You are on page 1of 23

Marine Georesources & Geotechnology, 0: 1–23

Copyright # 2015 Taylor & Francis Group, LLC


ISSN: 1064-119X print/1521-0618 online
DOI: 10.1080/1064119X.2015.1033070

Analysis and Design of Monopile Foundations for Offshore


Wind-Turbine Structures
MUHAMMAD ARSHAD1,2 and BRENDAN C. O’KELLY1
1
Department of Civil, Structural and Environmental Engineering, Trinity College Dublin, Dublin, Ireland
2
Department of Geological Engineering, University of Engineering & Technology, Lahore, Pakistan
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

Received 10 December 2012, Accepted 19 March 2015

Offshore wind turbines (OWTs) are generally supported by large-diameter monopiles, with the combination of axial forces, lateral
forces, bending moments, and torsional moments generated by the OWT structure and various environmental factors resisted by
earth pressures mobilized in the soil foundation. The lateral loading on the monopile foundation is essentially cyclic in nature and
typically of low amplitude. This state-of-the-art review paper presents details on the geometric design, nominal size, and structural
and environmental loading for existing and planned OWT structures supported by monopile foundations. Pertinent ocean-environ-
ment loading conditions, including methods of calculation using site-specific data, are described along with wave particle kinemat-
ics, focusing on correlations between the loading frequency and natural vibration frequency of the OWT structure. Existing
methods for modeling soil under cyclic loading are reviewed, focusing in particular on strain accumulation models that consider
pile–soil interaction under cyclic lateral loading. Inherent limitations=shortcomings of these models for the analysis and design
of existing and planned OWT monopile foundations are discussed. A design example of an OWT support structure having a mono-
pile foundation system is presented. Target areas for further research by the wind-energy sector, which would facilitate the devel-
opment of improved analyses=design methods for offshore monopiles, are identified.
Keywords: foundation, lateral load, monopile, ocean environment, soil, strain accumulation

Introduction design, the proportion and importance of the different load-


ing types essentially depend upon the kind of foundation
There has been a rapid growth in the use of offshore (and system being considered. For gravity-base foundations
onshore) wind farms for the production of clean and renew- (Figure 1a), potential failure modes may be in bearing
able energy. The economic development of wind farms capacity or excess settlement; hence, the vertical load is gen-
depends on efficient solutions being available for a number erally the major design consideration (Malhotra 2011). For
of technical issues, one aspect being the foundations. Off- monopile foundations (Figure 1b), lateral deflection
shore wind-turbine (OWT) structures may be found on grav- (rotation) of the monopile controls the serviceability limit
ity base, suction caisson, monopile, tripod or braced frame state of the whole structure; hence, lateral loads are more
(jacket) foundations or, more recently, using floating plat- critical when compared with the vertical loads.
forms tethered to the seabed (Figure 1). The foundation In general, offshore foundations must be designed to
choice largely depends on the water depth, seabed character- resist large numbers of aerodynamic and hydrodynamic load
istics, loading characteristics, available construction technol- cycles of varying direction, amplitude, and frequency
ogies, and importantly, economic costs (Malhotra 2011). (Figure 2) at the proposed site over the project’s lifetime of
Offshore foundations are subjected to a combination of typically 25 years or more (Şahin 2004). In most parts
loads, namely the axial (self-weight) forces of the of the world, the most popular foundation choice in terms
structure=machinery (V), repeating horizontal=lateral loads of ease of installation, economy, and logistics is the mono-
(H), and bending (M) and torsional moments. Apart from pile, with an estimated 75% of all installed OWTs using this
the self-weight, these loads are generated by environmental solution (Blanco 2009; Fischer 2011; Malhotra 2011). For
conditions (ocean waves, currents, tidal action, and aerody- OWT monopile foundation systems, the lateral loads (and
namic load cycles) and=or operation of the installed machin- resulting moments) are generally large in proportion to the
ery (Figure 2). The lateral loading is essentially cyclic in axial loads. Hence, the foundation response under repeating
nature and typically of low amplitude. For geotechnical lateral loading is a major consideration, with the foundation
design dominated by considerations of the dynamic and fati-
Address correspondence to Muhammad Arshad, Department gue responses under working loads, rather than the ultimate
of Geological Engineering, University of Engineering & Tech- load-carrying capacity. The overall feasibility of a wind-farm
nology, GT Road Lahore, Pakistan. E-mail: arshadm@tcd.ie project is invariably determined by upfront capital costs,
2 M. Arshad & B. C. O’Kelly
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

Fig. 2. Environmental impacts on offshore wind turbine.


Fig. 1. Support structure options showing their range of appli-
cable water depths.
generation capacity, which is an indirect measure of the
with the material, fabrication, and installation costs for applied loading. Other significant influencing factors include
OWT foundations, typically accounting for 25% of the the soil condition and severity of environmental loading.
overall project cost (Blanco 2009). This significant invest- Existing offshore monopiles are typically 5–7.5 m in diam-
ment demands reliable foundation designs to achieve the eter (Achmus, Kuo, and Abdel-Rahman 2009), with slender-
overall goals of pollution-free wind-generated power at ness [embedded length (L) to outer diameter (D)] ratio
economical rates. values of <10, and are; therefore, considered to behave as
The aims of this review article are to present: rigid structures for which rotation is prominent over bending
(Tomlinson 2001; Peng, Clarke, and Rouainia 2011). For a
. Details on the geometric design and structural and typical OWT having a rated power of 5 MW and installed
environmental loading for existing and planned OWTs in the North Sea, a hub height (see Figure 3) of 95 m above
supported by monopile foundations. the mean sea level and rotor diameter of 125 m lead to the
. Describe ocean-environment loading conditions, including approximate quasi-static loading scenario summarized in
methods of calculation using site-specific data. Table 1 (Lesny and Wiemann 2005; IEC [International
. Discuss wave particle kinematics, focusing on correlations Electrotechnical Commission] 2009) and which acts at the
between the loading frequency and natural vibration fre- seabed level. This example demonstrates that torsional
quency of OWT structures. moments are negligibly small, whereas lateral loading causes
. Critically review existing semi-empirical methods for mod- extremely high bending moments and principally controls
eling soil under cyclic loading, focusing on cyclic design the foundation design.
for fatigue life. The inherent limitations=shortcomings of The monopile foundation transfers the horizontal forces
these models for the analyses=design of existing and and bending moments generated by wind and wave loading
planned OWT monopile foundations are identified along (via horizontal earth pressure acting along the pile) into the
with further research needs. surrounding ground through cantilever action. Factors
. Present a typical example of the geometric design for an affecting the cyclic response include the monopile size and
OWT-monopile foundation system. embedment length, soil properties, soil–pile relative stiffness,
Other aspects related to offshore monopile structures loading characteristics, and pile installation method (Malho-
(e.g., construction materials, manufacturing advancements, tra 2011). Depending on subsoil conditions, the monopile is
corrosion assessment, and control) are beyond the scope of typically installed into the seabed by drilling and grouting,
this article. driving using large impact or vibratory hammers, or a com-
bination of drilling and driving (Malhotra 2011).
In current practice, OWT monopile foundations are
General Aspects of OWT-Monopile Foundation usually designed using general geotechnical standards in
System combination with more specific guidelines and semi-
empirical formulas that have been largely developed by the
A monopile is a single large-diameter steel tube that pene- offshore oil=gas industry. The latter are usually based on
trates into the seabed. The required monopile length and limited field data obtained for piles having significantly
diameter are primarily dependent on the OWT’s power smaller diameters (DIN [Deutsches Institut fur Normung]
Analysis and Design of Offshore Monopile Foundations 3

2.0 and 1.4 MN, respectively: these scenarios correspond to


the serviceability and fatigue limit states respectively (GL
2005). Further, the strategy of predictions and in-situ
measurement of the behavior of every OWT monopile foun-
dation in a wind farm project is technically and economically
laborious and; hence, not practical. The application of fast
countermeasures to mitigate against unexpected conditions
developing into potentially alarming situations also cannot
be guaranteed (Richwien, Lesny, and Wiemann 2002).
Many constitutive models have been proposed to estimate
the soil response under large numbers of load cycles, although
very few of these specifically consider simulation of the soil–
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

pile interaction, instead focusing on the soil’s strength and


stiffness responses. The most popular method of analysis for
predicting the pile deflection under lateral loading, and the
method recommended in offshore design codes including
API (2010) and DNV (2011), are based on the Winkler model;
commonly referred to as the p-y method. Although it has a
long history of use in design practice, the original database
on which this method is based largely consists of a single set
of pile tests, involving limited numbers of lateral load cycles,
performed at a medium-dense sand site. In contrast, many
millions of low-amplitude cycles of lateral loading are applied
over the operational life of OWT structures.
As a general rule, under field loading, rotation of the
monopile of up to 0.5 from its vertical alignment (Achmus,
Kuo, and Abdel-Rahman 2009; LeBlanc, Houlsby, and
Byrne 2010; DNV 2011) or lateral pile deflections (based
on practical experience) of up to 120 mm occurring at seabed
level (Malhotra 2011) are considered as limiting values for
the proper operation of the wind turbine.

Geometric Details of OWTs Supported by Monopile


Foundations
The main components of the wind-turbine system are its
foundation, support structure, transition piece, tower, rotor
Fig. 3. Main components of offshore wind turbine system. blades, and nacelle (Figure 3). Together, the support struc-
ture and foundation maintain the turbine in its proper oper-
2005; GL [Germanischer Lloyd] 2005; API [American ational position. The transition piece provides a means of
Petroleum Institute] 2010; DNV [Det Norske Veritas] correcting any misalignment of the monopile that may arise
2011). Hence, extrapolation of these formulations for during its installation. In some instances, the monopile can
present and future design of OWT monopile foundations extend from the seabed level to above the water surface level,
requires careful consideration of the applied loading and connecting with the transition piece or tower (Arshad and
the inherent limitations underlying such approaches. O’Kelly 2013). The nacelle (hub) contains the key electro-
Another significant shortcoming is that current design mechanical components of the wind turbine, including the
standards and guidelines on the serviceability of piles under gearbox and generator (Malhotra 2011). Rotor diameters
cyclic lateral loading are limited. For instance, over its ser- of existing OWTs typically range 90–120 m, producing
vice life, a typical 2 MW OWT structure is usually subjected power-generation capacities of 3–6 MW respectively (Tong
to 102 and 107 cycles of lateral loading with magnitudes of 2010). Gravitational loading typically ranges 2–8 MN. For
instance, Byrne and Houlsby (2003) reported vertical load
Table 1. Loading at seabed level for 5 MW wind turbine of 6 MN for an anticipated 3.5 MW OWT in the UK region,
supported by monopile foundation (Lesny and Wiemann 2005) with lateral loading from aerodynamic and hydrodynamic
Component Magnitude factors accounting for up to 67% of the vertical loading.
These loads will vary with the size of the installation, the
Vertical load 35 MN detailed design, and local environmental conditions. OWTs
Horizontal load 16 MN having rotor diameters of 250 m and a power-generation
Bending moment 562 MN  m capacity of 20 MW are currently in the research and develop-
Torsional moment 4 MN  m ment phase (EWEA [European Wind Energy Association]
4 M. Arshad & B. C. O’Kelly
 a
2011), creating even more onerous design loading scenarios Z
for (monopile) foundations. U z ¼ U 10 ð1Þ
10
OWT monopile foundations are typically manufactured
from steel tubular sections having outer diameters of up to ln Z=Z0
7.5 m, wall thicknesses of up to 150 mm, and embedment U z ¼ U 10 ð2Þ
ln 10=Z0
(penetration) depths of between 15 and 30 m (Achmus,
Kuo, and Abdel-Rahman 2009). Monopiles are generally where the values of exponent a and the roughness parameter,
used in shallow water depths (i.e., typically <30 m), gener- Z0, are dependent on the site location, with typical values of
ally becoming too flexible for water depths between 30 0.11 and 0.0001, respectively, for open-sea environments.
and 40 m, in which case, monopiles fitted with guy wires or The wind velocity fluctuates with time for a given height,
tripod solutions (see Figure 1) are considered as economical and also with change in height, above the sea level or ground
alternatives. For greater depths, time-consuming installation surface (Figure 4). However, a mean value (U z) can be estab-
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

and the effects of soil degradation (‘potholing’), which lished for a particular site location provided sufficient wind
occurs at seabed level around the pile, make monopile foun- speed data are available. The instantaneous wind velocity
dation solutions prohibitive (Irvine et al. 2003). Other foun- can be considered as the superposition of the turbulent
dation options, as illustrated in Figure 1, are then considered component (wz) on the mean value. Hence the total aerody-
as viable options. The serviceability limit state is largely namic drag force (FDWind) acting on an offshore structure
determined by the lateral deflection (rotation) of the mono- can be determined by (Jang and Shinn 1999; API 2010):
pile under many millions of load cycles; e.g., 107 lateral load  2  
cycles of 1.4 MN magnitude (corresponding to the fatigue FDWind ¼ 0:5qair Cdair Aw U z þ qair Cdair Aw U z wz
loading for design) are expected to occur over the service life þ 0:5qair Cdair Aw ðjwz jwz Þ ð3Þ
of OWT structures (GL 2005).

