Sei sulla pagina 1di 24

Marine Georesources & Geotechnology

ISSN: 1064-119X (Print) 1521-0618 (Online) Journal homepage: http://www.tandfonline.com/loi/umgt20

Analysis and Design of Monopile Foundations for


Offshore Wind-Turbine Structures
Muhammad Arshad & Brendan C. O'Kelly
To cite this article: Muhammad Arshad & Brendan C. O'Kelly (2015): Analysis and Design
of Monopile Foundations for Offshore Wind-Turbine Structures, Marine Georesources &
Geotechnology, DOI: 10.1080/1064119X.2015.1033070
To link to this article: http://dx.doi.org/10.1080/1064119X.2015.1033070

Accepted author version posted online: 28


Jul 2015.
Published online: 28 Jul 2015.
Submit your article to this journal

Article views: 162

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=umgt20
Download by: [University of Engineering & Technology Lahore]

Date: 03 February 2016, At: 01:19

Marine Georesources & Geotechnology, 0: 123


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. OKELLY1
1

Department of Civil, Structural and Environmental Engineering, Trinity College Dublin, Dublin, Ireland
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-environment loading conditions, including methods of calculation using site-specic data, are described along with wave particle kinematics, 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
pilesoil 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 monopile foundation system is presented. Target areas for further research by the wind-energy sector, which would facilitate the development of improved analyses=design methods for offshore monopiles, are identied.
Keywords: foundation, lateral load, monopile, ocean environment, soil, strain accumulation

Introduction
There has been a rapid growth in the use of offshore (and
onshore) wind farms for the production of clean and renewable energy. The economic development of wind farms
depends on efcient solutions being available for a number
of technical issues, one aspect being the foundations. Offshore wind-turbine (OWT) structures may be found on gravity base, suction caisson, monopile, tripod or braced frame
(jacket) foundations or, more recently, using oating platforms tethered to the seabed (Figure 1). The foundation
choice largely depends on the water depth, seabed characteristics, loading characteristics, available construction technologies, and importantly, economic costs (Malhotra 2011).
Offshore foundations are subjected to a combination of
loads, namely the axial (self-weight) forces of the
structure=machinery (V), repeating horizontal=lateral loads
(H), and bending (M) and torsional moments. Apart from
the self-weight, these loads are generated by environmental
conditions (ocean waves, currents, tidal action, and aerodynamic load cycles) and=or operation of the installed machinery (Figure 2). The lateral loading is essentially cyclic in
nature and typically of low amplitude. For geotechnical
Address correspondence to Muhammad Arshad, Department
of Geological Engineering, University of Engineering & Technology, GT Road Lahore, Pakistan. E-mail: arshadm@tcd.ie

design, the proportion and importance of the different loading types essentially depend upon the kind of foundation
system being considered. For gravity-base foundations
(Figure 1a), potential failure modes may be in bearing
capacity or excess settlement; hence, the vertical load is generally the major design consideration (Malhotra 2011). For
monopile foundations (Figure 1b), lateral deection
(rotation) of the monopile controls the serviceability limit
state of the whole structure; hence, lateral loads are more
critical when compared with the vertical loads.
In general, offshore foundations must be designed to
resist large numbers of aerodynamic and hydrodynamic load
cycles of varying direction, amplitude, and frequency
(Figure 2) at the proposed site over the projects lifetime of
typically 25 years or more (S ahin 2004). In most parts
of the world, the most popular foundation choice in terms
of ease of installation, economy, and logistics is the monopile, with an estimated 75% of all installed OWTs using this
solution (Blanco 2009; Fischer 2011; Malhotra 2011). For
OWT monopile foundation systems, the lateral loads (and
resulting moments) are generally large in proportion to the
axial loads. Hence, the foundation response under repeating
lateral loading is a major consideration, with the foundation
design dominated by considerations of the dynamic and fatigue responses under working loads, rather than the ultimate
load-carrying capacity. The overall feasibility of a wind-farm
project is invariably determined by upfront capital costs,

Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

M. Arshad & B. C. OKelly

Fig. 2. Environmental impacts on offshore wind turbine.


Fig. 1. Support structure options showing their range of applicable water depths.

with the material, fabrication, and installation costs for


OWT foundations, typically accounting for 25% of the
overall project cost (Blanco 2009). This signicant investment demands reliable foundation designs to achieve the
overall goals of pollution-free wind-generated power at
economical rates.
The aims of this review article are to present:
. Details on the geometric design and structural and

.
.

environmental loading for existing and planned OWTs


supported by monopile foundations.
Describe ocean-environment loading conditions, including
methods of calculation using site-specic data.
Discuss wave particle kinematics, focusing on correlations
between the loading frequency and natural vibration frequency of OWT structures.
Critically review existing semi-empirical methods for modeling soil under cyclic loading, focusing on cyclic design
for fatigue life. The inherent limitations=shortcomings of
these models for the analyses=design of existing and
planned OWT monopile foundations are identied along
with further research needs.
Present a typical example of the geometric design for an
OWT-monopile foundation system.

Other aspects related to offshore monopile structures


(e.g., construction materials, manufacturing advancements,
corrosion assessment, and control) are beyond the scope of
this article.

General Aspects of OWT-Monopile Foundation


System
A monopile is a single large-diameter steel tube that penetrates into the seabed. The required monopile length and
diameter are primarily dependent on the OWTs power

generation capacity, which is an indirect measure of the


applied loading. Other signicant inuencing factors include
the soil condition and severity of environmental loading.
Existing offshore monopiles are typically 57.5 m in diameter (Achmus, Kuo, and Abdel-Rahman 2009), with slenderness [embedded length (L) to outer diameter (D)] ratio
values of <10, and are; therefore, considered to behave as
rigid structures for which rotation is prominent over bending
(Tomlinson 2001; Peng, Clarke, and Rouainia 2011). For a
typical OWT having a rated power of 5 MW and installed
in the North Sea, a hub height (see Figure 3) of 95 m above
the mean sea level and rotor diameter of 125 m lead to the
approximate quasi-static loading scenario summarized in
Table 1 (Lesny and Wiemann 2005; IEC [International
Electrotechnical Commission] 2009) and which acts at the
seabed level. This example demonstrates that torsional
moments are negligibly small, whereas lateral loading causes
extremely high bending moments and principally controls
the foundation design.
The monopile foundation transfers the horizontal forces
and bending moments generated by wind and wave loading
(via horizontal earth pressure acting along the pile) into the
surrounding ground through cantilever action. Factors
affecting the cyclic response include the monopile size and
embedment length, soil properties, soilpile relative stiffness,
loading characteristics, and pile installation method (Malhotra 2011). Depending on subsoil conditions, the monopile is
typically installed into the seabed by drilling and grouting,
driving using large impact or vibratory hammers, or a combination of drilling and driving (Malhotra 2011).
In current practice, OWT monopile foundations are
usually designed using general geotechnical standards in
combination with more specic guidelines and semiempirical formulas that have been largely developed by the
offshore oil=gas industry. The latter are usually based on
limited eld data obtained for piles having signicantly
smaller diameters (DIN [Deutsches Institut fur Normung]

Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

Analysis and Design of Offshore Monopile Foundations

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 foundation 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 specically consider simulation of the soil
pile interaction, instead focusing on the soils strength and
stiffness responses. The most popular method of analysis for
predicting the pile deection 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 eld 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 deections (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

Fig. 3. Main components of offshore wind turbine system.

2005; GL [Germanischer Lloyd] 2005; API [American


Petroleum Institute] 2010; DNV [Det Norske Veritas]
2011). Hence, extrapolation of these formulations for
present and future design of OWT monopile foundations
requires careful consideration of the applied loading and
the inherent limitations underlying such approaches.
Another signicant shortcoming is that current design
standards and guidelines on the serviceability of piles under
cyclic lateral loading are limited. For instance, over its service life, a typical 2 MW OWT structure is usually subjected
to 102 and 107 cycles of lateral loading with magnitudes of
Table 1. Loading at seabed level for 5 MW wind turbine
supported by monopile foundation (Lesny and Wiemann 2005)
Component

Magnitude

Vertical load
Horizontal load
Bending moment
Torsional moment

35 MN
16 MN
562 MN  m
4 MN  m

The main components of the wind-turbine system are its


foundation, support structure, transition piece, tower, rotor
blades, and nacelle (Figure 3). Together, the support structure and foundation maintain the turbine in its proper operational position. The transition piece provides a means of
correcting any misalignment of the monopile that may arise
during its installation. In some instances, the monopile can
extend from the seabed level to above the water surface level,
connecting with the transition piece or tower (Arshad and
OKelly 2013). The nacelle (hub) contains the key electromechanical components of the wind turbine, including the
gearbox and generator (Malhotra 2011). Rotor diameters
of existing OWTs typically range 90120 m, producing
power-generation capacities of 36 MW respectively (Tong
2010). Gravitational loading typically ranges 28 MN. For
instance, Byrne and Houlsby (2003) reported vertical load
of 6 MN for an anticipated 3.5 MW OWT in the UK region,
with lateral loading from aerodynamic and hydrodynamic
factors accounting for up to 67% of the vertical loading.
These loads will vary with the size of the installation, the
detailed design, and local environmental conditions. OWTs
having rotor diameters of 250 m and a power-generation
capacity of 20 MW are currently in the research and development phase (EWEA [European Wind Energy Association]

Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

4
2011), creating even more onerous design loading scenarios
for (monopile) foundations.
OWT monopile foundations are typically manufactured
from steel tubular sections having outer diameters of up to
7.5 m, wall thicknesses of up to 150 mm, and embedment
(penetration) depths of between 15 and 30 m (Achmus,
Kuo, and Abdel-Rahman 2009). Monopiles are generally
used in shallow water depths (i.e., typically <30 m), generally becoming too exible for water depths between 30
and 40 m, in which case, monopiles tted with guy wires or
tripod solutions (see Figure 1) are considered as economical
alternatives. For greater depths, time-consuming installation
and the effects of soil degradation (potholing), which
occurs at seabed level around the pile, make monopile foundation solutions prohibitive (Irvine et al. 2003). Other foundation options, as illustrated in Figure 1, are then considered
as viable options. The serviceability limit state is largely
determined by the lateral deection (rotation) of the monopile under many millions of load cycles; e.g., 107 lateral load
cycles of 1.4 MN magnitude (corresponding to the fatigue
loading for design) are expected to occur over the service life
of OWT structures (GL 2005).

M. Arshad & B. C. OKelly


 a
Z
U z U 10
10
U z U 10

ln Z=Z0
ln 10=Z0

1
2

where the values of exponent a and the roughness parameter,


Z0, are dependent on the site location, with typical values of
0.11 and 0.0001, respectively, for open-sea environments.
The wind velocity uctuates with time for a given height,
and also with change in height, above the sea level or ground
surface (Figure 4). However, a mean value (U z) can be established for a particular site location provided sufcient wind
speed data are available. The instantaneous wind velocity
can be considered as the superposition of the turbulent
component (wz) on the mean value. Hence the total aerodynamic drag force (FDWind) acting on an offshore structure
can be determined by (Jang and Shinn 1999; API 2010):
 2


FDWind 0:5qair Cdair Aw U z qair Cdair Aw U z wz
0:5qair Cdair Aw jwz jwz

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 loading) 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 assessments of the OWT structures fatigue life.

Aerodynamic Loading
Wind conditions are important in dening, not only the
loads imposed on a wind turbines 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 distribution, 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
law given by:

Fig. 4. Fluctuation of wind speed about the mean value. (a) In


time domain. (b) In three-dimensional space.

Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

Analysis and Design of Offshore Monopile Foundations


where qair is the density of air, Aw is the exposed area of the
offshore structure, Cdair is an aerodynamic drag coefcient
that is dependent on the structures 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 rst term on the right-hand side of Eq. (3) represents
the mean drag force, with the second and third terms denoting the drag force components associated with wind velocity
uctuations.
Wind loads on offshore structures are seldom greater than
the hydrodynamic loads and often play a relatively minor
role, compared with the hydrodynamic loads. However wind
loading has a longer lever arm when considering moments
generated about the structures 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 calculation 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 compared with variations in the wave period. Fluctuations about
the mean wind speed impose aerodynamic forces, although
when considering the dynamic behavior of offshore structures, these forces are also generally insignicant compared
with the hydrodynamic forces (Journee and Massie 2001).
Hence wind speed is generally considered as steady (i.e., without uctuation), resulting in constant force(s) and; hence,
constant moment(s). Dynamic analysis of offshore structures
is necessary in cases where the wind eld contains energy at
frequencies near the offshore structures natural frequencies
of vibration (API 2010), although for monopile foundations,
the difference between these frequencies is usually high [see
later section on Correlation between Loading Frequency
and Natural Frequency of OWT Structures]. Some statistical
manipulations can be applied to the wind data (e.g., refer to
Jang and Shinn 1999; Haritos 2007) for the inclusion of
dynamic forces due to wind velocity turbulence.
Hydrodynamic Loading
For OWT monopile foundations, the dominant hydrodynamic loads are generated by ocean waves and currents, with
minor contributions arising from other sources such as sea
level variations due to tides or swell. The main sources of
wave generation are wind and astronomical forces. Earthquakes, submarine landslides, and shipping or other nearby
oating structures may also be important sources of wave
generation. Irregular waves are usually observed on the
ocean surface, although these can be treated as a superposition of many regular harmonic waves, each having its
own amplitude (na), wavelength (k), frequency (f), and direction of propagation (see Figure 5). This concept can be very

Fig. 5. Denitions of harmonic wave propagation and parameters. (a) In time domain. (b) In three-dimensional space.

useful in many applications since it allows one to predict


complex irregular behavior in terms of the theory of regular
waves; e.g., see St. Denis and Pierson (1953).

Wave Particle Kinematics


Wave Theory
Many regular wave theories have been developed to capture
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
among the offshore engineering community (Chakrabarti
2005). Airy wave theory is the simplest of the routinely-used
regular wave theories. Its key features are introduced in this
paper in the context of describing its role in modeling the
character of irregular sea states. Referring to Figure 5, with
the wave moving in the positive horizontal direction (x), the
form of the water surface (wave prole) can be expressed as
a function of x and time (t) by:
f x; t fa cos kx  xt

and with the wave moving in the opposite direction by:


f x; t fa cos kx xt

M. Arshad & B. C. OKelly

Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

where k is the wave number, x is the circular wave


frequency, and fa is the maximum amplitude of the water
particle measured from the mean water surface level.
The resulting water particle velocity components u(t) and
v(t) in the respective x and vertical directions can be
expressed by:
ut fa x

coshkhw kz
cos kx  xt
sinhkhw

vt fa x

sinhkhw kz
sin kx  xt
sinhkhw

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


elevation of the water particle above the mean surface level
in the vertical direction.
For deep water conditions, these velocity equations
reduce to:
ut fa xekz cos kx  xt

vt fa xekz sinkx  xt

From Eqs. (9) and (10), a simple interpretation is that for


deep water conditions, the elliptical orbits of water particles
associated with Airy wave theory reduce to circular orbits.
Different methods are available to determine the horizontal
velocity prole with depth, including Wheeler prole stretching, extrapolation, and constant extension methods. Since
the wave loads are dependent on the water particle velocity,
horizontal wave loads for shallow waters may be scrupulously undervalued if the calculations are based on deep
water equations. When considering maximum velocity
values, the time function of Eq. (6) can be ignored without
any signicant loss in accuracy. The time derivatives of
Eqs. (69) give the respective accelerations, which are key
parameters for the estimation of the inertial forces acting
on offshore structures located in the ow eld.
Morisons Equation
Morison, Johnson, and Schaff (1950) formulated an equation to predict the wave forces acting on a vertical pile
exposed to horizontal sinusoidal oscillatory ow. In their
equation, the linear inertial force and adapted quadratic
drag force (from real ows and constant currents) are superimposed to obtain the resultant force acting on the projected
area. For larger offshore structures (e.g., gravity foundations
and also large diameter monopiles), the wave eld is signicantly inuenced and a diffraction regime emerges. The
potential ow theory is more suitable for the calculation of
wave loads on such structures (Batchelor 1967). Morisons
formula is strictly limited for use with slender structural elements; i.e., characterized by D=k < 0.2: where D is the diameter of the structural element between the seabed level and
the transition piece (see Figure 3) and k is the impinging
wavelength of the ocean wave. However a signicant number of existing offshore structure designs have employed
Morisons equation even though the criterion of D=k < 0.2
may not have always been fully satised (Haritos 2007).

In the time domain, the total horizontal load per unit


length [F(t)] exerted on a xed object (e.g., OWT monopile
foundation) as a result of wave motion and currents can
be considered as the linear addition of the inertial [Finertia(t)]
and drag [Fdrag(t)] forces given by Morison, Johnson, and
Schaff (1950):
F t Finertia t Fdrag t

Finertia t

p
q CM D2  u_ t
4 w

1
Fdrag t qw CD D  ut  jutj
2

10

11

where qw is the density of sea water, D is the cylinder (i.e.,


monopile) diameter over the distance between the seabed
and ocean surface, and u(t) uacos(xt), ju(t)j and u_ t are
the horizontal water particle velocity, its absolute value
and its time derivative (acceleration), respectively, with ua
equal to the maximum along wave water-particle velocity
[discussed earlier in section on Wave theory], CM and CD
are the inertial and hydrodynamic drag force coefcients,
respectively, whose values are dependent on Reynolds number (Re) and KeuleganCarpenters number (KC):
Re

ua D
W

12

KC

ua T
D

13

where W is the kinematic viscosity and T is the oscillating


wave period.
A correct evaluation of the total hydrodynamic load acting on an offshore structure must consider the combined
current ow and wave particle velocities. Wave particle velocities and accelerations are sinusoidal in time, and when seen
as functions of time, are out of phase by 90 from each other.
Adopting the absolute value of the velocity term (i.e., ju(t)j)
in Eq. (11) synchronizes the drag force and ow direction.
Irregular variations of the total hydrodynamic load are produced against time for an idealized record, as demonstrated
by the experimental data shown in Figure 6 for a monopile
at reduced laboratory scale. For inertia dominated wave
loads and (or) small current velocity, the loads arising from
the ocean current can be ignored (DNV 2011).
The dominancy of the drag force or inertial force has been
assessed in terms of the KeuleganCarpenter number. Typical recommendations give the inertial force dominating for
KC < 10, comparable inertial and drag forces for
10 < KC < 20, with the drag force dominating for KC > 20.
In general, the CM and CD coefcients are dependent on
both the Reynolds and KeuleganCarpenters numbers
(Haritos 2007). For OWT monopile foundations, widely
used design codes such as API (2010) and DNV (2011)
specify=recommend appropriate CD and CM values, depending on the pile surface roughness and (or) the Re number.

Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

Analysis and Design of Offshore Monopile Foundations

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

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


the rst natural frequency (fnat, in Hz) of the whole structure
can be approximated by (Tong 2010):
2
fnat

Fig. 6. Graphical representations of variations in (a) velocity


and acceleration, and (b) horizontal force against time for an
idealized record related to a monopile at reduced laboratory
scale (Morison, Johnson, and Schaff 1950).

Correlation between Loading Frequency and Natural


Frequency of OWT Structures
It is essential to consider the fundamental natural frequency
of an OWT structure for a proper description and evaluation
of its dynamic behavior. For dynamic systems, resonance
occurs if an excitation frequency gets close to the structures
natural frequency. For OWTs, this invariably leads to the
development of higher stresses and, more signicantly, to a
higher range of stresses in the support structure, which is
an unfavorable situation with respect to fatigue life. Hence
it is important to ensure that excitation frequencies having
high energy levels do not coincide with the support structures natural frequency. As a rst approximation, the support structures natural frequency can be determined by
considering a simplied geometry for the whole structure
(Figure 7). The turbine mass (Mt) is concentrated at the top
(free end), similar to a cantilevered strut, which represents an

3:04
2

EI

4p 0:227 lLi Mt Li 3

14

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
wind and waves, with the effective wind load determined
by a complex interaction between the structural dynamics
of the turbine and the wind eld, which contains turbulent
gusts caused by eddies in the ow. The turbulent wind eld
originates from atmospheric turbulence, with nearby
turbines also contributing to the disturbance of the ow.
Energy-rich wind turbulence occurs at below 0.1 Hz
(LeBlanc 2009). The height of the ocean waves is usually
dened in terms of signicant wave height and determined
as the mean value of the highest one-third of the waves in a
given wave record. Ocean waves that induce fatigue loading
with high frequency on offshore structures usually have signicant wave heights ranging 1.01.5 m and also have a
zero-crossing period (Tz) of 45 s (De Vries 2007).
Modern OWTs are installed with either pitch-regulated
blades or variable rotational-speed systems to enable optimization of their power production over a wide range of wind
speeds. The rotational speed of the main rotor shaft is typically in the range 1020 rpm. Hence the rst excitation frequency 1P (i.e., corresponding to one full revolution)
occurs in the range 0.170.33 Hz and, in general, should only
be lightly excited. Large excitations in the 1P frequency
range generally arise from excessive turbine mass or

Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

M. Arshad & B. C. OKelly


comprehensive literature supports inherent anisotropic
behavior (or the general cross-anisotropy present in many
sedimentary deposits) for the soils strength=deformation
characteristics and also seepage properties (Yasin and
Tatsuoka 2000; Wang and Lade 2001; OKelly 2005, 2006).
Due to the initial and stress-induced anisotropy of most
deposits (Naughton and OKelly 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
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


having rated power-generation capacities ranging 2.03.6 MW
(LeBlanc 2009). (a) Deected shape of monopile. (b) Soil pressure pt exerted due to pile deection yt for a specic depth xt. (c)
Winkler model approach and change in shape of py curves
with depth.

aerodynamic imbalances. For a three-bladed turbine, the


blade passing-frequency of typically 0.51.0 Hz is denoted
by the 3P frequency, which is heavily excited, mainly on
account of the impulse-like excitation arising from the blades
passing by the tower. Site-specic spectral densities for wind
and waves can be derived from measured site data, met-ocean
databases or using numerical models (LeBlanc, Houlsby, and
Byrne 2010). Figure 8 illustrates the 1P and 3P excitation
ranges, along with realistic normalized power-spectra
representing aerodynamic and hydrodynamic excitations.
The regions before the 1P frequency range and after the
3P frequency range are referred to as the SoftSoft and
StiffStiff zones, respectively. The structure will be too exible if its natural frequency falls within the SoftSoft zone
and too rigid (heavy and expensive) if its natural frequency
falls within the StiffStiff zone; both of these scenarios
making it unsuitable for the design. Wind and wave-turbulence excitation frequencies usually fall within the SoftSoft
zone: another important reason for avoiding this frequency
region (LeBlanc, Houlsby, and Byrne 2010).

