012. Soil
properties at the UCD geotechnical research site at Blessington. In: Proceedings of the
Bridge and Concrete Research in Ireland Conference (BCRI 2012), Dublin, Ireland, 6e7th
September 2012, vol. 1, pp. 499e504. http://www.tara.tcd.ie/handle/2262/67119.
D€
uhrkop, J., 2009. On the Influence of Expanders and Cyclic Loads on the Deformation
Behavior of Lateral Stressed Piles in Sand (Ph.D. thesis). Hamburg University of Tech-
nology, Hamburg, Germany.
Offshore wind turbine foundations e analysis and design 607

El Naggar, M.H., Wei, J.Q., 1999. Response of tapered piles subjected to lateral loading.
Canadian Geotechnical Journal 36 (1), 52e71.
Esteban, M., Lopes-Gutierrez, J., Diez, J., Negro, V., 2011. Foundations for offshore wind
farms. In: Proceedings of the 12th International Conference on Environmental Science and
Technology, Rhodes, Greece, pp. 516e523.
Fischer, T., 2011. Executive Summary e Upwind Project WP4: Offshore Foundations and
Support Structures. See. http://www.upwind.eu/Publications/w/media/UpWind/Documents/
Publications/4%20-%20Offshore%20Foundations/WP4_Executive_Summary_Final.ashx
(accessed 15.03.15.).
Gavin, K.G., O’Kelly, B.C., 2007. Effect of friction fatigue on pile capacity in dense sand.
Geotechnical and Geoenvironmental Engineering 133 (1), 63e71.
GL (Germanischer Lloyd WindEnergie GmbH), 2005. Rules and Guidelines, IV Industrial Services,
Guideline for the Certification of Offshore Wind Turbines, Ch. 2 and 6. See. http://onlinepubs.
trb.org/onlinepubs/mb/Offshore%20Wind/Guideline.pdf (accessed 17.09.15.).
Guo, W.D., 2006. On limiting force profile, slip depth and response of lateral piles. Computer
and Geotechnics 33 (1), 47e67.
Guo, W.D., 2013. Pu-based solutions for slope stabilizing piles. International Journal of Geo-
mechanics 13 (3), 292e310.
Haiderali, A., Cilingir, U., Madabhushi, G., 2013. Lateral and axial capacity of monopiles for
offshore wind turbines. Indian Geotechnical Journal 43 (3), 181e194.
Haritos, N., 2007. Introduction to the analysis and design of offshore structuresdan overview.
Electronic Journal of Structural Engineering 2007, 55e65.
IEC (International Electrotechnical Commission), 2005. IEC 61400e1: Wind Turbines d Part 1:
Design requirements, third ed. IEC, Geneva, Switzerland.
Igoe, D.J.P., Gavin, K.G., O’Kelly, B.C., 2011. Shaft capacity of open-ended piles in sand.
Geotechnical and Geoenvironmental Engineering 137 (10), 903e913.
Igoe, D., Gavin, K., O’Kelly, B., 2013a. An investigation into the use of push-in pile foundations
by the offshore wind sector. International Journal of Environmental Studies 70 (5),
777e791.
Igoe, D.J.P., Gavin, K.G., O’Kelly, B.C., Byrne, B., 2013b. The use of in-situ site investigation
techniques for the axial design of offshore piles. In: Coutinho, R.Q., Mayne, P.W. (Eds.),
Proceedings of the Fourth International Conference on Geotechnical and Geophysical Site
Characterization (ISC’4), 18the21st September 2012. Pernambuco, Brazil, vol. 2,
pp. 1123e1129.
Igoe, D., Gavin, K., O’Kelly, B., 2010. Field tests using an instrumented model pipe pile in sand.
In: Springman, S., Laue, J., Seward, L. (Eds.), Proceedings of the Seventh International
Conference on Physical Modelling in Geotechnics, Zurich, Switzerland, 28th Junee1st
July 2010, vol. 2. CRC Press/Balkema, Leiden, The Netherlands, pp. 775e780.
Irvine, J.H., Allan, P.G., Clarke, B.G., Peng, J.R., 2003. Improving the lateral stability of
monopile foundations. In: Newson, T.A. (Ed.), Proceedings of the International Conference
on Foundations: Innovations, Observations, Design and Practice. 2nde5th September
2003, Dundee, UK. Thomas Telford, London, pp. 371e380.