Ocean-Environment Loading on Monopile Foundation


System
In the ocean-environment, the primary sources of variable=
cyclic loading on OWTs are aerodynamic, hydrodynamic,
and (depending on the geographic setting) ice loading.
Seismic loads (considered as a special type of dynamic load-
ing) and ship impact may threaten the serviceability of the
structure or even ultimately lead to its collapse. However,
the focus of this paper is on aerodynamic and hydrodynamic
loading, since these are considered more important in assess-
ments of the OWT structure’s fatigue life.

Aerodynamic Loading
Wind conditions are important in defining, not only the
loads imposed on a wind turbine’s structural components,
but also in predicting the amount of future energy produced
as a function of wind velocity. A realistic assessment of wind
direction through statistical analysis of recorded wind data
must be based on a realistic representation of wind speed
(preferably occurring at hub height), speed frequency distri-
bution, wind shear (i.e., rate of change in wind speed with
height), turbulence intensity (i.e., standard deviation of wind
speeds sampled over a 10 min period as a function of the
mean speed), wind direction distribution, and also extreme
wind gusts with return periods of up to 100 years (DNV
2011). The mean value of the 10 min period wind speed data
measured at a reference elevation of 10 m above mean sea
level (usually determined at hub height for OWTs) is referred
to as the wind speed U 10 , from which the mean wind speed
U z for some other height, Z, above mean sea level can be
approximated using either the power law or logarithmic Fig. 4. Fluctuation of wind speed about the mean value. (a) In
law given by: time domain. (b) In three-dimensional space.
Analysis and Design of Offshore Monopile Foundations 5

where qair is the density of air, Aw is the exposed area of the


offshore structure, Cdair is an aerodynamic drag coefficient
that is dependent on the structure’s geometry and size, jwzj
is the turbulent component (with the absolute value taken
in order to make the direction of the drag force coincident
with the wind velocity).
The first term on the right-hand side of Eq. (3) represents
the mean drag force, with the second and third terms denot-
ing the drag force components associated with wind velocity
fluctuations.
Wind loads on offshore structures are seldom greater than
the hydrodynamic loads and often play a relatively minor
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

role, compared with the hydrodynamic loads. However wind


loading has a longer lever arm when considering moments
generated about the structure’s foundation. Hence, it is best
not to neglect the effects of wind loading, particularly in the
calculation of overturning (bending) moments acting on the
foundation. Wind loads on offshore structures are seldom
greater than the hydrodynamic loads. However, in the calcu-
lation of the overturning (bending) moments acting on the
foundation, wind loading has a longer lever arm compared
with the wave load interaction with the OWT structure. For
instance, Byrne and Houlsby (2003) reported that in a typical
North Sea environment, the rotor thrust reaction due to the
wind loads contributes 25% of the total horizontal load
but generates 75% of the total overturning moments.
Variations in mean wind speed are relatively small com-
pared with variations in the wave period. Fluctuations about
the mean wind speed impose aerodynamic forces, although
when considering the dynamic behavior of offshore struc-
tures, these forces are also generally insignificant compared
with the hydrodynamic forces (Journée and Massie 2001).
Hence wind speed is generally considered as steady (i.e., with- Fig. 5. Definitions of harmonic wave propagation and para-
out fluctuation), resulting in constant force(s) and; hence, meters. (a) In time domain. (b) In three-dimensional space.
constant moment(s). Dynamic analysis of offshore structures
is necessary in cases where the wind field contains energy at useful in many applications since it allows one to predict
frequencies near the offshore structure’s natural frequencies complex irregular behavior in terms of the theory of regular
of vibration (API 2010), although for monopile foundations, waves; e.g., see St. Denis and Pierson (1953).
the difference between these frequencies is usually high [see
later section on Correlation between Loading Frequency Wave Particle Kinematics
and Natural Frequency of OWT Structures]. Some statistical Wave Theory
manipulations can be applied to the wind data (e.g., refer to
Jang and Shinn 1999; Haritos 2007) for the inclusion of Many regular wave theories have been developed to capture
dynamic forces due to wind velocity turbulence. the water particle kinematics associated with ocean waves.
Each theory pertains to a certain scenario with different
degrees of complexity. Hence, they have different ranking
Hydrodynamic Loading
among the offshore engineering community (Chakrabarti
For OWT monopile foundations, the dominant hydrody- 2005). Airy wave theory is the simplest of the routinely-used
namic loads are generated by ocean waves and currents, with regular wave theories. Its key features are introduced in this
minor contributions arising from other sources such as sea paper in the context of describing its role in modeling the
level variations due to tides or swell. The main sources of character of irregular sea states. Referring to Figure 5, with
wave generation are wind and astronomical forces. Earth- the wave moving in the positive horizontal direction (x), the
quakes, submarine landslides, and shipping or other nearby form of the water surface (wave profile) can be expressed as
floating structures may also be important sources of wave a function of x and time (t) by:
generation. Irregular waves are usually observed on the
ocean surface, although these can be treated as a superpo- f ðx; tÞ ¼ fa cos ðkx  xtÞ ð4Þ
sition of many regular harmonic waves, each having its and with the wave moving in the opposite direction by:
own amplitude (na), wavelength (k), frequency (f), and direc-
tion of propagation (see Figure 5). This concept can be very f ðx; tÞ ¼ fa cos ðkx þ xtÞ ð5Þ
6 M. Arshad & B. C. O’Kelly

where k is the wave number, x is the circular wave In the time domain, the total horizontal load per unit
frequency, and fa is the maximum amplitude of the water length [F(t)] exerted on a fixed object (e.g., OWT monopile
particle measured from the mean water surface level. foundation) as a result of wave motion and currents can
The resulting water particle velocity components u(t) and be considered as the linear addition of the inertial [Finertia(t)]
v(t) in the respective x and vertical directions can be and drag [Fdrag(t)] forces given by Morison, Johnson, and
expressed by: Schaff (1950):
coshðkhw þ kzÞ F ðtÞ ¼ Finertia ðtÞ þ Fdrag ðtÞ
uðtÞ ¼ fa x cos ðkx  xtÞ ð6Þ
sinhðkhw Þ
sinhðkhw þ kzÞ p
vðtÞ ¼ fa x sin ðkx  xtÞ ð7Þ Finertia ðtÞ ¼ q CM D2  u_ ðtÞ ð10Þ
sinhðkhw Þ 4 w
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

where hw is the mean depth of the water body and z is the


elevation of the water particle above the mean surface level 1
Fdrag ðtÞ ¼ qw CD D  uðtÞ  juðtÞj ð11Þ
in the vertical direction. 2
For deep water conditions, these velocity equations
where qw is the density of sea water, D is the cylinder (i.e.,
reduce to:
monopile) diameter over the distance between the seabed
and ocean surface, and u(t) ¼ uacos(xt), ju(t)j and u_ ðtÞ are
uðtÞ ¼ fa xekz cos ðkx  xtÞ ð8Þ
the horizontal water particle velocity, its absolute value
vðtÞ ¼ fa xekz sinðkx  xtÞ ð9Þ and its time derivative (acceleration), respectively, with ua
equal to the maximum along wave water-particle velocity
From Eqs. (9) and (10), a simple interpretation is that for [discussed earlier in section on Wave theory], CM and CD
deep water conditions, the elliptical orbits of water particles are the inertial and hydrodynamic drag force coefficients,
associated with Airy wave theory reduce to circular orbits. respectively, whose values are dependent on Reynold’s num-
Different methods are available to determine the horizontal ber (Re) and Keulegan–Carpenter’s number (KC):
velocity profile with depth, including Wheeler profile stretch-
ing, extrapolation, and constant extension methods. Since ua D
the wave loads are dependent on the water particle velocity, Re ¼ ð12Þ
W
horizontal wave loads for shallow waters may be scrupu-
lously undervalued if the calculations are based on deep ua T
water equations. When considering maximum velocity KC ¼ ð13Þ
D
values, the time function of Eq. (6) can be ignored without
any significant loss in accuracy. The time derivatives of where W is the kinematic viscosity and T is the oscillating
Eqs. (6–9) give the respective accelerations, which are key wave period.
parameters for the estimation of the inertial forces acting A correct evaluation of the total hydrodynamic load act-
on offshore structures located in the flow field. ing on an offshore structure must consider the combined
current flow and wave particle velocities. Wave particle velo-
cities and accelerations are sinusoidal in time, and when seen
Morison’s Equation
as functions of time, are out of phase by 90 from each other.
Morison, Johnson, and Schaff (1950) formulated an equa- Adopting the absolute value of the velocity term (i.e., ju(t)j)
tion to predict the wave forces acting on a vertical pile in Eq. (11) synchronizes the drag force and flow direction.
exposed to horizontal sinusoidal oscillatory flow. In their Irregular variations of the total hydrodynamic load are pro-
equation, the linear inertial force and adapted quadratic duced against time for an idealized record, as demonstrated
drag force (from real flows and constant currents) are super- by the experimental data shown in Figure 6 for a monopile
imposed to obtain the resultant force acting on the projected at reduced laboratory scale. For inertia dominated wave
area. For larger offshore structures (e.g., gravity foundations loads and (or) small current velocity, the loads arising from
and also large diameter monopiles), the wave field is signifi- the ocean current can be ignored (DNV 2011).
cantly influenced and a diffraction regime emerges. The The dominancy of the drag force or inertial force has been
potential flow theory is more suitable for the calculation of assessed in terms of the Keulegan–Carpenter number. Typi-
wave loads on such structures (Batchelor 1967). Morison’s cal recommendations give the inertial force dominating for
formula is strictly limited for use with slender structural ele- KC < 10, comparable inertial and drag forces for
ments; i.e., characterized by D=k < 0.2: where D is the diam- 10 < KC < 20, with the drag force dominating for KC > 20.
eter of the structural element between the seabed level and In general, the CM and CD coefficients are dependent on
the transition piece (see Figure 3) and k is the impinging both the Reynold’s and Keulegan–Carpenter’s numbers
wavelength of the ocean wave. However a significant num- (Haritos 2007). For OWT monopile foundations, widely
ber of existing offshore structure designs have employed used design codes such as API (2010) and DNV (2011)
Morison’s equation even though the criterion of D=k < 0.2 specify=recommend appropriate CD and CM values, depend-
may not have always been fully satisfied (Haritos 2007). ing on the pile surface roughness and (or) the Re number.
Analysis and Design of Offshore Monopile Foundations 7
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

Fig. 7. Structural model of flexible wind-turbine system.

equivalent steel pipe for the support structure. In this case,


the first natural frequency (fnat, in Hz) of the whole structure
can be approximated by (Tong 2010):

2 3:04 EI
fnat ¼ ð14Þ
4ðpÞ ð0:227 lLi þ Mt Þ ðLi Þ3
2

where l is the structure mass to length ratio (kg=m); Li is the


structure height=(strut length) and EI is its bending stiffness
(Nm2).
Offshore wind-turbine OWT structures are excited by
Fig. 6. Graphical representations of variations in (a) velocity wind and waves, with the effective wind load determined
and acceleration, and (b) horizontal force against time for an by a complex interaction between the structural dynamics
idealized record related to a monopile at reduced laboratory of the turbine and the wind field, which contains turbulent
scale (Morison, Johnson, and Schaff 1950). gusts caused by eddies in the flow. The turbulent wind field
originates from atmospheric turbulence, with nearby
turbines also contributing to the disturbance of the flow.
Correlation between Loading Frequency and Natural
Energy-rich wind turbulence occurs at below 0.1 Hz
Frequency of OWT Structures
(LeBlanc 2009). The height of the ocean waves is usually
It is essential to consider the fundamental natural frequency defined in terms of ‘significant wave height’ and determined
of an OWT structure for a proper description and evaluation as the mean value of the highest one-third of the waves in a
of its dynamic behavior. For dynamic systems, resonance given wave record. Ocean waves that induce fatigue loading
occurs if an excitation frequency gets close to the structure’s with high frequency on offshore structures usually have sig-
natural frequency. For OWTs, this invariably leads to the nificant wave heights ranging 1.0–1.5 m and also have a
development of higher stresses and, more significantly, to a zero-crossing period (Tz) of 4–5 s (De Vries 2007).
higher range of stresses in the support structure, which is Modern OWTs are installed with either pitch-regulated
an unfavorable situation with respect to fatigue life. Hence blades or variable rotational-speed systems to enable optimi-
it is important to ensure that excitation frequencies having zation of their power production over a wide range of wind
high energy levels do not coincide with the support struc- speeds. The rotational speed of the main rotor shaft is typi-
ture’s natural frequency. As a first approximation, the sup- cally in the range 10–20 rpm. Hence the first excitation fre-
port structure’s natural frequency can be determined by quency ‘1P’ (i.e., corresponding to one full revolution)
considering a simplified geometry for the whole structure occurs in the range 0.17–0.33 Hz and, in general, should only
(Figure 7). The turbine mass (Mt) is concentrated at the top be lightly excited. Large excitations in the 1P frequency
(free end), similar to a cantilevered strut, which represents an range generally arise from excessive turbine mass or
8 M. Arshad & B. C. O’Kelly

comprehensive literature supports inherent anisotropic


behavior (or the general cross-anisotropy present in many
sedimentary deposits) for the soil’s strength=deformation
characteristics and also seepage properties (Yasin and
Tatsuoka 2000; Wang and Lade 2001; O’Kelly 2005, 2006).
Due to the initial and stress-induced anisotropy of most
deposits (Naughton and O’Kelly 2004), soil deformations
can occur whenever the magnitudes of the three principal
stresses change and (or) the principal stress axes reorientate
on account of the cyclic lateral loading. An analogy may be
drawn between the behavior of the soil in the immediate
vicinity of an OWT monopile foundation and that of a
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

pavement under repeated wheel loading of varying intensity.