Modeling Soil Behavior Under Cyclic Loading

Simulation of In-situ Stress Conditions in the Geotechnical


Laboratory
The values of pertinent parameters used to describe the soil
response under cyclic loading can be determined in the geotechnical laboratory using cyclic triaxial tests (Das 2008),
although the system of cyclic axial loading and lateral connement pressure acting on the test-specimen is axisymmetric. An advancement on the cyclic triaxial apparatus is
the hollow cylinder apparatus (HCA) (OKelly and
Naughton 2005a) which allows independent control of the
magnitudes of the three principal stresses and also the orientation of the majorminor principal stress axis. The HCA is
ideal for simulating cyclic multi-directional loading conditions on cross-anisotropic test specimens. Generalized
stress-path testing can be performed in which the stress history and in-service loading conditions at specic locations in
the soil foundation can be simulated under stress- (OKelly
and Naughton 2009) or strain-controlled conditions
(Naughton and OKelly 2005; Das 2008). Special preparation techniques (OKelly and Naughton 2005b) are
required to prepare=reconstitute the test specimen (which
is often disturbed during the sampling procedure) in a physically identical condition to that of the in-situ deposit. In
many practical situations, laboratory testing may become
too laborious, expensive, and (or) time consuming. In-situ
testing techniques, including the Cone Penetration Test
method (Igoe et al. 2013) are used for offshore site investigations and afford another approach in the determination of
the pertinent design parameter values.

Real Soil Behavior

Modeling Soil Under Cyclic Loading

Soil deposits can be classied in many ways; e.g., by formation process, grain size (ne or coarse), plasticity index, age
(recent or aged deposits), mineralogical content, etc. Apart
from at very small strain levels of <103 strain (Atkinson
and Sallfors 1991; OKelly and Naughton 2008), the stress
strain relationship for soil is generally highly nonlinear
(inelastic) (Budhu 2011), with the strength and stiffness
properties strongly dependent on stress history, drainage
conditions (drained or undrained) and the stress path
followed during loading. For undisturbed deposits,

The strain accumulation occurring in soil under repeated


loading is dependent on the material properties, stress
path=level, and number of load cycles (Niemunis,
Wichtmann, and Triantafyllidis 2005; Karg 2007). Many
models with different complexity and acceptability have
been developed for the prediction of the strain accumulation
occurring in a soil element under cyclic loading. These
include models that are broadly based on: the number of
load cycles (Barksdale 1972; Sweere 1990; Hornych, Corte,
and Paute 1993); the stress level (Paute, Hornych, and

Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

Analysis and Design of Offshore Monopile Foundations

Benaben 1996); the number of load cycles and the stress level
(Pappin 1979; Lentz and Baladi 1981; Li and Selig 1996;
Lekarp and Dawson 1998; Chai and Miura 2002) or the
number of load cycles, stress level and material properties
(Niemunis, Wichtmann, and Triantafyllidis 2005; Karg
2007). A major limitation to their application in offshore
foundation design calculations is that none of these models
explicitly consider the (mono)pilesoil interaction under lateral loading. In reality, the pile deection (rotation) response
under lateral loading arises from the soil behavior, which is
dependent on the loading conditions. A few models have
been tailored for this particular scenario; these are described
in the next section, in the context of the design of OWT
monopile foundations. A discussion on numerical modeling,
considering dynamic constitutive soil models, torsional loading, and damping related issues, is beyond the scope of this
review article. Details on these topics can be found in Basack
and Dey (2012), Basack and Sen (2014), Guo (2006, 2013)
and Rani and Prashant (2014).

in-service over many decades. However, caution is necessary


in applying this methodology to OWT monopile foundation
design, since the approach may often be applied outside of
its veried range and several important design issues may
not be properly considered. The API (2010) and DNV
(2011) standards rely on methods (models) built upon
empirical data obtained for long exible piles, for which
bending (deection) is signicant. In contrast, existing and
planned OWT monopile foundations invariably have slenderness ratios of <10 (typical range of 56), indicating rigid
pile behavior (Achmus, Kuo, and Abdel-Rahman 2009;
Peng, Clarke, and Rouainia 2011). Under these circumstances, the pile rotation is generally more prominent over
bending, with the rotation occurring about a point (axis
of rotation) located approximately one pile diameter above
the pile base. Hence the design criteria and analyses appropriate for exible and rigid piles are considerably different
(Dobry et al. 1982), casting doubt on the application of
the API (2010) and DNV (2011) methods based on py
curves in predicting the in-service behavior of offshore
monopiles. Further, the py curves for cyclic loading presented in API (2010) and DNV (2011) were primarily formulated for the evaluation of the ultimate lateral load-carrying
capacity mobilized under relatively few load cycles. In contrast, OWT monopile foundations experience many millions
of low-amplitude cycles over their in-service life. Further, the

Strain Accumulation Models Considering PileSoil


Interaction
py Model
In general terms, a py curve is typically obtained by plotting
the soil pressure (p) response against the piles lateral
deection (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 corresponding pile deection (yt) response. In the literature curves (e.g.,
Matlock 1970; Reese, Cox, and Koop 1975; Ismael 1990; API
2010), py curves can be categorized on the basis of soil type
(granular or cohesive), loading type (monotonically increasing or cyclically repeated) and the groundwater table level.
The effects of soil stratication, nonlinearity, and other
soil properties are automatically considered by determining
py curves specic to different depth ranges along the length
of the pile (Figure 9c), which is typically modeled using
Winklers 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 reected by increasing values of the spring stiffness (Epy),
dened as the secant modulus of the py curve (Figure 9c).
The pile deection that develops under given loading conditions and constrains can be predicted by implementing
the related py curve in a simple non-dimensional framework, such as Randolphs method (Randolph 1981), or by
numerical methods using computer software such as
COM624P (1993). However, despite the stiffness Epy being
a soilstructure interaction parameter, API (2010) and
DNV (2011) only consider the soil properties in formulating
the py curves, and the inuence of the pile properties on the
mobilized py curves remains an open question.
Current pile design methodology based on py curves, as
described in API (2010) and DNV (2011), has gained
broad recognition on account of the low failure rate for piles

Fig. 9. py method of analysis for laterally loaded pile.

10

M. Arshad & B. C. OKelly

API (2010) and DNV (2011) methods do not provide a


means of calculating the accumulated pile deection
(rotation) that occurs during cyclic loading. Changes in the
foundation stiffness as a result of long-term cyclic loading,
which typically densies (but in some circumstances may
loosen) the surrounding soil (LeBlanc, Houlsby, and Byrne
2010), are also poorly accounted for in current design methodologies based on py curves.

Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

Model of Little and Briaud (1988)


The Little and Briaud (1988) model is based on experimental
results obtained from six cyclic lateral load tests performed
on long pre-stressed concrete and steel pipes (D 0.61.06 m)
installed in a 22 m thick stratum of loose to medium dense sand,
overlying stiff clay. Twenty load cycles were performed for each
test pile. The following power function was proposed for estimating the accumulated lateral strain (en) developed at ground
surface level for N number of load cycles:
en e 1 N m

15

where e1 is the strain response at the end of the rst load cycle
and exponent m is a parameter that accounts for the inuence
of the soil properties, pile material, installation method, and
loading characteristics, with experimentally derived m values
ranging 0.040.09 for the exible test piles investigated (Little
and Briaud 1988). In this context, the lateral strain is calculated
as the piles lateral deection occurring at the seabed level divided by its outer diameter; a dimensionless quantity.
Numerical studies by Kuo, Achmus, and Abdel-Rahman
(2012) indicated an m value of 0.07 for exible piles (agreeing with the experimental range reported by Little and Briaud
(1988)) and 0.135 for rigid piles. The validity of Eq. (15) has
been veried using cyclic lateral load test data obtained for
model piles installed in medium-dense quartz sand (Peralta
and Achmus 2010). It has also been shown that the value of
exponent m can be evaluated from consolidateddrained cyclic triaxial test data (Peralta and Achmus 2010). Nevertheless,
Eq. (15) is empirical and formulated on limited test data
(N < 20) for a loose to medium dense sand deposit.
Model of Long and Vanneste (1994)
The Long and Vanneste (1994) model is based on a closedform solution for a pile in an elastic foundation soil, with
the soil reaction modulus increasing linearly with depth.
By applying this approach and using input parameters
derived from reported eld data for cyclic lateral load tests
performed on thirty-four piles of different lengths, diameters, materials, and installation techniques, Long and
Vanneste (1994) proposed that the accumulated lateral strain
occurring at ground surface level for N load cycles can be
estimated by:
en

AH
EI0:4 nhN 0:6

BM
EI0:6 nhN 0:4

16

where EI is the exural rigidity of the pile; nhN is the


coefcient of soil reaction for the Nth cycle of loading; H
and M are the horizontal load and moment, respectively,
applied at the ground surface level; A and B are scalars

determined by the pile length and the relative stiffness Tr


[(EI=nhN)0.2] value.
In Eq. (16), all of the variables can be taken as known
except for coefcient nhN, which is calculated as:
nhN nh1 N t

17

where nh1 is the coefcient of soil reaction for the rst cycle
of loading, its value dependent on the relative density of the
soil and the location of the groundwater table (Terzaghi
1955; Reese, Cox, and Koop 1974), with the exponent t (having a recommended range of 0.20.4 (Broms 1964; Davisson
1970)) determined empirically by:
t 0:17 FL Fi FD

18

where FL, Fi, and FD are empirically determined factors that


take into account the inuence of the cyclic load ratio, pile
installation method, and soil density, respectively. Note
nh1 nh; i.e., its value is the same as that for the monotonic
(static) loading condition.
The Long and Vanneste (1994) model provides a relatively straightforward procedure for predicting the effect of
cyclic lateral loading, but specically restricted to scenarios
involving the soil reaction modulus increasing linearly with
depth. Soil stratication, nonlinearity, and other fundamental parameters (e.g., unit weight and strength) that affect the
lateral load response are not explicitly taken into account by
this model. Further, most of the thirty-four pile tests
considered in the development of Eq. (16) involved the
application of less than fty lateral load cycles.
Model of Lin and Liao (1999)
Based on the results of twenty-six full-scale piles tested under
cyclic lateral loading (N < 50), Lin and Liao (1999) proposed
a logarithmic trend to capture the accumulated strain en
developed at the ground surface level for N load cycles:
en e1 1 t lnN 

19

where t is a degradation parameter determined by:


t 0:032

L
FL Fi FD
Tt

20

where Tt EI=nh 0:2 , L is the embedded length of the pile


and nh is the coefcient of soil reaction for the monotonic
(static) loading condition.
The other contributing factors in Eq. (20) are similar to
those in Eq. (18), with insignicant differences in values
from those reported for the Long and Vanneste (1994)
model. However, considering the various inuencing factors,
caution is urged regarding empirical parameters used in
describing the logarithmic change in the pile deection=
rotation.
In their study, Lin and Liao (1999) considered two loading scenarios: one-way loading, in which the applied horizontal load increases from zero to its maximum value
(Hmax) before reducing back to zero, thereby completing
one cycle of loading; two-way loading, in which the horizontal load is applied in the opposite direction for half of each

11

Analysis and Design of Offshore Monopile Foundations

model explicitly considers the effects of soil density, mean


effective conning pressure, cyclic stress amplitude, simultaneous oscillation of stresses from different directions,
and the number of load cycles:

Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

e_acc fampl f_N fe fp fY fp

Fig. 10. Different modes of load variation against time.

