Sie sind auf Seite 1von 16

Soil Dynamics and Earthquake Engineering 49 (2013) 165180

Contents lists available at SciVerse ScienceDirect

Soil Dynamics and Earthquake Engineering


journal homepage: www.elsevier.com/locate/soildyn

Dynamic soilstructure interaction of monopile supported wind turbines in


cohesive soil
Domenico Lombardi a, Subhamoy Bhattacharya b,n, David Muir Wood c
a
Research Student, Department of Civil Engineering, University of Bristol, Bristol, UK
b
Chair in Geomechanics, University of Surrey, Department of Civil Engineering, Guildford, UK
c
Professor of Geotechnical Engineering, University of Dundee, Dundee, UK

a r t i c l e i n f o abstract

Article history: Offshore wind turbines supported on monopile foundations are dynamically sensitive because the
Received 31 July 2012 overall natural frequencies of these structures are close to the different forcing frequencies imposed
Received in revised form upon them. The structures are designed for an intended life of 25 to 30 years, but little is known about
20 January 2013
their long term behaviour. To study their long term behaviour, a series of laboratory tests were
Accepted 23 January 2013
conducted in which a scaled model wind turbine supported on a monopile in kaolin clay was subjected
Available online 20 March 2013
to between 32,000 and 172,000 cycles of horizontal loading and the changes in natural frequency and
Keywords: damping of the model were monitored. The experimental results are presented using a non-
Offshore dimensional framework based on an interpretation of the governing mechanics. The change in natural
Wind turbine
frequency was found to be strongly dependent on the shear strain level in the soil next to the pile.
Monopile
Practical guidance for choosing the diameter of monopile is suggested based on element test results
Cyclic loading
Dynamics using the concept of volumetric threshold shear strain.
Long-term performance Crown Copyright & 2013 Published by Elsevier Ltd. All rights reserved.
Laboratory test
Clay

1. Introduction turbines (e.g. the approach suggested by DNV-OS-J101 [15] or


IEC61400-1 [21]) are based on the methods originally developed
Offshore wind turbines are providing an increasing proportion for the offshore oil and gas industry [4]. Fig. 1 shows a typical
of wind energy generation capacity because these sites are monopile supported wind turbine and a pile supported xed
characterised by stronger and more stable wind conditions than offshore jacket structure. There are, however, obvious differences
comparable onshore sites. Offshore sites also have a higher between those two types of foundations.
capacity factor (the ratio of the actual amount of power produced Piles for offshore structures are typically 60110 m long and 1.8
over a period of time to the rated turbine power) when compared 2.7 m diameter. By contrast, monopiles for offshore wind turbines are
to equivalent onshore sites. commonly 3040 m long and 3.56 m diameter. Degradation in the
The design and construction of foundations for offshore upper soil layers resulting from cyclic loading is less severe for
turbines are challenging because of the harsh environmental offshore jacket piles which are signicantly restrained from pile head
conditions and as a result provide a focus of major research in rotation causing lower pile head deections. However, the over-
Europe, see for example Achmus et al. [1], Kuhn [29], Kuo et al. turning moments generated in the jacket superstructure are resisted
[30], Bhattacharya et al. [5]. Different types of foundations have by pairs of equal and opposite axial resultants in the piles. Such cyclic
been proposed: including monopile, gravity base, jacket, suction axial loads can produce a loss of shaft capacity because of the
caisson and oating systems. However, most of the offshore development of friction fatigue down the piles. Monopiles are free-
turbines currently in operation (UK Round 1 development) are headed which encourages more pile head deection. A design method
supported on driven monopiles. The choice of monopiles results using a beam on non-linear Winkler springs (py method in API code
from their simplicity of installation and the proven success of [4] or DNV code) may be used to obtain pile head deection under
driven piles in supporting offshore oil and gas infrastructures. The cyclic loading, but its use is limited for wind turbines because:
available methods for designing monopiles for offshore wind

(a) the widely used API model is calibrated against response of a


n few small diameter piles (length to diameter ratio of 30 to 50)
Corresponding author.
E-mail addresses: S.Bhattacharya@surrey.ac.uk, subjected to small numbers of cycles (maximum 200 cycles)
subhamoy.bhattacharya@gmail.com (S. Bhattacharya). suited for offshore xed platform applications, e.g. Matlock [38],