For both scenarios, changes in both the magnitudes and the
directions of the principal stresses occur in a single cycle.

Fig. 8. Excitation frequency ranges for offshore wind turbines Simulation of In-situ Stress Conditions in the Geotechnical
having rated power-generation capacities ranging 2.0–3.6 MW Laboratory
(LeBlanc 2009). (a) Deflected shape of monopile. (b) Soil press-
The values of pertinent parameters used to describe the soil
ure pt exerted due to pile deflection yt for a specific depth xt. (c)
Winkler model approach and change in shape of p–y curves
response under cyclic loading can be determined in the geo-
with depth. technical laboratory using cyclic triaxial tests (Das 2008),
although the system of cyclic axial loading and lateral con-
finement pressure acting on the test-specimen is axisym-
aerodynamic imbalances. For a three-bladed turbine, the metric. An advancement on the cyclic triaxial apparatus is
blade passing-frequency of typically 0.5–1.0 Hz is denoted the hollow cylinder apparatus (HCA) (O’Kelly and
by the ‘3P’ frequency, which is heavily excited, mainly on Naughton 2005a) which allows independent control of the
account of the impulse-like excitation arising from the blades magnitudes of the three principal stresses and also the orien-
passing by the tower. Site-specific spectral densities for wind tation of the major–minor principal stress axis. The HCA is
and waves can be derived from measured site data, met-ocean ideal for simulating cyclic multi-directional loading con-
databases or using numerical models (LeBlanc, Houlsby, and ditions on cross-anisotropic test specimens. Generalized
Byrne 2010). Figure 8 illustrates the 1P and 3P excitation stress-path testing can be performed in which the stress his-
ranges, along with realistic normalized power-spectra tory and in-service loading conditions at specific locations in
representing aerodynamic and hydrodynamic excitations. the soil foundation can be simulated under stress- (O’Kelly
The regions before the 1P frequency range and after the and Naughton 2009) or strain-controlled conditions
3P frequency range are referred to as the ‘Soft–Soft’ and (Naughton and O’Kelly 2005; Das 2008). Special prep-
‘Stiff–Stiff’ zones, respectively. The structure will be too flex- aration techniques (O’Kelly and Naughton 2005b) are
ible if its natural frequency falls within the ‘Soft–Soft’ zone required to prepare=reconstitute the test specimen (which
and too rigid (heavy and expensive) if its natural frequency is often disturbed during the sampling procedure) in a phy-
falls within the ‘Stiff–Stiff’ zone; both of these scenarios sically identical condition to that of the in-situ deposit. In
making it unsuitable for the design. Wind and wave-turbu- many practical situations, laboratory testing may become
lence excitation frequencies usually fall within the ‘Soft–Soft’ too laborious, expensive, and (or) time consuming. In-situ
zone: another important reason for avoiding this frequency testing techniques, including the Cone Penetration Test
region (LeBlanc, Houlsby, and Byrne 2010). method (Igoe et al. 2013) are used for offshore site investiga-
tions and afford another approach in the determination of
the pertinent design parameter values.
Modeling Soil Behavior Under Cyclic Loading
Real Soil Behavior Modeling Soil Under Cyclic Loading
Soil deposits can be classified in many ways; e.g., by forma- The strain accumulation occurring in soil under repeated
tion process, grain size (fine or coarse), plasticity index, age loading is dependent on the material properties, stress
(recent or aged deposits), mineralogical content, etc. Apart path=level, and number of load cycles (Niemunis,
from at very small strain levels of <103 strain (Atkinson Wichtmann, and Triantafyllidis 2005; Karg 2007). Many
and Sallfors 1991; O’Kelly and Naughton 2008), the stress– models with different complexity and acceptability have
strain relationship for soil is generally highly nonlinear been developed for the prediction of the strain accumulation
(inelastic) (Budhu 2011), with the strength and stiffness occurring in a soil element under cyclic loading. These
properties strongly dependent on stress history, drainage include models that are broadly based on: the number of
conditions (drained or undrained) and the stress path load cycles (Barksdale 1972; Sweere 1990; Hornych, Corte,
followed during loading. For undisturbed deposits, and Paute 1993); the stress level (Paute, Hornych, and
Analysis and Design of Offshore Monopile Foundations 9

Benaben 1996); the number of load cycles and the stress level in-service over many decades. However, caution is necessary
(Pappin 1979; Lentz and Baladi 1981; Li and Selig 1996; in applying this methodology to OWT monopile foundation
Lekarp and Dawson 1998; Chai and Miura 2002) or the design, since the approach may often be applied outside of
number of load cycles, stress level and material properties its verified range and several important design issues may
(Niemunis, Wichtmann, and Triantafyllidis 2005; Karg not be properly considered. The API (2010) and DNV
2007). A major limitation to their application in offshore (2011) standards rely on methods (models) built upon
foundation design calculations is that none of these models empirical data obtained for long flexible piles, for which
explicitly consider the (mono)pile–soil interaction under lat- bending (deflection) is significant. In contrast, existing and
eral loading. In reality, the pile deflection (rotation) response planned OWT monopile foundations invariably have slen-
under lateral loading arises from the soil behavior, which is derness ratios of <10 (typical range of 5–6), indicating rigid
dependent on the loading conditions. A few models have pile behavior (Achmus, Kuo, and Abdel-Rahman 2009;
been tailored for this particular scenario; these are described Peng, Clarke, and Rouainia 2011). Under these circum-
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

in the next section, in the context of the design of OWT stances, the pile rotation is generally more prominent over
monopile foundations. A discussion on numerical modeling, bending, with the rotation occurring about a point (‘axis
considering dynamic constitutive soil models, torsional load- of rotation’) located approximately one pile diameter above
ing, and damping related issues, is beyond the scope of this the pile base. Hence the design criteria and analyses appro-
review article. Details on these topics can be found in Basack priate for flexible and rigid piles are considerably different
and Dey (2012), Basack and Sen (2014), Guo (2006, 2013) (Dobry et al. 1982), casting doubt on the application of
and Rani and Prashant (2014). the API (2010) and DNV (2011) methods based on p–y
curves in predicting the in-service behavior of offshore
monopiles. Further, the p–y curves for cyclic loading pre-
Strain Accumulation Models Considering Pile–Soil sented in API (2010) and DNV (2011) were primarily formu-
Interaction lated for the evaluation of the ultimate lateral load-carrying
capacity mobilized under relatively few load cycles. In con-
p–y Model trast, OWT monopile foundations experience many millions
In general terms, a p–y curve is typically obtained by plotting of low-amplitude cycles over their in-service life. Further, the
the soil pressure (p) response against the pile’s lateral
deflection (y) arising from the action of a horizontal load
(H) applied at the pile head (Figure 9a). Figure 9b shows
the soil pressure (pt) distribution generated around the pile
circumference at a particular depth (xt) and the correspond-
ing pile deflection (yt) response. In the literature curves (e.g.,
Matlock 1970; Reese, Cox, and Koop 1975; Ismael 1990; API
2010), p–y curves can be categorized on the basis of soil type
(granular or cohesive), loading type (monotonically increas-
ing or cyclically repeated) and the groundwater table level.
The effects of soil stratification, nonlinearity, and other
soil properties are automatically considered by determining
p–y curves specific to different depth ranges along the length
of the pile (Figure 9c), which is typically modeled using
Winkler’s approach; i.e., the pile member acts as a beam
supported by a series of uncoupled nonlinear elastic springs
that represent the soil reaction. For instance, soil stiffness
generally increases with depth (overburden pressure), which
is reflected by increasing values of the spring stiffness (Epy),
defined as the secant modulus of the p–y curve (Figure 9c).
The pile deflection that develops under given loading con-
ditions and constrains can be predicted by implementing
the related p–y curve in a simple non-dimensional frame-
work, such as Randolph’s method (Randolph 1981), or by
numerical methods using computer software such as
COM624P (1993). However, despite the stiffness Epy being
a soil–structure interaction parameter, API (2010) and
DNV (2011) only consider the soil properties in formulating
the p–y curves, and the influence of the pile properties on the
mobilized p–y curves remains an open question.
Current pile design methodology based on p–y curves, as
described in API (2010) and DNV (2011), has gained
broad recognition on account of the low failure rate for piles Fig. 9. p–y method of analysis for laterally loaded pile.
10 M. Arshad & B. C. O’Kelly

API (2010) and DNV (2011) methods do not provide a determined by the pile length and the relative stiffness Tr
means of calculating the accumulated pile deflection [¼(EI=nhN)0.2] value.
(rotation) that occurs during cyclic loading. Changes in the In Eq. (16), all of the variables can be taken as known
foundation stiffness as a result of long-term cyclic loading, except for coefficient nhN, which is calculated as:
which typically densifies (but in some circumstances may
loosen) the surrounding soil (LeBlanc, Houlsby, and Byrne nhN ¼ nh1 N t ð17Þ
2010), are also poorly accounted for in current design meth-
where nh1 is the coefficient of soil reaction for the first cycle
odologies based on p–y curves.
of loading, its value dependent on the relative density of the
soil and the location of the groundwater table (Terzaghi
Model of Little and Briaud (1988) 1955; Reese, Cox, and Koop 1974), with the exponent t (hav-
The Little and Briaud (1988) model is based on experimental ing a recommended range of 0.2–0.4 (Broms 1964; Davisson
1970)) determined empirically by:
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