cycle, reducing its value below zero to a minimum of Hmin


(see Figure 10).
According to the Long and Vanneste (1994) and Lin and
Liao (1999) models, one-way loading produces greater accumulated strain compared with two-way loading. Similar
overall behavior has been reported from studies on model
monopiles (Peng, Clarke, and Rouainia 2011), with balanced
two-way loading (Hmax Hmin) producing about 80% of
the lateral deection recorded under one-way loading of
the same magnitude. Unbalanced two-way loading
(Hmax 6 Hmin) was reported to produce even lower
amounts of lateral deection. However some recent physical
model studies suggest that unbalanced two-way loading is
more damaging to the lateral stability of the monopile
(Zhu, Byrne, and Houlsby 2013; Arshad and OKelly
2014). An increase in soil stiffness was also observed with
increasing number of load cycles. Similar formulations to
that given by Eq. (19) have been proposed by Verdure,
Garnier, and Levacher (2003) and Li, Haigh, and Bolton
(2010) based on centrifuge experiments performed on miniature piles installed in sandy deposits, with the degradation
parameter t estimated by curve-tting analysis of the experimental data. The change in soil stiffness was found highly
dependent on the cyclic load amplitude. Up to certain levels
of cyclic load amplitude and numbers of load cycles, the stiffness was found to increase rapidly, but no further change was
observed for higher load amplitudes, even during the rst few
cycles (Verdure, Garnier, and Levacher 2003).
High Cycle Accumulation Model (2005, 2010)
The high cycle accumulation model is based on the experimental data obtained from a large collection of experimental
work, comprising index tests and drained cyclic triaxial, resonant column and simple shear tests performed on sand specimens having a wide range of initial relative densities
(Niemunis, Wichtmann, and Triantafyllidis 2005). Although
this model was initially developed without consideration of
pilesoil interaction, its potential application in modeling a
pilesoil system under low-amplitude stress cycles has been
demonstrated by nite-element calculations using ABAQUS
software (Wichtmann et al. 2010). The formulation of this

21

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 conning pressure function, fY is the average stress ratio (i.e.,
mobilized deviatoric stress to average mean conning pressure applied during a load cycle) function and fp is the polarization (load cyclic shape) function. Further details on the
determination of these functions and commonly adopted
values are reported by Wichtmann, Niemunis, and Triantafyllidis (2009), Wichtmann et al. (2010).
Wichtmann et al. (2010) applied the high cycle accumulation
model for typical North Sea ne sand deposits and found that
greater permanent (plastic) lateral deection occurred for:
. Decreasing relative density, with the average value of the

applied moment (Mave) and its amplitude (Mampl) held


constant;
. Increasing Mampl, with the relative density and Mave held
constant;
. Increasing Mave, with the relative density and Mampl held
constant.
These outcomes are qualitatively conceivable and broadly
consistent with the behavior observed in laboratory studies
on model monopiles by LeBlanc, Houlsby, and Byrne
(2010) and Peng, Clarke, and Rouainia (2011), although
further physical model studies and in-situ testing are
required to conrm these ndings. Potentially signicant
shortcomings of the high cycle accumulation model include:
the multiplicative approach used in Eq. (21) means that
potentially small errors in the values of the individual functions may produce an objectionable error in the nal calculation value; the formulation includes no measure of loading
frequency which is an important factor in connection with
the natural vibration frequency of OWT structures [see
Section on Correlation between Loading Frequency and
Natural Frequency of OWT Structures] and the strain
accumulation rate.
Model of Achmus, Kuo, and Abdel-Rahman (2009)
The Achmus, Kuo, and Abdel-Rahman (2009) model is
based on the concept of the soil stiffness degradation
approach suggested by different researchers (e.g., Little
and Briaud 1988; Long and Vanneste 1994), with emphasis
on the degradation of the soil secant stiffness (Es), in keeping
with formulations proposed by Huurman (1996) and Werkmeister (2003). According to this model, an increase in the
plastic strain can be interpreted as a decrease in the Es value.
In cyclic triaxial tests (for which elastic strains are usually
negligible) and referring to Figure 11, the plastic axial strains
produced at the end of the rst and Nth cycles (epN 1 and
epN N) can be determined from the rate of degradation of
the secant modulus measured for the rst and Nth cycles

Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

12

M. Arshad & B. C. OKelly

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

(EsN 1 and EsN N respectively) by:


epN1
EsNN

epNN
EsN1

22

In the Achmus, Kuo, and Abdel-Rahman (2009) study,


the characteristics of the permanent strains derived from
drained cyclic triaxial test data for medium and dense sand
were implemented in a 3D nite-element model of a laterally
loaded pile. The Achmus, Kuo, and Abdel-Rahman (2009)
model suggests that the accumulated strain en developed
for N load cycles can be estimated using the semi-empirical
approach proposed by Huurman (1996) for the calculation
of the accumulated plastic strains in cyclic triaxial tests, as:
e1
en
23
b1 xc b2
N
where b1 and b2 are model parameters and xc is the characteristic cyclic stress ratio (ranging 01 (Achmus, Kuo, and
Abdel-Rahman 2009)), dened as:
cyclic stress ratio at loading
xc

 cylic stress ratio at unloading


1  cylic stress ratio at unloading

The cyclic stress ratio at constant conning pressure is


dened as the ratio of the major principal stress under cyclic
loading to the major principal stress at failure under static
loading. Since the horizontal loading varies cyclically for
OWT monopile foundations, the connement pressure acting on soil elements surrounding the pile shaft also changes.
Hence the cyclic stress ratio values for cyclic triaxial tests (in
which the axial load is varied cyclically) and OWT monopile
foundations (where the axial load remains constant and the
horizontal load varies cyclically) are different. To overcome
this, the characteristic cyclic stress ratio is used in Eq. (23)

and accounts for the loading and unloading phases emerging


due to the cyclic horizontal loading. Achmus, Kuo, and
Abdel-Rahman (2009) found good agreement between
numerical simulations performed using cyclic triaxial test
data (Timmerman and Wu 1969) and experimental data
for laterally loaded model monopiles installed in sand.
Model of LeBlanc (2009)
The LeBlanc (2009) model is based on lateral load tests,
involving between 8  103 and 6  104 load cycles, performed
on rigid model piles having a slenderness ratio of 4.5 and
which were installed in unsaturated, very loose and medium
dense Leighton Buzzard sand beds. A non-dimensional
framework was developed to identify realistic pile dimensions and loading ranges for the laboratory study. Some
particular loading characteristics were found to cause signicant increases in the accumulated pile rotation to occur, with
the soil stiffness increasing with the number of load cycles.
This contrasts starkly with the Long and Vanneste (1994),
API (2010) and Achmus, Kuo, and Abdel-Rahman (2009)
models for which the static loaddisplacement curves are
degraded to account for cyclic loading.
LeBlanc (2009) proposed that the accumulated rotation hN
(and hence the lateral displacement) of a rigid pile developed
for N load cycles can be estimated by (see Figure 12):
hN h0 DhN

24

where h0 is the pile rotation achieved when the applied


moment reaches its maximum value during the rst load cycle
and Dh(N) is given by:
DhN hs Tb fb ; Rd Tc fc N 0:31

25

where hs is the pile rotation produced by a static load having


the same magnitude as the maximum cyclic load (Figure 12b),
Tb and Tc are dimensionless functions, Rd is the relative
density of the sand, with fb (a measure of the cyclic load

Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

Analysis and Design of Offshore Monopile Foundations

13

Fig. 12. Variation in pile rotation h with uctuation in moment


M. (a) Cyclic test (b) Static test.

Fig. 13. Functions relating (a) Tb and (b) Tc to the cyclic load
characteristics in terms of fb and fc respectively.

amplitude normalized with respect to the static moment


capacity, MR) and fc (quanties characteristics of the cyclic
load) given by:
fb

Mmax
MR

26

fc

Mmin
Mmax

27

where Mmin and Mmax are the minimum and maximum


moments developed in a load cycle.
It follows that 0 <fb < 1, with fc equal to unity for a static
load test, zero for one-way loading and 1 for balanced twoway loading. Values of Tb and Tc, determined from experimental data of Dh(N)=hs against N, were plotted against fb
and fc, as shown in Figure 13. Note that with Tc 0 for
fc  1.0, Eq. (25) proposes that the rotation occurring
after the rst load cycle is equal to zero for balanced twoway cyclic loading, which is fairly unrealistic in practice.

Further, this model cannot account for the multidirectional


loading typical of offshore wind-farm environments.
Considering the accumulated strain en produced at the
ground surface level for a rigid pile is proportional to the
accumulated rotation, then the predictions by this model
(with due consideration for the cyclic character of the loading) are contrary to those for exible piles according to the
Long and Vanneste (1994) and Lin and Liao (1999) model.
For loose and medium dense sand, LeBlanc (2009) also
reported an increase in soil stiffness due to cyclic loading
of the rigid pile, although further research for higher densication levels, other loading frequencies and degrees of saturation are necessary to conrm this preliminary experimental
nding.
Model of Bienen et al. (2012)
The Bienen et al. (2012) model is based on experimental data
from miniature monopiles installed in dry medium-dense
sand and tested at 1g and 200g, using a centrifuge apparatus
for the latter. The test piles were representing a prototype

14

M. Arshad & B. C. OKelly

Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

pile 2.4 m in diameter (D), with embedment lengths (L) of 9.6


and 30 m; i.e., 4D (rigid) and 12.5D (exible) piles, respectively. Using a sinusoidal waveform, different magnitudes of
one-way lateral loading were investigated considering 104
load cycles and a single model frequency of 0.25 Hz. The
accumulated lateral deection y occurring at the pile head
(assumed ush with the surrounding ground surface) for N
load cycles can be estimated by:

as
H
y D fN As 100 2
28
D LQc
where H is the horizontal load applied to the pile head, Qc is
the rate of change in CPT cone-tip resistance with depth, As
and as are dimensionless parameters dependent on the soil
properties and pile dimensions (Dyson and Randolph
2001; Duhrkop 2009), with reported values of As 1.4 and
as 0.0072 for monopiles (Bienen et al. 2012), and fN (factor
to modify the monotonic response (deection) into a cyclic
response) approximated by:
N 1
29
BN1 lnBN2 1
N
where the parameters BN1 and BN2 are determined from a
plot of [(y=D)n=(y=D)1] against number of load cycles N.
This model, based on the one-way lateral loading and single frequency of 0.25 Hz investigated, was found to provide
reasonable predictions of the accumulated strain produced
under cyclic loading, although the soil stiffness response
was not clear: increasing for up to certain levels of accumulated strain, but then decreasing with further load cycles.
Hence these ndings should be veried for greater numbers
of load cycles occurring in multiple directions and for other
frequency ranges, degrees of soil saturation and densication
levels.
fN 1

Model of Klinkvort and Hededal (2013)


Klinkvort and Hededal (2013) applied the model proposed
by LeBlanc (2009) to experimental data for rigid model
piles (slenderness ratio of 6) installed in saturated and dry
dense sand that were tested at ng using a centrifuge apparatus. These tests involved a maximum of 500 load cycles
and were performed at the same frequency and soil density.
They proposed that the piles accumulated lateral strain produced at the ground surface level after N number of load
cycles (en) can be estimated as:
en e1 N

Tb fbb Tc fcc

Design Procedure for OWT Foundation System


Design Motive
The basic driving motive for the design procedure is to avoid
the occurrence of resonance in the dynamic behavior of the
structural system under in-service loading (Jaimes 2010). An
iterative procedure is usually adopted in design, with the
basic steps involved in the design process for an OWT monopile foundation system illustrated in Figure 14. The data
required include environmental, turbine and site stratigraphy (soil prole) data. The environmental data is used to
determine the required work-platform and hub elevations
(see Figure 3) for the proposed OWT, and to select initial=
trial dimensions=geometry for the monopile foundation,
leading to the determination of the natural frequency of
the whole structural system. Checks on resonance frequency
are applied along with predictions of the anticipated
rotation, deection, and settlement responses of the proposed foundation system produced by the applied loads=
moments that are determined from the environmental and
wind turbine data. Fatigue and buckling checks are performed at a more advanced stage of the design process,
usually using some computer software package, and hence
are beyond the scope of this paper. The following sections
present a design example in which the work-platform and
hub elevations along with the required embedment length
for a monopile foundation supporting a typical 5 MW
OWT (Jonkman et al. 2009) are determined.
Determination of Work Platform and Hub Elevations
Referring to Figure 14, the rst step in the design process
requires the determination of the turbine platform and hub
elevations, usually with reference to the lowest astronomical
tide (LAT); dened as the lowest tide level to occur under
average meteorological conditions and any combination of
astronomical conditions. LAT is below the mean sea level
(MSL). The high astronomical tide (HAT) is above the
MSL, but below the storm surge level.
The work-platform level is located at the top of the
transition piece and it determines the location of the ange

30

with fbb and fcc dened as:


fbb

Pmax
Pu

31

fcc

Pmin
Pmax

32

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,
applied to the pilehead during the cyclic loading.