Jaimes, O.G., 2010. Design Concepts for Offshore Wind Turbines: A Technical and Economic
Study on the Trade-off between Stall and Pitch Controlled Systems (M.Sc. thesis). Delft
University of Technology, Delft, The Netherlands.
Journée, J.M.J., Massie, W.W., 2001. Offshore Hydrodynamics, first ed. Delft University of
Technology, Delft, The Netherlands.
Junginger, M., Agterbosch, S., Faaij, A., Turkenburg, W., 2004. Renewable electricity in the
Netherlands. Energy Policy 32 (9), 1053e1073.
608 Offshore Wind Farms

Karg, C., 2007. Modelling of Strain Accumulation Due to Low Level Vibrations in Granular
Soils (Ph.D. thesis). Ghent University, Ghent, Belgium.
Karthigeyan, S., Ramakrishna, V.V.G.S.T., Rajagopal, K., 2006. Influence of vertical load on
the lateral response of piles in sand. Computers and Geotechnics 33 (2), 121e131.
Karthigeyan, S., Ramakrishna, V.V.G.S.T., Rajagopal, K., 2007. Numerical investigation of the
effect of vertical load on the lateral response of piles. Geotechnical and Geoenvironmental
Engineering 133 (5), 512e521.
Klinkvort, R.T., Hededal, O., 2013. Lateral response of monopile supporting an offshore wind
turbine. Proceedings of the Institution of Civil Engineers, Geotechnical Engineering
166 (2), 147e158.
Kopp, D.R., 2010. Foundations for an Offshore Wind Turbine (M.Sc. thesis). Massachusetts
Institute of Technology, Cambridge, MA.
Kuo, Y.S., Achmus, M., Abdel-Rahman, K., 2012. Minimum embedded length of cyclic horizon-
tally loaded monopiles. Geotechnical and Geoenvironmental Engineering 138 (3), 357e363.
Lai, S., 1989. Similitude for shaking table test on soil-structure-fluid model in 1-g gravitational
field. Soils and Foundations 29 (1), 105e118.
LeBlanc, C., 2009. Design of Offshore Wind Turbine Support Structures: Selected Topics in the
Field of Geotechnical Engineering (Ph.D. thesis). Aalborg University, Aalborg, Denmark.
LeBlanc, C., Houlsby, G.T., Byrne, B.W., 2010. Response of stiff piles in sand to long-term
cyclic lateral loading. Géotechnique 60 (2), 79e90.
Lesny, K., Wiemann, J., 2005. Design aspects of monopiles in German offshore wind farms. In:
Gourvenec, S., Cassidy, M. (Eds.), Proceedings of the First International Symposium on
Frontiers in Offshore Geotechnics. Perth, Australia, 19the21st September 2005. Balkema,
Leiden, The Netherlands, pp. 383e389.
Li, D., Haigh, S.K., Bolton, M.D., 2010. Centrifuge modelling of mono-pile under cyclic lateral
loads. In: Springman, S., Laue, J., Seward, L. (Eds.), Proceedings of the Seventh Inter-
national Conference on Physical Modelling in Geotechnics, 28th Junee1st July 2010,
Zurich, Switzerland, vol. 2. CRC Press, Leiden, The Netherlands, pp. 965e970.
Lin, S.-S., Liao, J.-C., 1999. Permanent strains of piles in sand due to cyclic lateral loads.
Geotechnical and Geoenvironmental Engineering 125 (9), 798e802.
Little, R.L., Briaud, J.L., 1988. Full Scale Cyclic Lateral Load Tests on Six Single Piles in Sand.
Miscellaneous paper GL-88e27. Geotechnical Division, Civil Engineering Department,
Texas A&M University, College Station, TX.
Lombardi, D., Bhattacharya, S., Muir Wood, D., 2013. Dynamic soilestructure interaction of
monopile supported wind turbines in cohesive soil. Soil Dynamics and Earthquake
Engineering 49 (2013), 165e180.
Long, J., Vanneste, G., 1994. Effects of cyclic lateral loads on piles in sand. Journal of
Geotechnical Engineering, ASCE 120 (1), 225e244.