0267-7261/$ - see front matter Crown Copyright & 2013 Published by Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.soildyn.2013.01.015
166 D. Lombardi et al. / Soil Dynamics and Earthquake Engineering 49 (2013) 165180

Nomenclature N number of load cycles


P net horizontal load
a parameter in the rational function tting t time
CSR cyclic stress ratio tw wall thickness of pile
D pile diameter V vertical load
E Youngs modulus of pile y distance between foundation level and application of
ff forcing frequency P
fn natural frequency a parameter in the logarithm tting
G shear modulus of soil D parameter in the rational function tting
Gmax shear modulus of soil at small strains ga average shear strain in the soil
Gsec secant shear modulus of soil gtl linear shear strain threshold
I second moment of area of pile gtv volumetric shear strain threshold
kh horizontal coefcient of soil permeability d lateral deection of pile head
KL stiffness of transverse spring ep strain in pile wall thickness
KR stiffness of rotational spring es average strain in soil
KV stiffness of vertical spring l parameter in the rational function tting
L penetration depth of pile x damping ratio of model
M external moment acting at the pile head s0v effective vertical on the soil at the same depth
MN secant modulus of the py curve after Nth as above
M1 secant modulus of the py curve after 1st cycle sy pile yield stress

ONeill and Murchison [42], Poulos and Hull [44], Reese et al. monopile in sandy soil will actually increase as a result of
[47,48]. In contrast, for a real offshore wind turbine, the length densication of the soil next to the pile.
to diameter ratio of piles is of the order of 4 to 8 and 107108 (d) The ratio of horizontal load (P) to vertical load (V) is very high
cycles of lateral and moment loading are expected over a in offshore wind turbines when compared with xed jacket
lifetime of 2025 years. structures. Therefore, the monopiles experience disproportio-
(b) It can be shown that the calibrated py curves used in the API nately higher moment loading in comparison to a jacket pile.
and DNV codes are based on exible pile behaviour where the This more extreme loading condition was not taking into
pile is expected to fail by formation of plastic hinges (struc- account during the calibration of the API and DNV py curves.
tural failure of piles). On the other hand, the squat nature of
monopiles makes them sufciently rigid that the formation of
plastic hinges is not expected. Rather, a monopile will rotate A similar problem of cyclic degradation of the soil surrounding a
like a rigid body (potentially including some reverse toe-kick) relatively short pile (2030 m) was encountered in designing the
and the soil next to the pile may fail. oating offshore platforms for the North Sea Alvheim eld [8]:
(c) under cyclic loading, the API or DNV model always predicts Mechanisms included post-holing and possible jetting action due to
degradation of foundation stiffness in sandy soil. However, the one-way cyclic loading on the anchoring piles. These near-
recent work by Bhattacharya and Adhikari [7], Cuellar et al. surface effects, Bhattacharya et al. [8] are much more signicant for
[13], LeBlanc [32] suggested that the foundation stiffness for a the shorter monopiles, affecting a greater proportion of their length.

Fig. 1. Typical monopile supported wind turbine and a xed offshore jacket structure supported on piles.
D. Lombardi et al. / Soil Dynamics and Earthquake Engineering 49 (2013) 165180 167

The long term performance of offshore wind turbines involves [12], Little and Briaud [34], Kramer and Heavey [27] that one-way
the following two issues: loading develops more soil deformation and consequently more
change in foundation stiffness. Research on pilesoil interactions
without the effects of superstructure has also been recently
1. Change (degradation/stiffening) of soil stiffness leads to carried out by Cuellar et al. [13], Li et al. [33], LeBlanc [32],
changes in the natural frequency of the system, with the Achmus et al. [1] where it was observed that the pile-soil stiffness
potential for unplanned system resonances and consequent changes with cycles of loading. Research considering large num-
excessive cyclic displacements. bers of cycles of loading on a soil sample was investigated by
2. Degradation of soil stiffness may lead to permanent displace- Wichtmann et al. [54].
ment of the turbine which may jeopardise its performance: Wind By contrast, the present work is unique in its experimental
turbines typically cannot tolerate more than 0.5 degrees tilt. consideration of the dynamics of the overall system. This paper
aims:

Offshore wind turbines are relatively new structures without


any track record of long term performance. Monitoring of a (a) to summarise the external dynamic and cyclic loading acting
limited number of installed monopile-supported wind turbines on a typical offshore wind turbine.
has indicated a gradual departure of the overall system dynamics (b) to present results of tests on scaled model wind turbine
from the design requirements (e.g. the data from Lely wind farm supported on a monopile in kaolin clay in order to understand
data [28]). Clearly, the long-term performance and the uncertain- the long term foundation performance.
ties related to dynamic response of these dynamically sensitive (c) to assess the long term performance of prototype offshore
structures need to be understood. The current codes of practice wind turbines using similitude relationships for 1-g testing of
(for example API, ISO, DNV) for the design of monopile founda- models of such turbines.
tions of offshore wind turbines recommend the application of the (d) to highlight the practical implications for choice of monopile
py curves primarily for the evaluation of lateral pile capacity in dimensions.
the ultimate limit state (ULS). The codes provide limited guidance
in predicting the change of the foundation stiffness and conse-
quently the change of natural frequency which is an important
design driver for serviceability limit state (SLS) requirements.
Vibration of a monopile will induce cyclic strains in the soil in 2. Dynamic and cyclic loading on a wind turbine
its vicinity. Under moderate-to-high amplitudes of cyclic loading
most soils change their stiffness and strength. In order to study Offshore wind turbines are characterised by a unique set of
the changes in soil stiffness due to these cyclic strains, the dynamic loading conditions (Fig. 2). The principal external excita-
developing strain in the soil around the shear zone must be taken tions are:
into consideration. Soils in offshore sites can be fully saturated
and therefore pore pressures are likely to develop as a result of
these cyclic strains. The pore pressure developed may dissipate to (a) Environmental dynamic loads arising from wind and waves.
the surrounding soil depending on factors including frequency of Fig. 3 shows the plot of power spectral density of wind and
loading, permeability of the soil and diameter of the pile. Pore wave loading around the UK coastline (particularly in the
water pressure may also develop in unsaturated soils under cyclic North Sea). The predominant wave frequency is 0.1 Hz, which
shearing. corresponds to 10 s wave period.
The dynamic excitations caused by environmental loads (wind (b) Rotor loading at a frequency which is commonly referred to as
and wave) may impose either one-way or two-way cyclic loading 1P. Fig. 3 shows the rotor frequency for a 3.6 MW wind
on the monopile. It has been observed by Chang and Whitman turbine having an operational range between 5 and 13 rpm,

Fig. 2. External loads acting on offshore wind turbine supported on a monopile foundation.
168 D. Lombardi et al. / Soil Dynamics and Earthquake Engineering 49 (2013) 165180

Fig. 3. Simplied power spectral density of the forcing frequencies applied to typical three-bladed 3.6 MW offshore wind turbine with an operational interval in the range
of 0.140.31 Hz (513 rpm).

Table 1
Typical values of forcing and natural frequencies for real offshore wind turbines.

Offshore wind farm and the type of Foundation and soil data Forcing frequency (ff) Natural ff/fn Design
turbine frequency, fn [Hz] approach
1P [Hz] 2P/3P [Hz]

Lely A2 Turbine (Netherlands) Pile passes through soft layer to 0.53 1.06 (2P) 0.63nn From 0.84 Soft-stiff
Nedwind500KW/41 Zaaijer [57] stiff sandy layer to 1.68
Irene Vorrink (Netherlands) Pile passes through soft layer to 0.45 1.35 (3P) 0.56nn From 0.80 Soft-stiff
Nortdtank600/43 turbine Zaaijer [57] stiffer sandy layer to 2.41
n
Blyth (UK) Vestas V66 2 MW turbine Submerged rocky outcrop 0.180.41 0.541.23 (3P) 0.41 From 0.44 Soft-stiff
Camp et al. [10] to 3.00
Sheringham Shoal (UK) Siemens Gravelly sands overlaying rm to 0.080.22 0.240.66 (3P) 0.850.96n From 0.09 Stiff-stiff
SWT-3.6-107 turbine Hamre et al. [18] stiff sandy clays to 0.69
Kentish Flat (UK) Vestas V90 3 MW turbine. Soft and stiff clay 0.140.31 0.420.90 (3P) 0.38n From 0.37 Soft-stiff
Villalobos (2006) to 2.37
n
North Hoyle (UK) Vestas V80 2 MW turbine. The upper seabed layer comprises 0.150.32 0.450.96 (3P) 0.35 From 0.43 Soft-stiff
Carter [11] variations to 2.74
of sand and clay layers. Below is
mudstone/sandstone

n
Estimated based on Adhikari and Bhattacharya [2].
nn
Measured and reported in the literature.