results obtained from six cyclic lateral load tests performed


on long pre-stressed concrete and steel pipes (D ¼ 0.6–1.06 m)
installed in a 22 m thick stratum of loose to medium dense sand, t ¼ 0:17 FL Fi FD ð18Þ
overlying stiff clay. Twenty load cycles were performed for each where FL, Fi, and FD are empirically determined factors that
test pile. The following power function was proposed for esti- take into account the influence of the cyclic load ratio, pile
mating the accumulated lateral strain (en) developed at ground installation method, and soil density, respectively. Note
surface level for N number of load cycles: nh1 ¼ nh; i.e., its value is the same as that for the monotonic
en ¼ e 1 N m ð15Þ (static) loading condition.
The Long and Vanneste (1994) model provides a rela-
where e1 is the strain response at the end of the first load cycle tively straightforward procedure for predicting the effect of
and exponent m is a parameter that accounts for the influence cyclic lateral loading, but specifically restricted to scenarios
of the soil properties, pile material, installation method, and involving the soil reaction modulus increasing linearly with
loading characteristics, with experimentally derived m values depth. Soil stratification, nonlinearity, and other fundamen-
ranging 0.04–0.09 for the ‘flexible’ test piles investigated (Little tal parameters (e.g., unit weight and strength) that affect the
and Briaud 1988). In this context, the lateral strain is calculated lateral load response are not explicitly taken into account by
as the pile’s lateral deflection occurring at the seabed level div- this model. Further, most of the thirty-four pile tests
ided by its outer diameter; a dimensionless quantity. considered in the development of Eq. (16) involved the
Numerical studies by Kuo, Achmus, and Abdel-Rahman application of less than fifty lateral load cycles.
(2012) indicated an m value of 0.07 for ‘flexible’ piles (agree-
ing with the experimental range reported by Little and Briaud Model of Lin and Liao (1999)
(1988)) and 0.135 for ‘rigid’ piles. The validity of Eq. (15) has Based on the results of twenty-six full-scale piles tested under
been verified using cyclic lateral load test data obtained for cyclic lateral loading (N < 50), Lin and Liao (1999) proposed
model piles installed in medium-dense quartz sand (Peralta a logarithmic trend to capture the accumulated strain en
and Achmus 2010). It has also been shown that the value of developed at the ground surface level for N load cycles:
exponent m can be evaluated from consolidated–drained cyc-
lic triaxial test data (Peralta and Achmus 2010). Nevertheless, en ¼ e1 ½1 þ t lnðN Þ ð19Þ
Eq. (15) is empirical and formulated on limited test data where t is a degradation parameter determined by:
(N < 20) for a loose to medium dense sand deposit.
Model of Long and Vanneste (1994) L
t ¼ 0:032 FL Fi FD ð20Þ
The Long and Vanneste (1994) model is based on a closed- Tt
form solution for a pile in an elastic foundation soil, with where Tt ¼ ðEI=nh Þ0:2 , L is the embedded length of the pile
the soil reaction modulus increasing linearly with depth. and nh is the coefficient of soil reaction for the monotonic
By applying this approach and using input parameters (static) loading condition.
derived from reported field data for cyclic lateral load tests The other contributing factors in Eq. (20) are similar to
performed on thirty-four piles of different lengths, dia- those in Eq. (18), with insignificant differences in values
meters, materials, and installation techniques, Long and from those reported for the Long and Vanneste (1994)
Vanneste (1994) proposed that the accumulated lateral strain model. However, considering the various influencing factors,
occurring at ground surface level for N load cycles can be caution is urged regarding empirical parameters used in
estimated by: describing the logarithmic change in the pile deflection=
AH BM rotation.
en ¼ þ ð16Þ In their study, Lin and Liao (1999) considered two load-
EI0:4 ðnhN Þ0:6 EI0:6 ðnhN Þ0:4
ing scenarios: one-way loading, in which the applied hori-
where EI is the flexural rigidity of the pile; nhN is the zontal load increases from zero to its maximum value
coefficient of soil reaction for the Nth cycle of loading; H (Hmax) before reducing back to zero, thereby completing
and M are the horizontal load and moment, respectively, one cycle of loading; two-way loading, in which the horizon-
applied at the ground surface level; A and B are scalars tal load is applied in the opposite direction for half of each
Analysis and Design of Offshore Monopile Foundations 11

model explicitly considers the effects of soil density, mean


effective confining pressure, cyclic stress amplitude, simul-
taneous oscillation of stresses from different directions,
and the number of load cycles:

e_acc ¼ fampl f_N fe fp fY fp ð21Þ


where e_acc is the derivative of strain accumulation with
respect to the number of load cycles N, fampl is the strain
amplitude function, f_N is the load-cycle number function,
fe is the soil density function, fp is the mean effective confin-
ing pressure function, fY is the average stress ratio (i.e.,
mobilized deviatoric stress to average mean confining press-
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

ure applied during a load cycle) function and fp is the polar-


ization (load cyclic shape) function. Further details on the
determination of these functions and commonly adopted
values are reported by Wichtmann, Niemunis, and Trianta-
Fig. 10. Different modes of load variation against time.
fyllidis (2009), Wichtmann et al. (2010).
Wichtmann et al. (2010) applied the high cycle accumulation
cycle, reducing its value below zero to a minimum of Hmin model for typical North Sea fine sand deposits and found that
(see Figure 10). greater permanent (plastic) lateral deflection occurred for:
According to the Long and Vanneste (1994) and Lin and
. Decreasing relative density, with the average value of the
Liao (1999) models, one-way loading produces greater accu-
mulated strain compared with two-way loading. Similar applied moment (Mave) and its amplitude (Mampl) held
overall behavior has been reported from studies on model constant;
. Increasing Mampl, with the relative density and Mave held
monopiles (Peng, Clarke, and Rouainia 2011), with balanced
two-way loading (Hmax ¼ Hmin) producing about 80% of constant;
. Increasing Mave, with the relative density and Mampl held
the lateral deflection recorded under one-way loading of
the same magnitude. Unbalanced two-way loading constant.
(Hmax 6¼ Hmin) was reported to produce even lower These outcomes are qualitatively conceivable and broadly
amounts of lateral deflection. However some recent physical consistent with the behavior observed in laboratory studies
model studies suggest that unbalanced two-way loading is on model monopiles by LeBlanc, Houlsby, and Byrne
more damaging to the lateral stability of the monopile (2010) and Peng, Clarke, and Rouainia (2011), although
(Zhu, Byrne, and Houlsby 2013; Arshad and O’Kelly further physical model studies and in-situ testing are
2014). An increase in soil stiffness was also observed with required to confirm these findings. Potentially significant
increasing number of load cycles. Similar formulations to shortcomings of the high cycle accumulation model include:
that given by Eq. (19) have been proposed by Verdure, the multiplicative approach used in Eq. (21) means that
Garnier, and Levacher (2003) and Li, Haigh, and Bolton potentially small errors in the values of the individual func-
(2010) based on centrifuge experiments performed on minia- tions may produce an objectionable error in the final calcu-
ture piles installed in sandy deposits, with the degradation lation value; the formulation includes no measure of loading
parameter t estimated by curve-fitting analysis of the experi- frequency which is an important factor in connection with
mental data. The change in soil stiffness was found highly the natural vibration frequency of OWT structures [see
dependent on the cyclic load amplitude. Up to certain levels Section on Correlation between Loading Frequency and
of cyclic load amplitude and numbers of load cycles, the stiff- Natural Frequency of OWT Structures] and the strain
ness was found to increase rapidly, but no further change was accumulation rate.
observed for higher load amplitudes, even during the first few
cycles (Verdure, Garnier, and Levacher 2003). Model of Achmus, Kuo, and Abdel-Rahman (2009)
The Achmus, Kuo, and Abdel-Rahman (2009) model is
High Cycle Accumulation Model (2005, 2010) based on the concept of the soil stiffness degradation
The high cycle accumulation model is based on the experi- approach suggested by different researchers (e.g., Little
mental data obtained from a large collection of experimental and Briaud 1988; Long and Vanneste 1994), with emphasis
work, comprising index tests and drained cyclic triaxial, res- on the degradation of the soil secant stiffness (Es), in keeping
onant column and simple shear tests performed on sand spe- with formulations proposed by Huurman (1996) and Werk-
cimens having a wide range of initial relative densities meister (2003). According to this model, an increase in the
(Niemunis, Wichtmann, and Triantafyllidis 2005). Although plastic strain can be interpreted as a decrease in the Es value.
this model was initially developed without consideration of In cyclic triaxial tests (for which elastic strains are usually
pile–soil interaction, its potential application in modeling a negligible) and referring to Figure 11, the plastic axial strains
pile–soil system under low-amplitude stress cycles has been produced at the end of the first and Nth cycles (epN ¼ 1 and
demonstrated by finite-element calculations using ABAQUS epN ¼ N) can be determined from the rate of degradation of
software (Wichtmann et al. 2010). The formulation of this the secant modulus measured for the first and Nth cycles
12 M. Arshad & B. C. O’Kelly
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

Fig. 11. Degradation of soil secant modulus (Es) under cyclic loading.

(EsN ¼ 1 and EsN ¼ N respectively) by: and accounts for the loading and unloading phases emerging
epN¼1 due to the cyclic horizontal loading. Achmus, Kuo, and
EsN¼N
¼ ð22Þ Abdel-Rahman (2009) found good agreement between
epN¼N EsN¼1 numerical simulations performed using cyclic triaxial test
data (Timmerman and Wu 1969) and experimental data
In the Achmus, Kuo, and Abdel-Rahman (2009) study, for laterally loaded model monopiles installed in sand.
the characteristics of the permanent strains derived from
Model of LeBlanc (2009)
drained cyclic triaxial test data for medium and dense sand
were implemented in a 3D finite-element model of a laterally The LeBlanc (2009) model is based on lateral load tests,
loaded pile. The Achmus, Kuo, and Abdel-Rahman (2009) involving between 8  103 and 6  104 load cycles, performed
model suggests that the accumulated strain en developed on rigid model piles having a slenderness ratio of 4.5 and
for N load cycles can be estimated using the semi-empirical which were installed in unsaturated, very loose and medium
approach proposed by Huurman (1996) for the calculation dense Leighton Buzzard sand beds. A non-dimensional
of the accumulated plastic strains in cyclic triaxial tests, as: framework was developed to identify realistic pile dimen-
sions and loading ranges for the laboratory study. Some
e1 particular loading characteristics were found to cause signifi-
en ¼ ð23Þ
b1 ðxc Þb2 cant increases in the accumulated pile rotation to occur, with
ðN Þ
the soil stiffness increasing with the number of load cycles.
where b1 and b2 are model parameters and xc is the charac- This contrasts starkly with the Long and Vanneste (1994),
teristic cyclic stress ratio (ranging 0–1 (Achmus, Kuo, and API (2010) and Achmus, Kuo, and Abdel-Rahman (2009)
Abdel-Rahman 2009)), defined as: models for which the static load–displacement curves are
ðcyclic stress ratio at loadingÞ degraded to account for cyclic loading.
LeBlanc (2009) proposed that the accumulated rotation hN
 ðcylic stress ratio at unloadingÞ (and hence the lateral displacement) of a ‘rigid’ pile developed
xc ¼
1  ðcylic stress ratio at unloadingÞ for N load cycles can be estimated by (see Figure 12):

The cyclic stress ratio at constant confining pressure is hN ¼ h0 þ DhðNÞ ð24Þ


defined as the ratio of the major principal stress under cyclic where h0 is the pile rotation achieved when the applied
loading to the major principal stress at failure under static moment reaches its maximum value during the first load cycle
loading. Since the horizontal loading varies cyclically for and Dh(N) is given by:
OWT monopile foundations, the confinement pressure act-
ing on soil elements surrounding the pile shaft also changes.
DhðN Þ ¼ hs Tb ðfb ; Rd ÞTc ðfc Þ N 0:31 ð25Þ
Hence the cyclic stress ratio values for cyclic triaxial tests (in
which the axial load is varied cyclically) and OWT monopile where hs is the pile rotation produced by a static load having
foundations (where the axial load remains constant and the the same magnitude as the maximum cyclic load (Figure 12b),
horizontal load varies cyclically) are different. To overcome Tb and Tc are dimensionless functions, Rd is the relative
this, the ‘characteristic cyclic stress ratio’ is used in Eq. (23) density of the sand, with fb (a measure of the cyclic load
Analysis and Design of Offshore Monopile Foundations 13
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

Fig. 13. Functions relating (a) Tb and (b) Tc to the cyclic load
Fig. 12. Variation in pile rotation h with fluctuation in moment
characteristics in terms of fb and fc respectively.
M. (a) Cyclic test (b) Static test.

amplitude normalized with respect to the static moment Further, this model cannot account for the multidirectional
capacity, MR) and fc (quantifies characteristics of the cyclic loading typical of offshore wind-farm environments.
load) given by: Considering the accumulated strain en produced at the
ground surface level for a rigid pile is proportional to the
Mmax accumulated rotation, then the predictions by this model
fb ¼ ð26Þ (with due consideration for the cyclic character of the load-
MR
ing) are contrary to those for flexible piles according to the
Mmin Long and Vanneste (1994) and Lin and Liao (1999) model.
fc ¼ ð27Þ For loose and medium dense sand, LeBlanc (2009) also
Mmax
reported an increase in soil stiffness due to cyclic loading
where Mmin and Mmax are the minimum and maximum of the rigid pile, although further research for higher densi-
moments developed in a load cycle. fication levels, other loading frequencies and degrees of satu-
It follows that 0 <fb < 1, with fc equal to unity for a static ration are necessary to confirm this preliminary experimental
load test, zero for one-way loading and 1 for balanced two- finding.
way loading. Values of Tb and Tc, determined from experi-
mental data of Dh(N)=hs against N, were plotted against fb Model of Bienen et al. (2012)
and fc, as shown in Figure 13. Note that with Tc ¼ 0 for The Bienen et al. (2012) model is based on experimental data
fc ¼  1.0, Eq. (25) proposes that the rotation occurring from miniature monopiles installed in dry medium-dense
after the first load cycle is equal to zero for balanced two- sand and tested at 1g and 200g, using a centrifuge apparatus
way cyclic loading, which is fairly unrealistic in practice. for the latter. The test piles were representing a prototype
14 M. Arshad & B. C. O’Kelly

pile 2.4 m in diameter (D), with embedment lengths (L) of 9.6 Design Procedure for OWT Foundation System
and 30 m; i.e., 4D (‘rigid’) and 12.5D (flexible) piles, respect-
ively. Using a sinusoidal waveform, different magnitudes of Design Motive
one-way lateral loading were investigated considering 104 The basic driving motive for the design procedure is to avoid
load cycles and a single model frequency of 0.25 Hz. The the occurrence of resonance in the dynamic behavior of the
accumulated lateral deflection y occurring at the pile head structural system under in-service loading (Jaimes 2010). An
(assumed flush with the surrounding ground surface) for N iterative procedure is usually adopted in design, with the
load cycles can be estimated by: basic steps involved in the design process for an OWT mono-
 as pile foundation system illustrated in Figure 14. The data
H
y ¼ D fN As 100 2 ð28Þ required include environmental, turbine and site stratigra-
D LQc phy (soil profile) data. The environmental data is used to
where H is the horizontal load applied to the pile head, Qc is determine the required work-platform and hub elevations
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