Fig. 14. Flowchart for the design of offshore wind-turbine


support structure and monopile foundation.

15

Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

Analysis and Design of Offshore Monopile Foundations


Table 2. Values of key parameters for typical 5 MW wind
turbine

Determining the Required Natural Frequency and


Preliminary Tower Dimensions

Parameter

From the rotational speed interval of the main rotor shaft,


as given in Table 2, the 1P region ranges 0.120.20 Hz and
the 3P region ranges 0.360.60 Hz (Jonkman et al. 2009).
For this particular design example, the target value of the
natural frequency [Eq. (14)] was set as 0.27 Hz; i.e., within
the wanted frequency range of 0.200.36 Hz. By taking into
account the hub height, wind-turbine mass and target natural frequency, the outer diameter and wall thickness at the
top of the steel tower were estimated as 3.9 m and 40 mm,
respectively, with corresponding values at the tower base
of 6.6 m and 100 mm. These preliminary selected dimensions
must also satisfy the buckling and fatigue checks; otherwise
the calculations are revised iteratively. The outer diameter of
the transition piece (DTP) subsequently depends on the diameter of the monopile, as:


DTP D 2 tTP tgrout
36

Magnitude

Turbine mass (rotor and nacelle)


Rotor diameter
Nominal rotor speed
Rotational speed interval
of the main rotor shaft
Cut-in wind speed
Nominal wind speed
Cut-out wind speed

350 tonnes
126 m
10.1 rpm
6.912.1 rpm
4 m=s
11 m=s
25 m=s

connection between the transition piece and the turbine


tower (see Figure 3). The magnitudes of the applied wind
loads and the location of the centre of gravity of the nacelle
mass is strongly dependent on the hub height. Representative values for key parameters of the typical 5 MW wind turbine considered in this design example are given in Table 2.
Environmental data characteristic of an offshore wind-farm
located in the North Sea are given in Table 3. For this
location, a tidal range (DZtide HATLAT) of 1.6 m and
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:
Zplatform LAT DZtide DZsurge DZair n

33

where DZair is the air gap of typically 1.5 m between the


work-platform level and the highest wave crest elevation (n).
For an assumed wave-height coefcient value of 0.65 and
a maximum wave height, relative to the MSL, having a 50
year return period (Hmax(50 year)) of 20.3 m, the value of
n is given in DNV (2011) as:
n 0:65 Hmax 50 year  13:2 m

34

And hence the value of Zplatform is determined as


LAT 19 m (rounding the latter up to the nearest whole
integer). A check should also be applied (De Vries and van
der Tempel 2007 DNV 2001) such that Hmax(50
year) < 0.78  30 m; the assumed MSL value for these calculations. The required hub height (Zhub) of LAT 87m above
seabed level is determined from:
Zhub Zplatform 0:5rotor diameter 5

35

With LAT MSL  Ztide=2, Zhub 116.2 m from seabed


level.
Table 3. Extreme wind velocity (Vw), current velocity (Uc),
signicant wave height (Hs), and maximum wave height (Hmax)
as a function of return period (Treturn) for reference offshore
wind-farm site (Jonkman et al. 2009)
Treturn (year)

Vw (m=s)

Uc (m=s)

Hs (m)

where D is the monopile diameter (usually equal to the tower


base diameter), tTP is the wall thickness of the transition
piece and tgrout is the thickness of grout connection.
Calculation of Design Loads
The design loads for the subject monopile foundation
system were estimated using the preliminary tower geometry,
the wind-turbine parameter values given in Table 2 and the
environmental data given in Table 3. For this example, the
software package RECAL was used to estimate the shear
force and bending moment acting on the monopile foundation at seabed level, as documented by De Vries and
van der Tempel (2007). The torsional moments are usually
only a small fraction (<1%) of the bending moment (Lesny
and Wiemann 2005); e.g., see typical loading values listed in
Table 1 for a 5 MW wind turbine having a monopile foundation system. Hence, provided checks on lateral forces
and bending moments are satised, torsional moments are
generally not critical and have not been specically considered in design examples for OWT monopile foundation
systems reported in the literature. RECAL can simulate both
wind and wave time-series, from which it calculates the
environmental loads acting on the support structure, as well
as the structures dynamic behavior. For the design example
under consideration, Table 4 presents load data in terms of
the lateral forces and bending moments generated by
RECAL. The reported load values take into account the
safety factors for hydrodynamic loading and aerodynamic
Table 4. Estimated design loads occurring at seabed level for
design example

Hmax (m)
Load type

1
5
10
50

33.9
38.0
39.8
43.9

0.70
0.80
0.84
0.94

7.72
9.03
9.60
10.91

14.35
16.80
17.85
20.29

Aerodynamic
Hydrodynamic
Total

Lateral force, H
(MN)

Bending moment, M
(MN.m)

1.42
9.27
10.69

127
298
425

16
thrust on the rotor, in addition to the different critical
combinations of wind and wave loading, as reported in
several guidelines=standards (API 2010; DNV 2011).

Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

Monopile Embedment Length and Foundation Stability


The monopiles embedment length must be sufcient to
ensure vertical and lateral stability. Studying the interaction
effects for piles under combined axial and lateral loading
would call for a systematic and sophisticated analysis,
although the pertinent literature is limited and sometimes
contradictory. For instance, analytical studies by Ramasamy
(1974) and Goryunov (1973) indicate that for a given lateral
load, the piles lateral deection increases for the combination with vertical loading but some experimental
(Sorochan and Bykov 1976; Jain, Ranjan, and Ramasamy
1987) and eld (Bartolomey 1977; Zhukov and Balov
1978) studies have indicated the contrary. Numerical analysis by Karthigeyan, Ramakrishna, and Rajagopal (2007)
indicated that for sandy soils, the presence of vertical loads
increases the piles lateral load-carrying capacity by as much
as 40% (depending on the magnitude of the axial loading),
although marginal reductions in the lateral load-carrying
capacity were found to occur for clayey soils.
According to current practice, for monopile design, separate analyses are performed that consider (i) the axial loading
only, to determine the bearing capacity and settlement
response, and (ii) the lateral loading only, to determine the
exural behavior through cantilever action (Karthigeyan,
Ramakrishna, and Rajagopal 2006; Moayed, Mehdipour,
and Judi 2012; Rahim and Stevens 2013). However, compared 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 sufcient 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 (displacement at the pile base) and excessive deection=rotation
of the pile itself from occurring. The pile size and embedment 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 subject monopile.
The input data for the monopile design includes the soil
strength and stiffness (e.g., elastic (Es) and shear (Gs) modulii and Poissons ratio) proles 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

M. Arshad & B. C. OKelly


design example are given in Figure A1 of the Appendix. It
was assumed that the monopile was made of steel having
yield stress and Youngs modulus values of 248 MPa and
207 GPa, respectively.
The embedment length of the monopile was initially
assumed equal to nine times its outer diameter. Checks on
the pile rotation and (or) its deection occurring at the
seabed level are applied and the monopile embedment length
is optimized accordingly. The monopiles rotation (h0) and
lateral deection (v0) occurring at the seabed level under
the design loads (horizontal force (H) and bending moment
(M)) are calculated by considering the closed-form solutions
proposed by Randolph (1981), Broms (1964) and Matlock
and Reese (1960). These methods are related to monotonic=
static loading conditions and hence there is no consideration
of the number of load cycles. Randolphs (1981) method,
illustrated in Figure 15, employs Eqs. (37) and (38) to calculate v0 and h0, respectively; with parameters Ep (effective
Youngs modulus of the pile), Gc (equivalent shear modulus
of the soil at a depth of 0.5Lc), Lc (active pile length) and Rc
(the soils equivalent shear modulus prole parameter)
dened in Figure 15. In this method, the monopile dimensions are incorporated in terms of its active length and
second moment of area, I; the latter an indirect involvement
of its cross-sectional area (i.e., its inner and outer diameter
dimensions). When the active length exceeds the actual=proposed monopile embedment length (L), the monopile tends

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


deecting; hence the monopile deection obtained through
Eq. (37) must be modied according to Eq. (39).

0:142
 1
 2 !
Ep =GC
Lc
Lc
v0
0:27H
0:3M
37
Rc Gc
2
2

0:142
 2
 3 !
Ep =GC
Lc
Lc
0:5
0:3H
0:8Rc M
h0
Rc Gc
2
2

embedded length of the pile. However, for a rigid pile, this


decreasing trend may not be applicable. According to this
method (Matlock and Reese 1960), beyond a certain value
of embedment length, there will be no further reduction in
the monopiles deection or rotation. Calculations are not
included for the present design example, but the pile embedment length should be sufcient to achieve tolerable values
of v0 and h0 occurring at the pile toe level. For the calculation of the v0 and h0 values occurring at the seabed level,
a parametric study was performed, varying the piles embedment length, outer diameter and inner diameter in the ranges
2040 m, 5.57.0 m, and 5.46.5 m, respectively. A summary
of the results of this study is presented in Figures 1719 and
conversed as:

Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

38
v0 adjusted value 0:8v0from Equation 37:

 
Lc
L

39

. The monopile deection=rotation values predicted by the

Bromss (1964) method of calculation for v0 is illustrated


in Figure 16. Matlock and Reeses (1960) method employs
Eqs. (40) and (41) to calculate the values of v0 and h0,
respectively, occurring at the seabed level.
v0 Av0

HTt 3
MTt 2
Bv0
EI
EI

three selected methods for the static design load conditions


grossly differed from one another; e.g., for the applied
loads and range of monopile embedment lengths and

40

HTt 2
MTt
Bh0
41
EI
EI
where Av0, Bv0, Ah0, and Bh0 are sets of non-dimensional
coefcients (scalars) whose values are dependent on the
depth along the embedded length of the pile and Tt (EI=
nh)0.2, with the value of the coefcient of soil modulus variation, 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 coefcients Av0, Bv0, Ah0, and Bh0 are 2.43, 1.62, 1.62, and 1.75,
respectively, with these coefcients decreasing in value with
increasing depth along the pile embedment length, reecting
the reduction in v0 and h0 as we move down along the
h0 Ah0

Fig. 16. Charts for calculating the lateral deection occurring at


the ground surface level for laterally loaded piles in cohesionless
soil (after Broms (1964)).