Malhotra, S., 2010. Design and construction considerations for offshore wind turbine founda-
tions in North America. In: Fratta, D.O., Puppala, A.J., Muhunthan, B. (Eds.), Proceedings
of GeoFlorida 2010: Advances in Analysis, Modeling & Design, Orlando, Florida,
20e24th February 2010, GSP 199, pp. 1533e1542.
Matlock, H., Reese, L.C., 1960. Generalized solutions for laterally loaded piles. Journal of the
Soil Mechanics and Foundations Division ASCE 86 (5), 63e94.
Moayed, R.Z., Mehdipour, I., Judi, A., 2012. Undrained lateral behaviour of short pile under
combination of axial, lateral and moment loading in clayey soils. Kuwait Journal of Science
and Engineering 39 (1B), 59e78.
Morison, J.R., Johnson, J.W., Schaff, S.A., 1950. The forces exerted by surface waves on piles.
Journal of Petroleum Technology 2 (5), 149e154.
Offshore wind turbine foundations e analysis and design 609

Muir Wood, D., Crewe, A.J., Taylor, C.A., 2002. Shaking table testing of geotechnical models.
International Journal of Physical Modelling in Geotechnics 2 (1), 1e13.
Naughton, P.J., O’Kelly, B.C., 2004. The induced anisotropy of Leighton Buzzard sand. In:
Jardine, R.J., Potts, D.M., Higgins, K.G. (Eds.), Proceedings of Advances in Geotechnical
Engineering: The Skempton Conference, 28the31st March 2004, London, UK, vol. 1.
Thomas Telford, London, UK, pp. 556e567.
Naughton, P.J., O’Kelly, B.C., 2005. Yield behaviour of sand under generalized stress condi-
tions. In: Proceedings of the 16th International Conference on Soil Mechanics and
Geotechnical Engineering, 12e16th September 2005, vol. 2. IOS Press, Osaka, Japan,
pp. 555e558.
Nicolai, G., Ibsen, L.B., 2014. Small-scale testing of cyclic laterally loaded monopiles in dense
saturated sand. Journal of Ocean and Wind Energy 1 (4), 240e245.
Niemunis, A., Wichtmann, T., Triantafyllidis, T., 2005. A high-cycle accumulation model for
sand. Computers and Geotechnics 32 (4), 245e263.
OSPAR Commission, 2014. Inventory of Measures to Mitigate the Emission and Environmental
Impact of Underwater Noise. In: Biodiversity Series. OSPAR Commission, London, UK.
See. http://www.ospar.org/documents/dbase/publications/p00626/p00626_inventory_of_
noise_mitigation.pdf (accessed 03.06.15.).
O’Kelly, B.C., Naughton, P.J., 2005. Development of a new hollow cylinder apparatus for stress
path measurements over a wide strain range. Geotechnical Testing Journal 28 (4),
345e354.
O’Kelly, B.C., 2006. Compression and consolidation anisotropy of some soft soils. Geotech-
nical and Geological Engineering 24 (6), 1715e1728.
O’Kelly, B.C., Naughton, P.J., 2008. Local measurements of the polar deformation response in a
hollow cylinder apparatus. Geomechanics and Geoengineering 3 (4), 217e229.
O’Kelly, B.C., Naughton, P.J., 2009. Study of the yielding of sand under generalized stress
conditions using a versatile hollow cylinder torsional apparatus. Mechanics of Materials
41 (3), 187e198.
Pappusetty, D., Pando, M.A., 2013. Numerical evaluation of long term monopile head behaviour
for ocean energy converters under sustained low amplitude lateral loading. International
Journal of Civil and Structural Engineering 3 (4), 669e684.
Peng, J., Clarke, B., Rouainia, M., 2006. A device to cyclic lateral loaded model piles.
Geotechnical Testing Journal 29 (4), 1e7.
Peng, J.-R., Rouainia, M., Clarke, B.G., 2010. Finite element analysis of laterally loaded fin
piles. Computers and Structures 88 (21e22), 1239e1247.
Peng, J., Clarke, B., Rouainia, M., 2011. Increasing the resistance of piles subject to cyclic
lateral loading. Geotechnical and Geoenvironmental Engineering 137 (10), 977e982.