the rate of change in CPT cone-tip resistance with depth, As (see Figure 3) for the proposed OWT, and to select initial=
and as are dimensionless parameters dependent on the soil trial dimensions=geometry for the monopile foundation,
properties and pile dimensions (Dyson and Randolph leading to the determination of the natural frequency of
2001; Dührkop 2009), with reported values of As ¼ 1.4 and the whole structural system. Checks on resonance frequency
as ¼ 0.0072 for monopiles (Bienen et al. 2012), and fN (factor are applied along with predictions of the anticipated
to modify the monotonic response (deflection) into a cyclic rotation, deflection, and settlement responses of the pro-
response) approximated by: posed foundation system produced by the applied loads=
moments that are determined from the environmental and
N 1 wind turbine data. Fatigue and buckling checks are per-
fN ¼ 1 þ BN1 ðlnðBN2 þ 1Þ ð29Þ formed at a more advanced stage of the design process,
N
usually using some computer software package, and hence
where the parameters BN1 and BN2 are determined from a are beyond the scope of this paper. The following sections
plot of [(y=D)n=(y=D)1] against number of load cycles N. present a design example in which the work-platform and
This model, based on the one-way lateral loading and sin- hub elevations along with the required embedment length
gle frequency of 0.25 Hz investigated, was found to provide for a monopile foundation supporting a typical 5 MW
reasonable predictions of the accumulated strain produced OWT (Jonkman et al. 2009) are determined.
under cyclic loading, although the soil stiffness response
was not clear: increasing for up to certain levels of accumu-
lated strain, but then decreasing with further load cycles. Determination of Work Platform and Hub Elevations
Hence these findings should be verified for greater numbers
Referring to Figure 14, the first step in the design process
of load cycles occurring in multiple directions and for other
requires the determination of the turbine platform and hub
frequency ranges, degrees of soil saturation and densification
elevations, usually with reference to the lowest astronomical
levels.
tide (LAT); defined as the lowest tide level to occur under
Model of Klinkvort and Hededal (2013) average meteorological conditions and any combination of
Klinkvort and Hededal (2013) applied the model proposed astronomical conditions. LAT is below the mean sea level
by LeBlanc (2009) to experimental data for ‘rigid’ model (MSL). The high astronomical tide (HAT) is above the
piles (slenderness ratio of 6) installed in saturated and dry MSL, but below the storm surge level.
dense sand that were tested at ng using a centrifuge appar- The work-platform level is located at the top of the
atus. These tests involved a maximum of 500 load cycles transition piece and it determines the location of the flange
and were performed at the same frequency and soil density.
They proposed that the pile’s accumulated lateral strain pro-
duced at the ground surface level after N number of load
cycles (en) can be estimated as:
Tb ðfbb ÞTc ðfcc Þ
en ¼ e1 N ð30Þ
with fbb and fcc defined as:

Pmax
fbb ¼ ð31Þ
Pu
Pmin
fcc ¼ ð32Þ
Pmax
where Pu is the ultimate lateral load-carrying capacity under
monotonic loading applied to the pilehead; Pmax and Pmin
are the maximum and minimum lateral loads, respectively, Fig. 14. Flowchart for the design of offshore wind-turbine
applied to the pilehead during the cyclic loading. support structure and monopile foundation.
Analysis and Design of Offshore Monopile Foundations 15
Table 2. Values of key parameters for typical 5 MW wind Determining the Required Natural Frequency and
turbine Preliminary Tower Dimensions
Parameter Magnitude From the rotational speed interval of the main rotor shaft,
as given in Table 2, the 1P region ranges 0.12–0.20 Hz and
Turbine mass (rotor and nacelle) 350 tonnes the 3P region ranges 0.36–0.60 Hz (Jonkman et al. 2009).
Rotor diameter 126 m For this particular design example, the target value of the
Nominal rotor speed 10.1 rpm natural frequency [Eq. (14)] was set as 0.27 Hz; i.e., within
Rotational speed interval 6.9–12.1 rpm the ‘wanted frequency’ range of 0.20–0.36 Hz. By taking into
of the main rotor shaft
account the hub height, wind-turbine mass and target natu-
Cut-in wind speed 4 m=s
ral frequency, the outer diameter and wall thickness at the
Nominal wind speed 11 m=s
top of the steel tower were estimated as 3.9 m and 40 mm,
Cut-out wind speed 25 m=s
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

respectively, with corresponding values at the tower base


of 6.6 m and 100 mm. These preliminary selected dimensions
connection between the transition piece and the turbine must also satisfy the buckling and fatigue checks; otherwise
tower (see Figure 3). The magnitudes of the applied wind the calculations are revised iteratively. The outer diameter of
loads and the location of the centre of gravity of the nacelle the transition piece (DTP) subsequently depends on the diam-
mass is strongly dependent on the hub height. Representa- eter of the monopile, as:
tive values for key parameters of the typical 5 MW wind tur-  
bine considered in this design example are given in Table 2. DTP ¼ D þ 2 tTP þ tgrout ð36Þ
Environmental data characteristic of an offshore wind-farm where D is the monopile diameter (usually equal to the tower
located in the North Sea are given in Table 3. For this base diameter), tTP is the wall thickness of the transition
location, a tidal range (DZtide ¼ HATLAT) of 1.6 m and piece and tgrout is the thickness of grout connection.
storm surge (DZsurge) of 2.0 m may be adopted (Jonkman
et al. 2009; Fischer 2011).
The required work-platform level (Zplatform) is given by: Calculation of Design Loads
The design loads for the subject monopile foundation
Zplatform ¼ LAT þ DZtide þ DZsurge þ DZair þ n ð33Þ system were estimated using the preliminary tower geometry,
the wind-turbine parameter values given in Table 2 and the
where DZair is the air gap of typically 1.5 m between the
environmental data given in Table 3. For this example, the
work-platform level and the highest wave crest elevation (n).
software package ‘RECAL’ was used to estimate the shear
For an assumed wave-height coefficient value of 0.65 and
force and bending moment acting on the monopile foun-
a maximum wave height, relative to the MSL, having a 50
dation at seabed level, as documented by De Vries and
year return period (Hmax(50 year)) of 20.3 m, the value of
van der Tempel (2007). The torsional moments are usually
n is given in DNV (2011) as:
only a small fraction (<1%) of the bending moment (Lesny
n ¼ 0:65 Hmax ð50 yearÞ  13:2 m ð34Þ and Wiemann 2005); e.g., see typical loading values listed in
Table 1 for a 5 MW wind turbine having a monopile foun-
And hence the value of Zplatform is determined as
dation system. Hence, provided checks on lateral forces
LAT þ 19 m (rounding the latter up to the nearest whole
and bending moments are satisfied, torsional moments are
integer). A check should also be applied (De Vries and van
generally not critical and have not been specifically con-
der Tempel 2007 DNV 2001) such that Hmax(50
sidered in design examples for OWT monopile foundation
year) < 0.78  30 m; the assumed MSL value for these calcula-
systems reported in the literature. RECAL can simulate both
tions. The required hub height (Zhub) of LAT þ 87m above
wind and wave time-series, from which it calculates the
seabed level is determined from:
environmental loads acting on the support structure, as well
Zhub ¼ Zplatform þ 0:5ðrotor diameterÞ þ 5 ð35Þ as the structure’s dynamic behavior. For the design example
under consideration, Table 4 presents load data in terms of
With LAT ¼ MSL  Ztide=2, Zhub ¼ 116.2 m from seabed the lateral forces and bending moments generated by
level. RECAL. The reported load values take into account the
safety factors for hydrodynamic loading and aerodynamic
Table 3. Extreme wind velocity (Vw), current velocity (Uc),
significant wave height (Hs), and maximum wave height (Hmax)
as a function of return period (Treturn) for reference offshore Table 4. Estimated design loads occurring at seabed level for
wind-farm site (Jonkman et al. 2009) design example

Treturn (year) Vw (m=s) Uc (m=s) Hs (m) Hmax (m) Lateral force, H Bending moment, M
Load type (MN) (MN.m)
1 33.9 0.70 7.72 14.35
5 38.0 0.80 9.03 16.80 Aerodynamic 1.42 127
10 39.8 0.84 9.60 17.85 Hydrodynamic 9.27 298
50 43.9 0.94 10.91 20.29 Total 10.69 425
16 M. Arshad & B. C. O’Kelly

thrust on the rotor, in addition to the different critical design example are given in Figure A1 of the Appendix. It
combinations of wind and wave loading, as reported in was assumed that the monopile was made of steel having
several guidelines=standards (API 2010; DNV 2011). yield stress and Young’s modulus values of 248 MPa and
207 GPa, respectively.
The embedment length of the monopile was initially
Monopile Embedment Length and Foundation Stability
assumed equal to nine times its outer diameter. Checks on
The monopile’s embedment length must be sufficient to the pile rotation and (or) its deflection occurring at the
ensure vertical and lateral stability. Studying the interaction seabed level are applied and the monopile embedment length
effects for piles under combined axial and lateral loading is optimized accordingly. The monopile’s rotation (h0) and
would call for a systematic and sophisticated analysis, lateral deflection (v0) occurring at the seabed level under
although the pertinent literature is limited and sometimes the design loads (horizontal force (H) and bending moment
contradictory. For instance, analytical studies by Ramasamy (M)) are calculated by considering the closed-form solutions
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

(1974) and Goryunov (1973) indicate that for a given lateral proposed by Randolph (1981), Broms (1964) and Matlock
load, the pile’s lateral deflection increases for the combi- and Reese (1960). These methods are related to monotonic=
nation with vertical loading but some experimental static loading conditions and hence there is no consideration
(Sorochan and Bykov 1976; Jain, Ranjan, and Ramasamy of the number of load cycles. Randolph’s (1981) method,
1987) and field (Bartolomey 1977; Zhukov and Balov illustrated in Figure 15, employs Eqs. (37) and (38) to calcu-
1978) studies have indicated the contrary. Numerical analy- late v0 and h0, respectively; with parameters Ep (effective
sis by Karthigeyan, Ramakrishna, and Rajagopal (2007) Young’s modulus of the pile), Gc (equivalent shear modulus
indicated that for sandy soils, the presence of vertical loads of the soil at a depth of 0.5Lc), Lc (active pile length) and Rc
increases the pile’s lateral load-carrying capacity by as much (the soil’s equivalent shear modulus profile parameter)
as 40% (depending on the magnitude of the axial loading), defined in Figure 15. In this method, the monopile dimen-
although marginal reductions in the lateral load-carrying sions are incorporated in terms of its active length and
capacity were found to occur for clayey soils. second moment of area, I; the latter an indirect involvement
According to current practice, for monopile design, separ- of its cross-sectional area (i.e., its inner and outer diameter
ate analyses are performed that consider (i) the axial loading dimensions). When the active length exceeds the actual=pro-
only, to determine the bearing capacity and settlement posed monopile embedment length (L), the monopile tends
response, and (ii) the lateral loading only, to determine the
flexural behavior through cantilever action (Karthigeyan,
Ramakrishna, and Rajagopal 2006; Moayed, Mehdipour,
and Judi 2012; Rahim and Stevens 2013). However, com-
pared with the axial loads, the lateral loads are considered
governing, as mentioned in several design guidelines (e.g.,
GL 2005; API 2010; DNV 2011) and documented by many
researchers (Achmus 2010; LeBlanc, Houlsby, and Byrne
2010; Malhotra 2011; Peng, Clarke, and Rouainia 2011;
Bhattacharya et al. 2012; Kuo, Achmus, and Abdel-Rahman
2012; Haiderali, Cilingir, and Madabhushi 2013; Lombardi,
Bhattacharya, and Wood 2013; Zhu, Byrne, and Houlsby
2013; Nicolai and Ibsen 2014; Carswell et al. 2015). For
OWT monopile foundations, the pile must mobilize suf-
ficient soil resistance over its embedded length to transfer
all the different types of applied loads to the surrounding
soil, with adequate safety factors, and prevent toe ‘kick’ (dis-
placement at the pile base) and excessive deflection=rotation
of the pile itself from occurring. The pile size and embed-
ment length necessary to satisfy the lateral load requirements
are generally greater compared with those necessary to
satisfy the axial loading requirements (Kopp 2010). Hence
this design example focuses on the lateral stability of the sub-
ject monopile.
The input data for the monopile design includes the soil
strength and stiffness (e.g., elastic (Es) and shear (Gs) modu-
lii and Poisson’s ratio) profiles against depth, the pile
properties (e.g., cross-sectional dimensions=properties and
material strength=stiffness), and the design loads (De Vries
2007; Jaimes 2010; Tong, 2010). The soil (medium dense
sand considered for the present case) properties used in the Fig. 15. Closed-form solution after Randolph (1981).
Analysis and Design of Offshore Monopile Foundations 17

to translate somewhat in the soil foundation as well as embedded length of the pile. However, for a ‘rigid’ pile, this
deflecting; hence the monopile deflection obtained through decreasing trend may not be applicable. According to this
Eq. (37) must be modified according to Eq. (39). method (Matlock and Reese 1960), beyond a certain value
 0:142  1  2 ! of embedment length, there will be no further reduction in
Ep =GC Lc Lc the monopile’s deflection or rotation. Calculations are not
v0 ¼ 0:27H þ 0:3M ð37Þ
Rc Gc 2 2 included for the present design example, but the pile embed-
ment length should be sufficient to achieve tolerable values
 0:142  2  3 ! of v0 and h0 occurring at the pile toe level. For the calcu-
Ep =GC Lc 0:5 Lc
h0 ¼ 0:3H þ 0:8ðRc Þ M lation of the v0 and h0 values occurring at the seabed level,
Rc Gc 2 2
a parametric study was performed, varying the pile’s embed-
ð38Þ ment length, outer diameter and inner diameter in the ranges
  20–40 m, 5.5–7.0 m, and 5.4–6.5 m, respectively. A summary
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