Fig. 17. Using Randolphs (1981) method, effects of piles


embedment length and diameter on its (a) lateral deection
occurring at the seabed level and (b) rotation.

18

M. Arshad & B. C. OKelly

Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

minor change in v0 was observed for pile embedment


lengths greater than 26 m, irrespective of the piles crosssectional area (Figure 18).
. For the Matlock and Reeses (1960) method, v0 and h0
were both found dependent on the inner and outer diameters (cross-sectional area) of the monopile, although,
their signicance tended to reduce substantially for pile
inner and outer diameters greater than 5.8 and 6.0 m
respectively (Figure 19).

Fig. 18. Summary of the results deduced using Bromss (1964)


method.

inner=outer diameters considered in the design example,


the predicted values of v0 ranged 70270, 25115 and
2590 mm for the Randolphs (1981), Bromss (1964)
and Matlock and Reese (1960) methods respectively. This
noticeable difference is perhaps due to the involvement of
marked empiricism in the methods used to evaluate the
lateral stability of the different monopile set-ups and the
limited available experimental data used in their
formulation=calibration.
. For Randolphs (1981) method, v0 was found sensitive to
the piles embedment length and its inner and outer diameters (incorporated in calculations of the cross-sectional
area and second moment of area, I) (Figure 17a), whereas
h0 was sensitive to the pile cross-sectional area only
(Figure 17b).
. For Bromss (1964) method, v0 was found dependent on
the embedment length and inner and outer diameters
(cross-sectional area) of the monopile, although only a

Fig. 19. Summary of the results deduced using the Matlock and
Reeses (1960) method.

The monopiles lateral response for the long-term cyclic


lateral loading condition was assessed for the presented
design example using the models after Little and Briaud
(1988), Long and Vanneste (1994), Lin and Liao (1999)
and Klinkvort and Hededal (2013), which have been discussed 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.
Figure 20 presents the accumulated lateral response of the
monopile predicted by the different models considered.
The accumulated lateral response is in fact a multiple of
the response for the end of the rst load cycle, realistic values
of which cannot be estimated using the models considered,
and for the purpose of these calculations, its value was
assumed equal to unity. Hence Figure 20 should be viewed
in the context of a qualitative comparison of the long-term
predictions by the different models. From Figure 20, it can
be concluded that for the same loading conditions, large differences can occur between the accumulated lateral
responses of the monopile predicted by the different models,
perhaps by many folds for large numbers of load cycles. One
reason for these large differences can be put down to the
empirical selection of values for the coefcients=non-dimensional parameters incorporated in these models for the prediction of the monopiles lateral response under long-term
cyclic loading.
The monopiles ultimate axial load-carrying capacity
(under static loading) (Qd) can be determined using

Fig. 20. Accumulated lateral strain responses predicted for the


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),


and has been employed by many researchers; e.g., De Vries
(2007); Igoe, Gavin, and OKelly (2010); Haiderali, Cilingir,
and Madabhushi (2013); Bisoi and Haldar (2014); to name
a few.

. Pile properties (e.g., diameter and slenderness ratio) and

Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

Qd Qf Qb fAsur qAb

42

where Qf is the shaft-friction capacity, Qb is the end-bearing


capacity, f is the unit skin friction, Asur is the shaft area
pertaining to the pile embedment length, q is the unit endbearing capacity, and Ab is the base area of the plugged
monopile.
For monopiles having a diameter of 4 m diameter or
greater, the pile plug resistance is usually not taken into
account in the calculations (van der Tempel 2006). Further,
it has been shown that degradation of the shaft-friction
capacity due to cyclic axial loading leads to accumulating
displacements and potentially a severe reduction in the axial
load-carrying capacity (Gavin and OKelly 2007), although
consideration of this degradation of the shaft friction in
the design process is still an open question. Numerous design
charts available in the literature make it possible to distinguish between stable and unstable loading levels to ensure
a design solution on the safe side. Further, no reduction in
axial load-carrying capacity is expected if a certain magnitude of the cyclic load amplitude is not exceeded (see Poulos
1988, and Abdel-Rahman and Achmus 2011, for further
details).

Discussion
There is no overall agreement in the literature regarding the
determination of the piles rotation=lateral deformation
response to the many millions of low-amplitude lateral load
cycles associated with OWT monopile foundation scenarios.
Considerable differences of opinion exist on the rate of cyclic
strain accumulation; e.g., power function (Little and Briaud
1988) and logarithmic-trend relationships (Lin and Liao
1999) have been proposed. Most of the reported eld and
laboratory pile tests were performed for medium dense sand
(Little and Briaud 1988; LeBlanc 2009; Bienen et al. 2012).
Compared with the Bienen et al. (2012) and LeBlanc
(2009) models which are based on N  104 lateral load cycles,
other widely used models for predicting the piles rotation=
lateral displacement response, including the API (2010),
DNV (2011), Little and Briaud (1988) and Long and
Vanneste (1994) approaches, are based on experimental data
for relatively few load cycles (N < 200). Since no reliable
model presently exists, design requirements regarding the
piles rotation=lateral displacement behavior under monotonic (static) extreme load are used as a substitute. Areas
warranting further in-depth research are the effects on
monopile behavior of:
. Soil properties; e.g., relative density, over-consolidation

ratio, and the relative signicance of the at-rest earth


pressure coefcient (K0) and the soil stiffness, including
their variations with depth;

the soilpile interaction;


. Varying the direction and frequency of the applied lateral

loading;
. Combined (e.g., axial and lateral) cyclic loading.

The scarcity of eld data for cyclic lateral loading of large


diameter (rigid) piles available in the literature, particularly
for high load cycling, makes the validation of current and
improved design methods=theories and the calibration of
numerical models for offshore monopiles difcult. Many
of the proposed models=formulations, with the principal
ones described earlier in the paper, were calibrated against
experimental data obtained from model-scale pile tests.
Further instrumented eld testing of full-scale monopiles,
ideally having comparable size and geometry (slenderness
ratio) with that of current and proposed OWT monopile
foundations, and subjected to high numbers of lateral load
cycles, is warranted and would provide a valuable source
of information in this regard for the wind-energy sector.

Summary and Conclusions


Sources, types, and methods of analyses for the determination of the magnitude of the environmental loading and
resulting moments exerted on offshore monopiles, including
for long-term and extreme conditions, are well documented
in the literature. However the behavior of the pilesoil system 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 proposed OWT structures is well outside the scope of present
experience and analysis=design methods, which are mainly
based on experimental data obtained for relatively smalldiameter exible piles subjected to low numbers of load
cycles (N < 200). This includes the widely used API (2010)
and DNV (2011) standards developed primarily for the offshore oil=gas industry.
The behavior of large-diameter monopiles in general, and
the long-term low-amplitude cyclic lateral loading response
in particular, are not well documented. Existing models for
estimating the accumulated lateral strain (rotation) response
of monopiles are based on very limited eld=laboratory data
and are currently not capable of explicitly accounting for
site-specic soil properties and environmental loading characteristics. There is also no consensus among researchers
regarding the severity of strain accumulation due to oneway and two-way loading scenarios, the effects of varying
the load direction and frequency or changes in soil stiffness
under long-term stress application.
Instrumented eld tests on full-scale rigid monopiles,
combined with a more extensive program of testing on
model-scale piles installed in different soil conditions, considering changing load characteristics (amplitude, frequency,
and direction) and subjected to high numbers of load cycles
of more than 106 are warranted. Such studies would provide
valuable information for the validation of current and
improved design methods=theories for offshore monopiles
and the calibration of pertinent numerical models. They

20
would also be helpful in generating computer code for
numerical simulations of more realistic conditions, especially
in-situ soil conditions and in-service loading characteristics.

Acknowledgment

Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

The rst author gratefully acknowledges a Postgraduate


Research Scholarship Award from Trinity College Dublin.
The writers also thank the reviewers for many helpful
comments.

References
Abdel-Rahman, K. and M. Achmus. 2011. Behavior of foundation
piles for offshore wind energy plants under axial cyclic loading.
Proceedings of the SIMULIA Customer Conference, Barcelona,
Spain, Dassault Syste`mes Press, Paris, France, 1719th May
2011, 33141.
Achmus, M. 2010. Design of axially and laterally loaded piles for the
support of offshore wind energy converters. Proceedings of the
Indian Geotechnical Conference GEOtrendz2010, Mumbai,
India, December 1618th, 2010, 92102.
Achmus, M., Y.-S. Kuo, and K. Abdel-Rahman. 2009. Behaviour of
monopile foundations under cyclic lateral load. Computer and
Geotechnics 36(5): 72535. doi:10.1016=j.compgeo.2008.12.003
API (American Petroleum Institute). 2010. Recommended practice for
planning, designing and constructing xed offshore platforms
Working stress design. API RP 2A-WSD (R2010), 22nd ed.,
API Publishing Services, Washington, DC, USA.
Arshad, M. and B. C. OKelly. 2013. Offshore wind-turbine structures:
A review. Proceedings of the ICE, Energy 166(4): 13952.
doi:10.1680=ener.12.00019
Arshad, M. and B. C. OKelly. 2014. Development of a rig to study
model pile behaviour under repeating lateral loads. International
Journal of Physical Modelling in Geotechnics 14(3): 5467.
doi:10.1680=ijpmg.13.00015
Atkinson, J. H. and G. Sallfors. 1991. Experimental determination of
stressstraintime characteristics in laboratory and in-situ tests.
Proceedings of the 10th European Conference on Soil Mechanics
and Foundation Engineering, Florence, Italy. Balkema: Rotterdam, The Netherlands, May 2630th, 1991, Vol. 3, 91556.
Barksdale, R. D. 1972. Laboratory evaluation of rutting in base course
materials. Proceedings of the 3rd International Conference on the
Structural Design of Asphalt Pavement, London, UK, September
1115th, 1972, Vol. 1, 16174.
Bartolomey, A. A. 1977. Experimental analysis of pile groups under lateral loads. Proceedings of the 9th International Conference on Soil
Mechanics and Foundation Engineering, Tokyo, Japan, July 10
15th, 1977, 18788.
Basack, S. and S. Dey. 2012. Inuence of relative pile-soil stiffness and
load eccentricity on single pile response in sand under lateral cyclic
loading. Geotechnical and Geological Engineering 30(4): 73751.
doi:10.1007=s10706-011-9490-1
Basack, S. and S. Sen. 2014. Numerical solution of single piles
subjected to pure torsion. Journal of Geotechnical and Geoenvironmental Engineering 140(1): 7490. doi:10.1061=(asce)gt.1943-5606.
0000964
Batchelor, G. K. 1967. An Introduction to Fluid Dynamics. Cambridge,
UK: Cambridge University Press.
Bhattacharya, S., J. A. Cox, D. Lombardi, and D. M. Wood. 2012.
Dynamics of offshore wind turbines supported on two foundations. Proceedings of the ICE, Geotechnical Engineering 166(2):
15969. doi:10.1680=geng.11.00015
Bienen, B., J. Duhrkop, J. Grabe, M. F. Randolph, and D. J. White.
2012. Response of piles with wings to monotonic and cyclic lateral
loading in sand. Journal of Geotechnical and Geoenvironmental