Peralta, P., Achmus, M., 2010. An experimental investigation of piles in sand subjected to lateral
cyclic loads. In: Springman, S., Laue, J., Seward, L. (Eds.), Proceedings of the Seventh
International Conference on Physical Modelling in Geotechnics, 28th Junee1st July 2010,
Zurich, Switzerland, vol. 2. CRC Press, Leiden, The Netherlands, pp. 985e990.
Poulos, H.G., 1988. Cyclic stability diagram for axially loaded piles. Journal of Geotechnical
Engineering, ASCE 114 (8), 877e895.
Rahim, A., Stevens, R.F., 2013. Design procedures for marine renewable energy foundations.
In: Proceedings of the First Marine Energy Technology Symposium, 10e11th April 2013,
Washington, DC. 10 p.
Ramakrishna, V.G.S.T., Rao, S.N., 1999. Critical cyclic load levels for laterally loaded piles in
soft clays. In: Sing, S.K., Lacasse, S. (Eds.), Proceedings of the International Conference on
Offshore and Nearshore Geotechnical Engineering, Panvil, Mumbai, India, pp. 301e307.
610 Offshore Wind Farms

Randolph, M.F., 1981. The response of flexible piles to lateral loading. Géotechnique 31 (2),
247e259.
Rani, S., Prashant, A., 2015. Estimation of the linear spring constant for a laterally loaded
monopile embedded in nonlinear soil. International Journal of Geomechanics 15 (6). http://
dx.doi.org/10.1061/(ASCE)GM.1943-5622.0000441.
Rosquoet, F., Thorel, L., Garnier, J., Canepa, Y., 2007. Lateral cyclic loading of sand-installed
piles. Soils and Foundations 47 (5), 821e832.
Şahin, A.D., 2004. Progress and recent trends in wind energy. Progress in Energy and Com-
bustion Science 30 (5), 501e543.
Saleem, Z., 2011. Alternatives and Modifications of Monopile Foundation or its Installation
Technique for Noise Mitigation. Delft University of Technology, Delft, The Netherlands.
See. http://www.vliz.be/imisdocs/publications/223688.pdf (accessed 03.08.15.).
Schaumann, P., B€oker, C., 2005. Can tripods and jackets compete with monopiles?. In: Pro-
ceedings of the European Offshore Wind Conference and Exhibition, 26e28th October 2005,
Copenhagen, Denmark, 10 p. See. http://wind.nrel.gov/public/SeaCon/Proceedings/
Copenhagen.Offshore.Wind.2005/documents/papers/Low_cost_foundations/P.Schaumann_
Can_jackets_and_tripods_compete_with_monopile.pdf (accessed 30.06.15.).
van der Tempel, J., Molenaar, D.-P., 2002. Wind turbine structural dynamics d a review of the
principles for modern power generation, onshore and offshore. Wind Engineering 26 (2),
211e220.
Tomlinson, M.J., 2001. Foundation Design and Construction, seventh ed. Pearson, Harlow,
Essex, UK.
Tong, W., 2010. Wind Power Generation and Wind Turbine Design. WIT Press, Southampton,
UK.
van der Tempel, J., 2006. Design of Support Structures for Offshore Wind Turbines (Ph.D.
thesis). Delft University of Technology, Delft, The Netherlands.
van der Tempel, J., Zaaijer, M.B., Subroto, H., 2004. The effects of scour on the design of
offshore wind turbines. In: Proceedings of the Third International Conference on Marine
Renewable Energy, 7e9th July 2004, Blyth, UK. Institute of Marine Engineering, Science,
and Technology, London, UK, pp. 27e35.
Verdure, L., Garnier, J., Levacher, D., 2003. Lateral cyclic loading of single piles in sand. In-
ternational Journal of Physical Modelling in Geotechnics 3 (3), 17e28.
Wichtmann, T., Rondon, H.A., Niemunis, A., Triantafyllidis, T., Lizcano, A., 2010. Prediction
of permanent deformations in pavements using a high-cycle accumulation model.
Geotechnical and Geoenvironmental Engineering 136 (5), 728e740.
Zhu, B., Byrne, B.W., Houlsby, G.T., 2013. Long-term lateral cyclic response of suction caisson
foundations in sand. Geotechnical and Geoenvironmental Engineering 139 (1), 73e83.

You might also like