Lc
v0 ðadjusted valueÞ ¼ 0:8v0ðfrom Equation 37Þ: ð39Þ of the results of this study is presented in Figures 17–19 and
L conversed as:
. The monopile deflection=rotation values predicted by the
Broms’s (1964) method of calculation for v0 is illustrated three selected methods for the static design load conditions
in Figure 16. Matlock and Reese’s (1960) method employs grossly differed from one another; e.g., for the applied
Eqs. (40) and (41) to calculate the values of v0 and h0, loads and range of monopile embedment lengths and
respectively, occurring at the seabed level.
HTt 3 MTt 2
v0 ¼ Av0 þ Bv0 ð40Þ
EI EI
HTt 2 MTt
h0 ¼ Ah0 þ Bh0 ð41Þ
EI EI
where Av0, Bv0, Ah0, and Bh0 are sets of non-dimensional
coefficients (scalars) whose values are dependent on the
depth along the embedded length of the pile and Tt ¼ (EI=
nh)0.2, with the value of the coefficient of soil modulus vari-
ation, nh, dependent on the soil relative density and the
location of the groundwater table (Terzaghi 1955; Reese,
Cox, and Koop 1974).
For the seabed (ground surface) level, the values of coeffi-
cients Av0, Bv0, Ah0, and Bh0 are 2.43, 1.62, 1.62, and 1.75,
respectively, with these coefficients decreasing in value with
increasing depth along the pile embedment length, reflecting
the reduction in v0 and h0 as we move down along the

Fig. 16. Charts for calculating the lateral deflection occurring at Fig. 17. Using Randolph’s (1981) method, effects of pile’s
the ground surface level for laterally loaded piles in cohesionless embedment length and diameter on its (a) lateral deflection
soil (after Broms (1964)). occurring at the seabed level and (b) rotation.
18 M. Arshad & B. C. O’Kelly

minor change in v0 was observed for pile embedment


lengths greater than 26 m, irrespective of the pile’s cross-
sectional area (Figure 18).
. For the Matlock and Reese’s (1960) method, v0 and h0
were both found dependent on the inner and outer dia-
meters (cross-sectional area) of the monopile, although,
their significance tended to reduce substantially for pile
inner and outer diameters greater than 5.8 and 6.0 m
respectively (Figure 19).
The monopile’s lateral response for the long-term cyclic
lateral loading condition was assessed for the presented
design example using the models after Little and Briaud
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

(1988), Long and Vanneste (1994), Lin and Liao (1999)


and Klinkvort and Hededal (2013), which have been dis-
cussed previously in this paper. It was assumed that the
monopile had been driven into the medium dense sand
deposit and was subjected to one-way cyclic loading.
Fig. 18. Summary of the results deduced using Broms’s (1964)
Figure 20 presents the accumulated lateral response of the
method.
monopile predicted by the different models considered.
The accumulated lateral response is in fact a multiple of
inner=outer diameters considered in the design example, the response for the end of the first load cycle, realistic values
the predicted values of v0 ranged 70–270, 25–115 and of which cannot be estimated using the models considered,
25–90 mm for the Randolph’s (1981), Broms’s (1964) and for the purpose of these calculations, its value was
and Matlock and Reese (1960) methods respectively. This assumed equal to unity. Hence Figure 20 should be viewed
noticeable difference is perhaps due to the involvement of in the context of a qualitative comparison of the long-term
marked empiricism in the methods used to evaluate the predictions by the different models. From Figure 20, it can
lateral stability of the different monopile set-ups and the be concluded that for the same loading conditions, large dif-
limited available experimental data used in their ferences can occur between the accumulated lateral
formulation=calibration. responses of the monopile predicted by the different models,
. For Randolph’s (1981) method, v0 was found sensitive to perhaps by many folds for large numbers of load cycles. One
the pile’s embedment length and its inner and outer dia- reason for these large differences can be put down to the
meters (incorporated in calculations of the cross-sectional empirical selection of values for the coefficients=non-dimen-
area and second moment of area, I) (Figure 17a), whereas sional parameters incorporated in these models for the pre-
h0 was sensitive to the pile cross-sectional area only diction of the monopile’s lateral response under long-term
(Figure 17b). cyclic loading.
. For Broms’s (1964) method, v0 was found dependent on The monopile’s ultimate axial load-carrying capacity
the embedment length and inner and outer diameters (under static loading) (Qd) can be determined using
(cross-sectional area) of the monopile, although only a

Fig. 19. Summary of the results deduced using the Matlock and Fig. 20. Accumulated lateral strain responses predicted for the
Reese’s (1960) method. monopile at seabed level under long-term cyclic loading.
Analysis and Design of Offshore Monopile Foundations 19

Eq. (43), which is the recommended practice by API (2010), . Pile properties (e.g., diameter and slenderness ratio) and
and has been employed by many researchers; e.g., De Vries the soil–pile interaction;
(2007); Igoe, Gavin, and O’Kelly (2010); Haiderali, Cilingir, . Varying the direction and frequency of the applied lateral
and Madabhushi (2013); Bisoi and Haldar (2014); to name loading;
a few. . Combined (e.g., axial and lateral) cyclic loading.
The scarcity of field data for cyclic lateral loading of large
Qd ¼ Qf þ Qb ¼ fAsur þ qAb ð42Þ
diameter (rigid) piles available in the literature, particularly
where Qf is the shaft-friction capacity, Qb is the end-bearing for high load cycling, makes the validation of current and
capacity, f is the unit skin friction, Asur is the shaft area improved design methods=theories and the calibration of
pertaining to the pile embedment length, q is the unit end- numerical models for offshore monopiles difficult. Many
bearing capacity, and Ab is the base area of the plugged of the proposed models=formulations, with the principal
ones described earlier in the paper, were calibrated against
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

monopile.
For monopiles having a diameter of 4 m diameter or experimental data obtained from model-scale pile tests.
greater, the pile plug resistance is usually not taken into Further instrumented field testing of full-scale monopiles,
account in the calculations (van der Tempel 2006). Further, ideally having comparable size and geometry (slenderness
it has been shown that degradation of the shaft-friction ratio) with that of current and proposed OWT monopile
capacity due to cyclic axial loading leads to accumulating foundations, and subjected to high numbers of lateral load
displacements and potentially a severe reduction in the axial cycles, is warranted and would provide a valuable source
load-carrying capacity (Gavin and O’Kelly 2007), although of information in this regard for the wind-energy sector.
consideration of this degradation of the shaft friction in
the design process is still an open question. Numerous design
Summary and Conclusions
charts available in the literature make it possible to dis-
tinguish between stable and unstable loading levels to ensure Sources, types, and methods of analyses for the determi-
a design solution on the safe side. Further, no reduction in nation of the magnitude of the environmental loading and
axial load-carrying capacity is expected if a certain magni- resulting moments exerted on offshore monopiles, including
tude of the cyclic load amplitude is not exceeded (see Poulos for long-term and extreme conditions, are well documented
1988, and Abdel-Rahman and Achmus 2011, for further in the literature. However the behavior of the pile–soil sys-
details). tem in response to these cyclic (dynamic) loading scenarios
is not entirely clear. Further, the analysis=design of large
diameter (rigid) monopile foundations for current and pro-
Discussion posed OWT structures is well outside the scope of present
experience and analysis=design methods, which are mainly
There is no overall agreement in the literature regarding the based on experimental data obtained for relatively small-
determination of the pile’s rotation=lateral deformation diameter flexible piles subjected to low numbers of load
response to the many millions of low-amplitude lateral load cycles (N < 200). This includes the widely used API (2010)
cycles associated with OWT monopile foundation scenarios. and DNV (2011) standards developed primarily for the off-
Considerable differences of opinion exist on the rate of cyclic shore oil=gas industry.
strain accumulation; e.g., power function (Little and Briaud The behavior of large-diameter monopiles in general, and
1988) and logarithmic-trend relationships (Lin and Liao the long-term low-amplitude cyclic lateral loading response
1999) have been proposed. Most of the reported field and in particular, are not well documented. Existing models for
laboratory pile tests were performed for medium dense sand estimating the accumulated lateral strain (rotation) response
(Little and Briaud 1988; LeBlanc 2009; Bienen et al. 2012). of monopiles are based on very limited field=laboratory data
Compared with the Bienen et al. (2012) and LeBlanc and are currently not capable of explicitly accounting for
(2009) models which are based on N  104 lateral load cycles, site-specific soil properties and environmental loading char-
other widely used models for predicting the pile’s rotation= acteristics. There is also no consensus among researchers
lateral displacement response, including the API (2010), regarding the severity of strain accumulation due to one-
DNV (2011), Little and Briaud (1988) and Long and way and two-way loading scenarios, the effects of varying
Vanneste (1994) approaches, are based on experimental data the load direction and frequency or changes in soil stiffness
for relatively few load cycles (N < 200). Since no reliable under long-term stress application.
model presently exists, design requirements regarding the Instrumented field tests on full-scale ‘rigid’ monopiles,
pile’s rotation=lateral displacement behavior under mono- combined with a more extensive program of testing on
tonic (static) extreme load are used as a substitute. Areas model-scale piles installed in different soil conditions, con-
warranting further in-depth research are the effects on sidering changing load characteristics (amplitude, frequency,
monopile behavior of: and direction) and subjected to high numbers of load cycles
. Soil properties; e.g., relative density, over-consolidation of more than 106 are warranted. Such studies would provide
ratio, and the relative significance of the at-rest earth valuable information for the validation of current and
pressure coefficient (K0) and the soil stiffness, including improved design methods=theories for offshore monopiles
their variations with depth; and the calibration of pertinent numerical models. They
20 M. Arshad & B. C. O’Kelly

would also be helpful in generating computer code for Engineering 138(3): 364–75. doi:10.1061=(asce)gt.1943-
numerical simulations of more realistic conditions, especially 5606.0000592
in-situ soil conditions and in-service loading characteristics. Bisoi, S. and S. Haldar. 2014. Dynamic analysis of offshore wind tur-
bine in clay considering soil–monopile–tower interaction. Soil
Dynamics and Earthquake Engineering 63: 19–35. doi:10.1016=
Acknowledgment j.soildyn.2014.03.006
Blanco, M. I. 2009. The economics of wind energy. Renewable and Sus-
The first author gratefully acknowledges a Postgraduate tainable Energy Reviews 13(6–7): 1372–82.
Research Scholarship Award from Trinity College Dublin. Broms, B. 1964. Lateral resistance of piles in cohesionless soils. Soil
The writers also thank the reviewers for many helpful Mechanics and Foundation Engineering Division, ASCE 90(3):
comments. 123–56.
Budhu, M. 2011. Soil Mechanics and Foundations. New York: John
Wiley & Sons.
References Byrne, B. W. and G. T. Houlsby. 2003. Foundations for offshore wind
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