M. Arshad & B. C. OKelly


Engineering
138(3):
36475.
doi:10.1061=(asce)gt.19435606.0000592
Bisoi, S. and S. Haldar. 2014. Dynamic analysis of offshore wind turbine in clay considering soilmonopiletower interaction. Soil
Dynamics and Earthquake Engineering 63: 1935. doi:10.1016=
j.soildyn.2014.03.006
Blanco, M. I. 2009. The economics of wind energy. Renewable and Sustainable Energy Reviews 13(67): 137282.
Broms, B. 1964. Lateral resistance of piles in cohesionless soils. Soil
Mechanics and Foundation Engineering Division, ASCE 90(3):
12356.
Budhu, M. 2011. Soil Mechanics and Foundations. New York: John
Wiley & Sons.
Byrne, B. W. and G. T. Houlsby. 2003. Foundations for offshore wind
turbines. Philosophical Transactions of the Royal Society of London
361(1813): 290930.
Carswell, W., S. R. Arwade, D. J. DeGroot, and M. A. Lackner. 2015.
Soilstructure reliability of offshore wind turbine monopile foundations. Wind Energy 18(3): 48398. doi:10.1002=we.1710
Chai, J.-C. and N. Miura. 2002. Trafc-load-induced permanent deformation of road on soft subsoil. Journal of Geotechnical and Geoenvironmental Engineering 128(11): 90716. doi:10.1061=(asce)10900241(2002)128:11(907)
Chakrabarti, S. K. 2005. Handbook of Offshore Engineering. London,
UK: Elsevier.
Das, B. M. 2008. Advanced Soil Mechanics. New York: CRC Press.
Davisson, M. T. 1970. Lateral Load Capacity of Piles, Vol. 333, 10412.
Washington, DC: Highway Research Record.
De Vries, W. E. 2007. Project UpWind WP4 deliverable D4.2.1: Assessment of bottom-mounted support structure types with conventional design stiffness and installation techniques for typical
deep water sites. http://www.upwind.eu/Publications/~/media/
UpWind/Documents/Publications/4%20-%20Offshore%20Foundations/UpwindWP4D421%20Assessment%20of%20bottommounted%20support%20structure%20types.ashx (accessed March
15th, 2015).
De Vries, W. E. and J. van der Tempel. 2007. Quick monopile design.
Proceedings of the European Offshore Wind Conference & Exhibition, Berlin, Germany, December 46th, 2007.
DIN (Deutsches Institut fur Normung). 2005. Baugrund Sicherheitsnachweise im Erd und Grundbau. DIN 1054: 2005. DIN: Berlin,
Germany.
http://www.baunormenlexikon.de/Normen/DIN/
DIN%201054/1b69b621-74a0-4c33-8b29-a82a24afd4d4 (accessed
March 15th, 2015).
DNV (Det Norske Veritas). 2011. Design of offshore wind turbine
structures. DNV-OS-J101, DNV, Oslo, Norway.
Dobry, R., E. Vicenti, M. J. ORourke, and J. M. Roesset. 1982. Horizontal stiffness and damping of single piles. Journal of Geotechnical Engineering Division, ASCE 108(3): 43959.
Duhrkop, J. 2009. On the inuence of expanders and cyclic loads on the
deformation behavior of lateral stressed piles in sand. PhD Thesis,
Hamburg University of Technology.
Dyson, G. J. and M. F. Randolph. 2001. Monotonic lateral loading of
piles in calcareous sand. Journal of Geotechnical and Geoenvironmental Engineering 127(4): 34652. doi:10.1061=(asce)10900241(2001)127:4(346)
EWEA (European Wind Energy Association). 2011. Design limits and
solutions for very large wind turbines. UpWind project. http://
www.ewea.org/leadmin/ewea_documents/documents/upwind/
21895_UpWind_Report_low_web.pdf (accessed March, 15th 2015).
Fischer, T. 2011. Executive summary Upwind project WP4: Offshore
foundations and support structures. http://www.upwind.eu/Publications/~/media/UpWind/Documents/Publications/4%20%20Offshore%20Foundations/WP4_Executive_Summary_Final.ashx (accessed March, 15th 2015).
Gavin, K. G. and OKelly, B. C. 2007. Effect of friction fatigue
on pile capacity in dense sand. Journal of Geotechnical and

Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

Analysis and Design of Offshore Monopile Foundations

21

Geoenvironmental Engineering 133(1): 6371. doi:10.1061=


(asce)1090-0241(2007)133:1(63)
GL (Germanischer Lloyd). 2005. Guideline for the Certication of Offshore Wind Turbines. Hamburg, Germany: GL.
Goryunov, B. F. 1973. Analysis of piles subjected to the combined
action of vertical and horizontal loads (discussion). Soil Mechanics
and Foundation Engineering 10(1): 1013. doi:10.1007=bf01706631
Guo, W. D. 2006. On limiting force prole, slip depth and response of
lateral piles. Computer and Geotechnics 33(1): 4767. doi:10.1016=
j.compgeo.2006.02.001
Guo, W. D. 2013. Pu is subscripted-based solutions for slope stabilizing
piles. International Journal of Geomechanics 13(3): 292310.
doi:10.1061=(asce)gm.1943-5622.0000201
Haiderali, A., U. Cilingir, and G. Madabhushi. 2013. Lateral and axial
capacity of monopiles for offshore wind turbines. Indian Geotechnical Journal 43(3): 18194. doi:10.1007=s40098-013-0056-4
Haritos, N. 2007. Introduction to the analysis and design of offshore
structures An overview. Electronic Journal of Structural
Engineering 7(Special Issue: Loading on Structures): 5565.
Hornych, P., J. F. Corte, and J. L. Paute. 1993. Etude des deformations
permanentes sous chargements repetes de trois graves non traitees.
Bulletin de Liaison des Laboratoires des Ponts et Chaussees, Presse
de lENPC, Paris, France 184: 4555.
Huurman, M. 1996. Development of trafc induced permanent strains
in concrete block pavements. Heron 41(1): 2952.
IEC (International Electrotechnical Commission). 2009. Wind turbines
Part 3: Design requirements for offshore wind turbines. IEC
614003: 2009, International Electrotechnical Commission, Geneva, Switzerland.
Igoe, D., K. Gavin, and B. OKelly. 2010. Field tests using an instrumented model pipe pile in sand. Proceedings of the Seventh International Conference on Physical Modelling in Geotechnics,
Zurich, Switzerland, June 28thJuly 1st, 2010, ed. S. Springman,
J. Laue, and L. Seward, Vol. 2, 775780. Leiden, The Netherlands:
CRC Press.
Igoe, D., K. Gavin and B. OKelly. 2013. An investigation into the use
of push-in pile foundations by the offshore wind sector. International Journal of Environmental Studies 70(5): 77791.
doi:10.1080=00207233.2013.798496
Igoe, D. J. P., K. G. Gavin, B. C. OKelly, and B. Byrne. 2013. The use
of in-situ site investigation techniques for the axial design of offshore piles. Proceedings of the 4th International Conference on
Geotechnical and Geophysical Site Characterization, Pernambuco, Brazil, September 1821st, 2012, ed. R. Q. Coutinho and
P. W. Mayne. Vol. 2, 112329. CRC Press.
Irvine, J. H., P. G. Allan, B. G. Clarke, and J. R. Peng. 2003. Improving the lateral stability of monopile foundations. Proceedings of
the BGA International Conference on Foundations: Innovations,
Observations, Design and Practice, Dundee, UK, September 2
5th, 2003, ed. T. A. Newson, 37180. London: Thomas Telford.
Ismael, N. F. 1990. Behavior of laterally loaded bored piles in cemented
sands. Journal of Geotechnical Engineering, ASCE 116(11): 1678
99. doi:10.1061=(asce)0733-9410(1990)116:11(1678)
Jaimes, O. G. 2010. Design concepts for offshore wind turbines:
A technical and economic study on the trade-off between stall
and pitch controlled systems. MSc Thesis, Delft University of
Technology.
Jain, N. K., G. Ranjan, and G. Ramasamy. 1987. Effect of vertical load
on exural behaviour of piles. Geotechnical Engineering: Journal of
the Southeast Asian Geotechnical Society 18(2): 185204.
Jang, J. J. and G. J. Shinn. 1999. Analysis of maximum wind force for
offshore structure design. Journal of Marine Science and Technology 7(1): 4351.
Jonkman, J., S. Buttereld, W. Musial, and G. Scott. 2009. Denition
of a 5-MW reference wind turbine for offshore system

development. Technical Report NREL=TP-50038060, National


Renewable Energy Laboratory (NREL), Colorado, USA.
Journee, J. M. J. and W. W. Massie. 2001. Offshore Hydrodynamics, 1st
ed. Delft, The Netherlands: Delft University of Technology.
Karg, C. 2007. Modelling of strain accumulation due to low level vibrations
in granular soils. PhD Thesis. Ghent University, Ghent, Belgium.
Karthigeyan, S., V. V. G. S. T. Ramakrishna, and K. Rajagopal. 2006.
Inuence of vertical load on the lateral response of piles in sand.
Computers and Geotechnics 33(2): 12131. doi:10.1016=j.compgeo.2005.12.002
Karthigeyan, S., V. V. G. S. T. Ramakrishna, and K. Rajagopal. 2007.
Numerical investigation of the effect of vertical load on the lateral
response of piles. Journal of Geotechnical and Geoenvironmental
Engineering 133(5): 51221. doi:10.1061=(asce)1090-0241(2007)
133:5(512)
Klinkvort, R. T. and O. Hededal. 2013. Lateral response of monopile
supporting an offshore wind turbine. Proceedings of the ICE, Geotechnical Engineering 166(2): 14758. doi:10.1680=geng.12.00033
Kopp, D. R. 2010. Foundations for an offshore wind turbine. MSc
Thesis, Massachusetts Institute of Technology.
Kuo, Y.-S., M. Achmus, and K. Abdel-Rahman. 2012. Minimum
embedded length of cyclic horizontally loaded monopiles. Journal
of Geotechnical and Geoenvironmental Engineering 138(3): 35763.
doi:10.1061=(asce)gt.1943-5606.0000602
LeBlanc, C. 2009. Design of offshore wind turbine support structures:
Selected topics in the eld of geotechnical engineering. PhD Thesis, Aalborg University, Aalborg, Denmark.
LeBlanc, C., G. T. Houlsby, and B. W. Byrne. 2010. Response of stiff
piles in sand to long-term cyclic lateral loading. Geotechnique
60(2): 7990. doi:10.1680=geot.7.00196
Lekarp, F. and A. Dawson. 1998. Modelling permanent deformation
behaviour of unbound granular materials. Construction and Building Materials 12(1): 918. doi:10.1016=s0950-0618(97)00078-0
Lentz, R. W. and G. Y. Baladi. 1981. Constitutive equation for permanent strain of sand subjected to cyclic loading. Transportation
Research Record 810: 5054.
Lesny, K. and J. Wiemann. 2005. Design aspects of monopiles in German offshore wind farms. Proceedings of the 1st International
Symposium on Frontiers in Offshore Geotechnics. Perth, Australia, September 1921st, 2005, ed. S. Gourvenec and M. Cassidy,
38389. Leiden, The Netherlands: Balkema.
Li, D. and E. Selig. 1996. Cumulative plastic deformation for negrained subgrade soils. Journal of Geotechnical Engineering, ASCE
122(12): 100613. doi:10.1061=(asce)0733-9410(1996)122:12(1006)
Li, Z., S. K. Haigh, and M. D. Bolton. 2010. Centrifuge modelling of
mono-pile under cyclic lateral loads. Proceedings of the 7th International Conference on Physical Modelling in Geotechnics, Zurich, Switzerland, 28th June1st July 2010, ed. S. Springman, J.
Laue, and L. Seward, Vol. 2, 96570. Leiden, The Netherlands:
CRC Press.
Lin, S.-S. and J.-C. Liao. 1999. Permanent strains of piles in sand due
to cyclic lateral loads. Journal of Geotechnical and Geoenvironmental Engineering 125(9): 798802. doi:10.1061=(asce)10900241(1999)125:9(798)
Little, R. L. and J.-L. Briaud. 1988. Full scale cyclic lateral load tests on
six single piles in sand. Miscellaneous paper GL-8827, Geotechnical Division, Civil Engineering Department, Texas A&M University, College Station, TX, USA.
Lombardi, D., S. Bhattacharya, and D. M. Wood. 2013. Dynamic soil
structure interaction of monopile supported wind turbines in
cohesive soil. Soil Dynamics and Earthquake Engineering 49:
16580. doi:10.1016=j.soildyn.2013.01.015
Long, J. H. and G. Vanneste. 1994. Effects of cyclic lateral loads on
piles in sand. Journal of Geotechnical Engineering, ASCE 120(1):
22544. doi:10.1061=(asce)0733-9410(1994)120:1(225)

Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

22
Malhotra, S. 2011. Design and construction considerations for offshore
wind turbine foundations in North America. Proceedings GeoFlorida 2010: Advances in Analysis, Modeling and Design, West
Palm Beach, Florida, USA, February 2024th, 2010, ed. D. O.
Fratta, A. J. Puppala, and B. Muhunthan, Vol. 2, 153342, GSP
199. Red Hook, NY, USA: Curran Associates, Inc.
Matlock, H. 1970. Correlation for design of laterally loaded piles in soft
clay. Proceedings of the 2nd Offshore Technology Conference, 22
24th April 1970, Houston, TX, USA. Paper Number 1204. Vol. 1,
7794.
Matlock, H., and L. C. Reese 1960. Generalized solutions for laterally
loaded piles. Journal of the Soil Mechanics and Foundations
Division ASCE 86(5): 6394.
Moayed, R. Z., I. Mehdipour, and A. Judi. 2012. Undrained lateral
behavior of short pile under combination of axial, lateral and
moment loading in clayey soils. Kuwait Journal of Science and
Engineering 39(1B): 5978.
Morison, J. R., J. W. Johnson, and S. A. Schaff. 1950. The forces
exerted by surface waves on piles. Journal of Petroleum Technology
2(5): 14954. doi:10.2118=950149-g
Naughton, P. J. and B. C. OKelly. 2004. The induced anisotropy of
Leighton Buzzard sand. Proceedings Advances in Geotechnical
Engineering: The Skempton Conference, London, UK, March
28th31st, 2004, ed. R. J. Jardine, D. M. Potts, and K. G. Higgins,
Vol. 1, 55667. London, UK: Thomas Telford.
Naughton, P. J. and B. C. OKelly. 2005. Yield behavior of sand under
generalized stress conditions. Proceedings of the 16th International
Conference on Soil Mechanics and Geotechnical Engineering, Osaka,
Japan, September 1216th, 2005. IOS Press. Vol. 2, 55558.
Nicolai, G. and L. B. Ibsen. 2014. Small-scale testing of cyclic laterally
loaded monopiles in dense saturated sand. Journal of Ocean and
Wind Energy 1(4): 24045.
Niemunis, A., T. Wichtmann, and Th. Triantafyllidis. 2005. A highcycle accumulation model for sand. Computers and Geotechnics
32(4): 24563. doi:10.1016=j.compgeo.2005.03.002
OKelly, B. C. 2005. Consolidation anisotropy of some natural soft
soils. Proceedings of the International Conference on Problematic
Soils, Famagusta, North Cyprus, May 2527th, 2005, ed. H. Bilsel
and N. Zalihe, Vol. 3, 118392. North Cyprus: Eastern Mediterranean University Press.
OKelly, B. C. 2006. Compression and consolidation anisotropy of
some soft soils. Geotechnical and Geological Engineering 24(6):
171528. doi:10.1007=s10706-005-5760-0
OKelly, B. C. and P. J. Naughton. 2005a. Development of a new hollow cylinder apparatus for stress path measurements over a wide
strain range. Geotechnical Testing Journal 28(4): 34554.
doi:10.1520=gtj12252
OKelly, B. C. and P. J. Naughton. 2005b. Engineering properties of
wet-pluviated hollow cylindrical specimens. Geotechnical Testing
Journal 28(6): 57076. doi:10.1520=gtj12325
OKelly, B. C. and P. J. Naughton. 2008. Local measurements of the
polar deformation response in a hollow cylinder apparatus. Geomechanics and Geoengineering 3(4): 21729. doi:10.1080=
17486020802400981
OKelly, B. C. and P. J. Naughton. 2009. Study of the yielding of sand
under generalized stress conditions using a versatile hollow cylinder torsional apparatus. Mechanics of Materials 41(3): 18798.
doi:10.1016=j.mechmat.2008.11.002
Pappin, J. W. 1979. Characteristics of granular material for pavement
analysis. PhD Thesis, University of Nottingham.
Paute, J.-L., P. Hornych, and J. P. Benaben. 1996. Repeated load triaxial testing of granular materials in the French Network of Laboratories des Ponts et Chaussees. Flexible Pavements: Proceedings of
the European Symposium Euroex, Lisbon, Portugal, September
2022nd, 1993, ed. A. G. Corriea, 5364. Rotterdam, The
Netherlands: Balkema.

M. Arshad & B. C. OKelly


Peng, J., B. G. Clarke, and M. Rouainia. 2011. Increasing the resistance
of piles subject to cyclic lateral loading. Journal of Geotechnical
and Geoenvironmental Engineering 137(10): 97782. doi:10.1061=
(asce)gt.1943-5606.0000504
Peralta, P. and M. Achmus. 2010. An experimental investigation of
piles in sand subjected to lateral cyclic loads. Proceedings of the
7th International Conference on Physical Modelling in Geotechnics, Zurich, Switzerland, 28th June1st July 2010, ed. S. Springman, J. Laue, and L. Seward, Vol. 2, 98590. Leiden, The
Netherlands: CRC Press.
Poulos, H. G. 1988. Cyclic stability diagram for axially loaded piles.
Journal of Geotechnical Engineering, ASCE, 114(8): 87795.
doi:10.1061=(asce)0733-9410(1988)114:8(877)
Rahim, A. and R. F. Stevens. 2013. Design procedures for marine
renewable energy foundations. Proceedings of the 1st Marine
Energy Technology Symposium, Washington, DC, April 10
11th, 2013, 10.
Ramasamy, G. 1974. Flexural behaviour of axially and laterally loaded
individual piles and groups of piles. PhD Thesis, Indian Institute
of Science, Bangalore.
Randolph, M. F. 1981. The response of exible piles to lateral loading.
Geotechnique 31(2): 24759. doi:10.1680=geot.1981.31.2.247
Rani, S. and A. Prashant. 2014. Estimation of the linear spring constant for a laterally loaded monopile embedded in nonlinear soil.
International
Journal
of
Geomechanics.
doi:10.1061=
(ASCE)GM.1943-5622.0000441. Online Publication Date: 29th
August 2014.
Reese, L. C, W. R. Cox, and F. D. Koop. (1974). Analysis of laterally
loaded piles in sand. Proceedings of the 6th Annual Offshore Technology Conference, Houston, Texas, May 68th, 1974, 47384.
Reese, L. C, W. R. Cox, and F. D. Koop. (1975). Field testing
and analysis of laterally loaded piles in stiff clay. Procedings
of the 7th Annual Offshore Technology Conference. Paper No.
2312, Houston, Texas, May 58th, 1975, 671690. doi: 10.4043/
2312-MS.
Richwien, W., K. Lesny, and J. Wiemann. 2002. Bau- und umwelttechnische Aspekte von Off-shore Windenergieanlagen (Construction
and environmental aspects of offshore wind turbines). Gigawind
Annual Report 2001, Chapter 6: Support structure Foundation,
3646. Leibniz Universitat Hannover, Germany.
S ahin, A. D. 2004. Progress and recent trends in wind energy. Progress
in Energy and Combustion Science 30(5): 50143. doi:10.1016=
j.pecs.2004.04.001
Sorochan, E. A. and V. I. Bykov. 1976. Performance of groups of
cast-in place piles subjected to horizontal loading. Soil Mechanics
and Foundation Engineering 13(3): 15761. doi:10.1007=
bf01705310
St. Denis, M. and W. J. Pierson. 1953. On the motions of ships in confused seas. Society of Naval Architects and Marine Engineers
Transactions 61: 280354.
Sweere, G. T. H. 1990. Unbound granular bases for roads. PhD Thesis,
Delft University of Technology.
Terzaghi, K. 1955. Evaluation of coefcients of subgrade reaction.
Geotechnique 5(4): 297326.
Timmerman, D. H. and T. H. Wu. 1969. Behavior of dry sands under
cyclic loading. Journal of the Soil Mechanics and Foundations
Division ASCE 95(4): 1097114.
Tomlinson, M. J. 2001. Foundation Design and Construction, 7th edn.
Harlow, England: Pearson Education.
Tong, W. 2010. Wind Power Generation and Wind Turbine Design.
Southampton, UK: WIT Press.
van der Tempel, J. 2006. Design of support structures for offshore wind
turbines. PhD Thesis, Delft University of Technology.
Verdure, L., J. Garnier, and D. Levacher. 2003. Lateral cyclic loading
of single piles in sand. International Journal of Physical Modelling
in Geotechnics 3(3): 1728.

23

Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

Analysis and Design of Offshore Monopile Foundations


Wang, Q. and P. V. Lade. 2001. Shear banding in true triaxial tests and
its effect on failure in sand. Journal of Engineering Mechanics
127(8): 75461. 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, Technical University Dresden.
Wichtmann, T., A. Niemunis, and Th. Triantafyllidis. 2009. Validation
and calibration of a high-cycle accumulation model based on cyclic triaxial tests on eight sands. Soils and Foundations 49(5): 711
28. doi:10.3208=sandf.49.711
Wichtmann, T., H. A. Rondon, A. Niemunis, Th. Triantafyllidis, and
A. Lizcano. 2010. Prediction of permanent deformations in pavements using a high-cycle accumulation model. Journal of Geotechnical and Geoenvironmental Engineering 136(5): 72840.
doi:10.1061=(asce)gt.1943-5606.0000275
Yasin, S. J. M. and F. Tatsuoka. 2000. Stress history-dependent deformation characteristics of dense sand in plane strain. Soils and
Foundations 40(2): 7798. doi:10.3208=sandf.40.2_77
Zhu, B., B. W. Byrne, and G. T. Houlsby. 2013. Long-term lateral cyclic response of suction caisson foundations in sand. Journal of
Geotechnical and Geoenvironmental Engineering 139(1): 7383.
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
pile columns to horizontal loads. Soil Mechanics and Foundation
Engineering 15(1): 1622. doi:10.1007=bf02145324

Appendix

Fig. A1. Pertinent stiffness proles for the medium dense sand
deposit considered in the OWT monopile foundation design
example.

Potrebbero piacerti anche