Abdel-Rahman, K. and M. Achmus. 2011. Behavior of foundation turbines. Philosophical Transactions of the Royal Society of London
piles for offshore wind energy plants under axial cyclic loading. 361(1813): 2909–30.
Proceedings of the SIMULIA Customer Conference, Barcelona, Carswell, W., S. R. Arwade, D. J. DeGroot, and M. A. Lackner. 2015.
Spain, Dassault Systèmes Press, Paris, France, 17–19th May Soil–structure reliability of offshore wind turbine monopile foun-
2011, 331–41. dations. Wind Energy 18(3): 483–98. doi:10.1002=we.1710
Achmus, M. 2010. Design of axially and laterally loaded piles for the Chai, J.-C. and N. Miura. 2002. Traffic-load-induced permanent defor-
support of offshore wind energy converters. Proceedings of the mation of road on soft subsoil. Journal of Geotechnical and Geoen-
Indian Geotechnical Conference GEOtrendz–2010, Mumbai, vironmental Engineering 128(11): 907–16. doi:10.1061=(asce)1090-
India, December 16–18th, 2010, 92–102. 0241(2002)128:11(907)
Achmus, M., Y.-S. Kuo, and K. Abdel-Rahman. 2009. Behaviour of Chakrabarti, S. K. 2005. Handbook of Offshore Engineering. London,
monopile foundations under cyclic lateral load. Computer and UK: Elsevier.
Geotechnics 36(5): 725–35. doi:10.1016=j.compgeo.2008.12.003 Das, B. M. 2008. Advanced Soil Mechanics. New York: CRC Press.
API (American Petroleum Institute). 2010. Recommended practice for Davisson, M. T. 1970. Lateral Load Capacity of Piles, Vol. 333, 104–12.
planning, designing and constructing fixed offshore platforms— Washington, DC: Highway Research Record.
Working stress design. API RP 2A-WSD (R2010), 22nd ed., De Vries, W. E. 2007. Project UpWind WP4 deliverable D4.2.1: Assess-
API Publishing Services, Washington, DC, USA. ment of bottom-mounted support structure types with conven-
Arshad, M. and B. C. O’Kelly. 2013. Offshore wind-turbine structures: tional design stiffness and installation techniques for typical
A review. Proceedings of the ICE, Energy 166(4): 139–52. deep water sites. http://www.upwind.eu/Publications/~/media/
doi:10.1680=ener.12.00019 UpWind/Documents/Publications/4%20-%20Offshore%20Foun-
Arshad, M. and B. C. O’Kelly. 2014. Development of a rig to study dations/UpwindWP4D421%20Assessment%20of%20bottom-
model pile behaviour under repeating lateral loads. International mounted%20support%20structure%20types.ashx (accessed March
Journal of Physical Modelling in Geotechnics 14(3): 54–67. 15th, 2015).
doi:10.1680=ijpmg.13.00015 De Vries, W. E. and J. van der Tempel. 2007. Quick monopile design.
Atkinson, J. H. and G. Sallfors. 1991. Experimental determination of Proceedings of the European Offshore Wind Conference & Exhi-
stress–strain–time characteristics in laboratory and in-situ tests. bition, Berlin, Germany, December 4–6th, 2007.
Proceedings of the 10th European Conference on Soil Mechanics DIN (Deutsches Institut fur Normung). 2005. Baugrund Sicherheits-
and Foundation Engineering, Florence, Italy. Balkema: Rotter- nachweise im Erd und Grundbau. DIN 1054: 2005. DIN: Berlin,
dam, The Netherlands, May 26–30th, 1991, Vol. 3, 915–56. Germany. http://www.baunormenlexikon.de/Normen/DIN/
Barksdale, R. D. 1972. Laboratory evaluation of rutting in base course DIN%201054/1b69b621-74a0-4c33-8b29-a82a24afd4d4 (accessed
materials. Proceedings of the 3rd International Conference on the March 15th, 2015).
Structural Design of Asphalt Pavement, London, UK, September DNV (Det Norske Veritas). 2011. Design of offshore wind turbine
11–15th, 1972, Vol. 1, 161–74. structures. DNV-OS-J101, DNV, Oslo, Norway.
Bartolomey, A. A. 1977. Experimental analysis of pile groups under lat- Dobry, R., E. Vicenti, M. J. O’Rourke, and J. M. Roesset. 1982. Hori-
eral loads. Proceedings of the 9th International Conference on Soil zontal stiffness and damping of single piles. Journal of Geotechni-
Mechanics and Foundation Engineering, Tokyo, Japan, July 10– cal Engineering Division, ASCE 108(3): 439–59.
15th, 1977, 187–88. Dührkop, J. 2009. On the influence of expanders and cyclic loads on the
Basack, S. and S. Dey. 2012. Influence of relative pile-soil stiffness and deformation behavior of lateral stressed piles in sand. PhD Thesis,
load eccentricity on single pile response in sand under lateral cyclic Hamburg University of Technology.
loading. Geotechnical and Geological Engineering 30(4): 737–51. Dyson, G. J. and M. F. Randolph. 2001. Monotonic lateral loading of
doi:10.1007=s10706-011-9490-1 piles in calcareous sand. Journal of Geotechnical and Geoenviron-
Basack, S. and S. Sen. 2014. Numerical solution of single piles mental Engineering 127(4): 346–52. doi:10.1061=(asce)1090-
subjected to pure torsion. Journal of Geotechnical and Geoenviron- 0241(2001)127:4(346)
mental Engineering 140(1): 74–90. doi:10.1061=(asce)gt.1943-5606. EWEA (European Wind Energy Association). 2011. Design limits and
0000964 solutions for very large wind turbines. UpWind project. http://
Batchelor, G. K. 1967. An Introduction to Fluid Dynamics. Cambridge, www.ewea.org/fileadmin/ewea_documents/documents/upwind/
UK: Cambridge University Press. 21895_UpWind_Report_low_web.pdf (accessed March, 15th 2015).
Bhattacharya, S., J. A. Cox, D. Lombardi, and D. M. Wood. 2012. Fischer, T. 2011. Executive summary – Upwind project WP4: Offshore
Dynamics of offshore wind turbines supported on two founda- foundations and support structures. http://www.upwind.eu/Pub-
tions. Proceedings of the ICE, Geotechnical Engineering 166(2): lications/~/media/UpWind/Documents/Publications/4%20-
159–69. doi:10.1680=geng.11.00015 %20Offshore%20Foundations/WP4_Executive_Summary_Fina-
Bienen, B., J. Dührkop, J. Grabe, M. F. Randolph, and D. J. White. l.ashx (accessed March, 15th 2015).
2012. Response of piles with wings to monotonic and cyclic lateral Gavin, K. G. and O’Kelly, B. C. 2007. Effect of friction fatigue
loading in sand. Journal of Geotechnical and Geoenvironmental on pile capacity in dense sand. Journal of Geotechnical and
Analysis and Design of Offshore Monopile Foundations 21
Geoenvironmental Engineering 133(1): 63–71. doi:10.1061= development. Technical Report NREL=TP-500–38060, National
(asce)1090-0241(2007)133:1(63) Renewable Energy Laboratory (NREL), Colorado, USA.
GL (Germanischer Lloyd). 2005. Guideline for the Certification of Off- Journée, J. M. J. and W. W. Massie. 2001. Offshore Hydrodynamics, 1st
shore Wind Turbines. Hamburg, Germany: GL. ed. Delft, The Netherlands: Delft University of Technology.
Goryunov, B. F. 1973. Analysis of piles subjected to the combined Karg, C. 2007. Modelling of strain accumulation due to low level vibrations
action of vertical and horizontal loads (discussion). Soil Mechanics in granular soils. PhD Thesis. Ghent University, Ghent, Belgium.
and Foundation Engineering 10(1): 10–13. doi:10.1007=bf01706631 Karthigeyan, S., V. V. G. S. T. Ramakrishna, and K. Rajagopal. 2006.
Guo, W. D. 2006. On limiting force profile, slip depth and response of Influence of vertical load on the lateral response of piles in sand.
lateral piles. Computer and Geotechnics 33(1): 47–67. doi:10.1016= Computers and Geotechnics 33(2): 121–31. doi:10.1016=j.comp-
j.compgeo.2006.02.001 geo.2005.12.002
Guo, W. D. 2013. Pu is subscripted-based solutions for slope stabilizing Karthigeyan, S., V. V. G. S. T. Ramakrishna, and K. Rajagopal. 2007.
piles. International Journal of Geomechanics 13(3): 292–310. Numerical investigation of the effect of vertical load on the lateral
doi:10.1061=(asce)gm.1943-5622.0000201 response of piles. Journal of Geotechnical and Geoenvironmental
Haiderali, A., U. Cilingir, and G. Madabhushi. 2013. Lateral and axial Engineering 133(5): 512–21. doi:10.1061=(asce)1090-0241(2007)
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

capacity of monopiles for offshore wind turbines. Indian Geotech- 133:5(512)


nical Journal 43(3): 181–94. doi:10.1007=s40098-013-0056-4 Klinkvort, R. T. and O. Hededal. 2013. Lateral response of monopile
Haritos, N. 2007. Introduction to the analysis and design of offshore supporting an offshore wind turbine. Proceedings of the ICE, Geo-
structures – An overview. Electronic Journal of Structural technical Engineering 166(2): 147–58. doi:10.1680=geng.12.00033
Engineering 7(Special Issue: Loading on Structures): 55–65. Kopp, D. R. 2010. Foundations for an offshore wind turbine. MSc
Hornych, P., J. F. Corte, and J. L. Paute. 1993. Etude des déformations Thesis, Massachusetts Institute of Technology.
permanentes sous chargements répétés de trois graves non traitées. Kuo, Y.-S., M. Achmus, and K. Abdel-Rahman. 2012. Minimum
Bulletin de Liaison des Laboratoires des Ponts et Chaussées, Presse embedded length of cyclic horizontally loaded monopiles. Journal
de l’ENPC, Paris, France 184: 45–55. of Geotechnical and Geoenvironmental Engineering 138(3): 357–63.
Huurman, M. 1996. Development of traffic induced permanent strains doi:10.1061=(asce)gt.1943-5606.0000602
in concrete block pavements. Heron 41(1): 29–52. LeBlanc, C. 2009. Design of offshore wind turbine support structures:
IEC (International Electrotechnical Commission). 2009. Wind turbines Selected topics in the field of geotechnical engineering. PhD The-
– Part 3: Design requirements for offshore wind turbines. IEC sis, Aalborg University, Aalborg, Denmark.
61400–3: 2009, International Electrotechnical Commission, Gen- LeBlanc, C., G. T. Houlsby, and B. W. Byrne. 2010. Response of stiff
eva, Switzerland. piles in sand to long-term cyclic lateral loading. Géotechnique
Igoe, D., K. Gavin, and B. O’Kelly. 2010. Field tests using an instru- 60(2): 79–90. doi:10.1680=geot.7.00196
mented model pipe pile in sand. Proceedings of the Seventh Inter- Lekarp, F. and A. Dawson. 1998. Modelling permanent deformation
national Conference on Physical Modelling in Geotechnics, behaviour of unbound granular materials. Construction and Build-
Zurich, Switzerland, June 28th–July 1st, 2010, ed. S. Springman, ing Materials 12(1): 9–18. doi:10.1016=s0950-0618(97)00078-0
J. Laue, and L. Seward, Vol. 2, 775–780. Leiden, The Netherlands: Lentz, R. W. and G. Y. Baladi. 1981. Constitutive equation for perma-
CRC Press. nent strain of sand subjected to cyclic loading. Transportation
Igoe, D., K. Gavin and B. O’Kelly. 2013. An investigation into the use Research Record 810: 50–54.
of push-in pile foundations by the offshore wind sector. Inter- Lesny, K. and J. Wiemann. 2005. Design aspects of monopiles in Ger-
national Journal of Environmental Studies 70(5): 777–91. man offshore wind farms. Proceedings of the 1st International
doi:10.1080=00207233.2013.798496 Symposium on Frontiers in Offshore Geotechnics. Perth, Austra-
Igoe, D. J. P., K. G. Gavin, B. C. O’Kelly, and B. Byrne. 2013. The use lia, September 19–21st, 2005, ed. S. Gourvenec and M. Cassidy,
of in-situ site investigation techniques for the axial design of off- 383–89. Leiden, The Netherlands: Balkema.
shore piles. Proceedings of the 4th International Conference on Li, D. and E. Selig. 1996. Cumulative plastic deformation for fine-
Geotechnical and Geophysical Site Characterization, Pernam- grained subgrade soils. Journal of Geotechnical Engineering, ASCE
buco, Brazil, September 18–21st, 2012, ed. R. Q. Coutinho and 122(12): 1006–13. doi:10.1061=(asce)0733-9410(1996)122:12(1006)
P. W. Mayne. Vol. 2, 1123–29. CRC Press. Li, Z., S. K. Haigh, and M. D. Bolton. 2010. Centrifuge modelling of
Irvine, J. H., P. G. Allan, B. G. Clarke, and J. R. Peng. 2003. Improv- mono-pile under cyclic lateral loads. Proceedings of the 7th Inter-
ing the lateral stability of monopile foundations. Proceedings of national Conference on Physical Modelling in Geotechnics, Zur-
the BGA International Conference on Foundations: Innovations, ich, Switzerland, 28th June–1st July 2010, ed. S. Springman, J.
Observations, Design and Practice, Dundee, UK, September 2– Laue, and L. Seward, Vol. 2, 965–70. Leiden, The Netherlands:
5th, 2003, ed. T. A. Newson, 371–80. London: Thomas Telford. CRC Press.
Ismael, N. F. 1990. Behavior of laterally loaded bored piles in cemented Lin, S.-S. and J.-C. Liao. 1999. Permanent strains of piles in sand due
sands. Journal of Geotechnical Engineering, ASCE 116(11): 1678– to cyclic lateral loads. Journal of Geotechnical and Geoenvironmen-
99. doi:10.1061=(asce)0733-9410(1990)116:11(1678) tal Engineering 125(9): 798–802. doi:10.1061=(asce)1090-
Jaimes, O. G. 2010. Design concepts for offshore wind turbines: 0241(1999)125:9(798)
A technical and economic study on the trade-off between stall Little, R. L. and J.-L. Briaud. 1988. Full scale cyclic lateral load tests on
and pitch controlled systems. MSc Thesis, Delft University of six single piles in sand. Miscellaneous paper GL-88–27, Geotech-
Technology. nical Division, Civil Engineering Department, Texas A&M Uni-
Jain, N. K., G. Ranjan, and G. Ramasamy. 1987. Effect of vertical load versity, College Station, TX, USA.
on flexural behaviour of piles. Geotechnical Engineering: Journal of Lombardi, D., S. Bhattacharya, and D. M. Wood. 2013. Dynamic soil–
the Southeast Asian Geotechnical Society 18(2): 185–204. structure interaction of monopile supported wind turbines in
Jang, J. J. and G. J. Shinn. 1999. Analysis of maximum wind force for cohesive soil. Soil Dynamics and Earthquake Engineering 49:
offshore structure design. Journal of Marine Science and Tech- 165–80. doi:10.1016=j.soildyn.2013.01.015
nology 7(1): 43–51. Long, J. H. and G. Vanneste. 1994. Effects of cyclic lateral loads on
Jonkman, J., S. Butterfield, W. Musial, and G. Scott. 2009. Definition piles in sand. Journal of Geotechnical Engineering, ASCE 120(1):
of a 5-MW reference wind turbine for offshore system 225–44. doi:10.1061=(asce)0733-9410(1994)120:1(225)
22 M. Arshad & B. C. O’Kelly

Malhotra, S. 2011. Design and construction considerations for offshore Peng, J., B. G. Clarke, and M. Rouainia. 2011. Increasing the resistance
wind turbine foundations in North America. Proceedings Geo- of piles subject to cyclic lateral loading. Journal of Geotechnical
Florida 2010: Advances in Analysis, Modeling and Design, West and Geoenvironmental Engineering 137(10): 977–82. doi:10.1061=
Palm Beach, Florida, USA, February 20–24th, 2010, ed. D. O. (asce)gt.1943-5606.0000504
Fratta, A. J. Puppala, and B. Muhunthan, Vol. 2, 1533–42, GSP Peralta, P. and M. Achmus. 2010. An experimental investigation of
199. Red Hook, NY, USA: Curran Associates, Inc. piles in sand subjected to lateral cyclic loads. Proceedings of the
Matlock, H. 1970. Correlation for design of laterally loaded piles in soft 7th International Conference on Physical Modelling in Geotech-
clay. Proceedings of the 2nd Offshore Technology Conference, 22– nics, Zurich, Switzerland, 28th June–1st July 2010, ed. S. Spring-
24th April 1970, Houston, TX, USA. Paper Number 1204. Vol. 1, man, J. Laue, and L. Seward, Vol. 2, 985–90. Leiden, The
77–94. Netherlands: CRC Press.
Matlock, H., and L. C. Reese 1960. Generalized solutions for laterally Poulos, H. G. 1988. Cyclic stability diagram for axially loaded piles.
loaded piles. Journal of the Soil Mechanics and Foundations Journal of Geotechnical Engineering, ASCE, 114(8): 877–95.
Division ASCE 86(5): 63–94. doi:10.1061=(asce)0733-9410(1988)114:8(877)
Moayed, R. Z., I. Mehdipour, and A. Judi. 2012. Undrained lateral Rahim, A. and R. F. Stevens. 2013. Design procedures for marine
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

behavior of short pile under combination of axial, lateral and renewable energy foundations. Proceedings of the 1st Marine
moment loading in clayey soils. Kuwait Journal of Science and Energy Technology Symposium, Washington, DC, April 10–
Engineering 39(1B): 59–78. 11th, 2013, 10.
Morison, J. R., J. W. Johnson, and S. A. Schaff. 1950. The forces Ramasamy, G. 1974. Flexural behaviour of axially and laterally loaded
exerted by surface waves on piles. Journal of Petroleum Technology individual piles and groups of piles. PhD Thesis, Indian Institute
2(5): 149–54. doi:10.2118=950149-g of Science, Bangalore.
Naughton, P. J. and B. C. O’Kelly. 2004. The induced anisotropy of Randolph, M. F. 1981. The response of flexible piles to lateral loading.
Leighton Buzzard sand. Proceedings Advances in Geotechnical Géotechnique 31(2): 247–59. doi:10.1680=geot.1981.31.2.247
Engineering: The Skempton Conference, London, UK, March Rani, S. and A. Prashant. 2014. Estimation of the linear spring con-
28th–31st, 2004, ed. R. J. Jardine, D. M. Potts, and K. G. Higgins, stant for a laterally loaded monopile embedded in nonlinear soil.
Vol. 1, 556–67. London, UK: Thomas Telford. International Journal of Geomechanics. doi:10.1061=
Naughton, P. J. and B. C. O’Kelly. 2005. Yield behavior of sand under (ASCE)GM.1943-5622.0000441. Online Publication Date: 29th
generalized stress conditions. Proceedings of the 16th International August 2014.
Conference on Soil Mechanics and Geotechnical Engineering, Osaka, Reese, L. C, W. R. Cox, and F. D. Koop. (1974). Analysis of laterally
Japan, September 12–16th, 2005. IOS Press. Vol. 2, 555–58. loaded piles in sand. Proceedings of the 6th Annual Offshore Tech-
Nicolai, G. and L. B. Ibsen. 2014. Small-scale testing of cyclic laterally nology Conference, Houston, Texas, May 6–8th, 1974, 473–84.
loaded monopiles in dense saturated sand. Journal of Ocean and Reese, L. C, W. R. Cox, and F. D. Koop. (1975). Field testing
Wind Energy 1(4): 240–45. and analysis of laterally loaded piles in stiff clay. Procedings
Niemunis, A., T. Wichtmann, and Th. Triantafyllidis. 2005. A high- of the 7th Annual Offshore Technology Conference. Paper No.
cycle accumulation model for sand. Computers and Geotechnics 2312, Houston, Texas, May 5–8th, 1975, 671–690. doi: 10.4043/
32(4): 245–63. doi:10.1016=j.compgeo.2005.03.002 2312-MS.
O’Kelly, B. C. 2005. Consolidation anisotropy of some natural soft Richwien, W., K. Lesny, and J. Wiemann. 2002. Bau- und umwelttech-
soils. Proceedings of the International Conference on Problematic nische Aspekte von Off-shore Windenergieanlagen (Construction
Soils, Famagusta, North Cyprus, May 25–27th, 2005, ed. H. Bilsel and environmental aspects of offshore wind turbines). Gigawind
and N. Zalihe, Vol. 3, 1183–92. North Cyprus: Eastern Mediterra- Annual Report 2001, Chapter 6: Support structure – Foundation,
nean University Press. 36–46. Leibniz Universität Hannover, Germany.
O’Kelly, B. C. 2006. Compression and consolidation anisotropy of Şahin, A. D. 2004. Progress and recent trends in wind energy. Progress
some soft soils. Geotechnical and Geological Engineering 24(6): in Energy and Combustion Science 30(5): 501–43. doi:10.1016=
1715–28. doi:10.1007=s10706-005-5760-0 j.pecs.2004.04.001
O’Kelly, B. C. and P. J. Naughton. 2005a. Development of a new hol- Sorochan, E. A. and V. I. Bykov. 1976. Performance of groups of
low cylinder apparatus for stress path measurements over a wide cast-in place piles subjected to horizontal loading. Soil Mechanics
strain range. Geotechnical Testing Journal 28(4): 345–54. and Foundation Engineering 13(3): 157–61. doi:10.1007=
doi:10.1520=gtj12252 bf01705310
O’Kelly, B. C. and P. J. Naughton. 2005b. Engineering properties of St. Denis, M. and W. J. Pierson. 1953. On the motions of ships in con-
wet-pluviated hollow cylindrical specimens. Geotechnical Testing fused seas. Society of Naval Architects and Marine Engineers
Journal 28(6): 570–76. doi:10.1520=gtj12325 Transactions 61: 280–354.
O’Kelly, B. C. and P. J. Naughton. 2008. Local measurements of the Sweere, G. T. H. 1990. Unbound granular bases for roads. PhD Thesis,
polar deformation response in a hollow cylinder apparatus. Geo- Delft University of Technology.
mechanics and Geoengineering 3(4): 217–29. doi:10.1080= Terzaghi, K. 1955. Evaluation of coefficients of subgrade reaction.
17486020802400981 Géotechnique 5(4): 297–326.
O’Kelly, B. C. and P. J. Naughton. 2009. Study of the yielding of sand Timmerman, D. H. and T. H. Wu. 1969. Behavior of dry sands under
under generalized stress conditions using a versatile hollow cylin- cyclic loading. Journal of the Soil Mechanics and Foundations
der torsional apparatus. Mechanics of Materials 41(3): 187–98. Division ASCE 95(4): 1097–114.
doi:10.1016=j.mechmat.2008.11.002 Tomlinson, M. J. 2001. Foundation Design and Construction, 7th edn.
Pappin, J. W. 1979. Characteristics of granular material for pavement Harlow, England: Pearson Education.
analysis. PhD Thesis, University of Nottingham. Tong, W. 2010. Wind Power Generation and Wind Turbine Design.
Paute, J.-L., P. Hornych, and J. P. Benaben. 1996. Repeated load triax- Southampton, UK: WIT Press.
ial testing of granular materials in the French Network of Labora- van der Tempel, J. 2006. Design of support structures for offshore wind
tories des Ponts et Chaussées. Flexible Pavements: Proceedings of turbines. PhD Thesis, Delft University of Technology.
the European Symposium Euroflex, Lisbon, Portugal, September Verdure, L., J. Garnier, and D. Levacher. 2003. Lateral cyclic loading
20–22nd, 1993, ed. A. G. Corriea, 53–64. Rotterdam, The of single piles in sand. International Journal of Physical Modelling
Netherlands: Balkema. in Geotechnics 3(3): 17–28.
Analysis and Design of Offshore Monopile Foundations 23
Wang, Q. and P. V. Lade. 2001. Shear banding in true triaxial tests and Appendix
its effect on failure in sand. Journal of Engineering Mechanics
127(8): 754–61. doi:10.1061=(asce)0733-9399(2001)127:8(754)
Werkmeister, S. 2003. Permanent deformation behaviour of unbound
granular materials in pavement constructions. PhD Thesis, Tech-
nical University Dresden.
Wichtmann, T., A. Niemunis, and Th. Triantafyllidis. 2009. Validation
and calibration of a high-cycle accumulation model based on cyc-
lic triaxial tests on eight sands. Soils and Foundations 49(5): 711–
28. doi:10.3208=sandf.49.711
Wichtmann, T., H. A. Rondón, A. Niemunis, Th. Triantafyllidis, and
A. Lizcano. 2010. Prediction of permanent deformations in pave-
ments using a high-cycle accumulation model. Journal of Geotech-
nical and Geoenvironmental Engineering 136(5): 728–40.
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

doi:10.1061=(asce)gt.1943-5606.0000275
Yasin, S. J. M. and F. Tatsuoka. 2000. Stress history-dependent defor-
mation characteristics of dense sand in plane strain. Soils and
Foundations 40(2): 77–98. doi:10.3208=sandf.40.2_77
Zhu, B., B. W. Byrne, and G. T. Houlsby. 2013. Long-term lateral cyc-
lic response of suction caisson foundations in sand. Journal of
Geotechnical and Geoenvironmental Engineering 139(1): 73–83.
doi:10.1061=(asce)gt.1943-5606.0000738
Zhukov, N. V. and I. L. Balov. 1978. Investigation of the effect of a
vertical surcharge on horizontal displacements and resistance of Fig. A1. Pertinent stiffness profiles for the medium dense sand
pile columns to horizontal loads. Soil Mechanics and Foundation deposit considered in the OWT monopile foundation design
Engineering 15(1): 16–22. doi:10.1007=bf02145324 example.

You might also like