i.e. 0.140.31 Hz. In the power spectral density plot the 1P They correspond to three different design approaches namely: soft-
frequency appears as a band. soft (natural frequency o1P), soft-stiff (natural frequency between
(c) The blade passing frequency (3P or 2P for a three-bladed or 1P and 2P or 3P) and stiff-stiff (natural frequency 42P or 3P). The
two-bladed turbine, respectively) is a forced loading gener- most common design, used for example in the Round 1 UK develop-
ated from the effect of wind deciency that occurs as each ment, is soft-stiff, which implies that the natural frequency lies
blade passes through the shadow of the tower. Fig. 3 shows between 1P and 3P.
the blade passing frequency for the 3.6 MW wind turbine The design procedure requires an accurate evaluation of the
generator. natural frequency, which is dependent on the support condition
(i.e. the stiffness of the foundation), which in turn relies on the
From Fig. 3 it may be observed that, in order to avoid the strength and stiffness of the surrounding soil. Furthermore, it
resonance of the system, the designed frequency of the overall system should be ensured that throughout the operational life the natural
must be kept away from the frequency content of applied loads. frequency of the system does not come close to any forcing
Specically, DNV [14] suggests that the natural frequency of the wind frequencies: This would lead to amplication of the dynamic
turbine should be at least 710% away from the 1P and 2P/3P response of the turbines, leading to larger tower deections
frequencies. Bearing these considerations in mind, there are three and/or rotations beyond the 0.5 degrees tilt which can typically
possible ranges in which the natural frequency of the system may lie. be tolerated.
D. Lombardi et al. / Soil Dynamics and Earthquake Engineering 49 (2013) 165180 169

Table 1 lists the details of six types of turbines (Nedwind, 3. Experimental apparatus
Nordtank, Seimens, Vestas V66, Vestas V80 and Vestas V90) from
six different locations. Dynamic measurements were carried out only The experimental investigation was conducted using the
on two turbines (Lely and Irene Vorrink wind farm). For the BLADE facilities (Bristol Laboratory for Advanced Dynamics Engi-
remaining four the assessment of the natural frequency is based on neering) at the University of Bristol. Tests were carried out on a
the calibrated mathematical model developed by Adhikari and scale wind turbine model supported on a monopile. The turbine
Bhattacharya [2,3], Bhattacharya and Adhikari [7]. Table 1 shows that was subjected to up to 172,000 cycles of 3P loading. The experi-
the natural frequency of these structures is of the same order as the mental setup is shown in Fig. 4.
excitation frequencies imposed by the external dynamic loads, The model turbine replicates the 3 MW Vestas V90 turbine
reinforcing the need for dynamic considerations in the design process. having a notional scale of 1:100. A homogeneous soil prole of
soft speswhite kaolin clay was used for the model tests. The soil
was prepared from slurry by mixing kaolin powder with de-
ionised water at a moisture content of about twice the liquid
limit. This slurry was then consolidated in a cylindrical concrete
tube (diameter 600 mm, height 600 mm). The small strain shear
modulus (Gmax) of the clay was assessed by a series of bender
element tests placed in the tube and an average value of 6 MPa
was measured. The undrained shear strength (Su) was estimated
from the moisture content using correlations and an average
value of 14 kPa was estimated. More details are given by Lom-
bardi [35].
The environmental dynamic loads were modelled using an
electro-dynamic actuator xed to the laboratory strong wall and
connected to the model wind turbine tower. The force (P) applied
to the wind turbine could be constantly monitored by a force
sensor. An electric motor powered by a DC supply was used to
rotate the blades to model the 1P loading and also provided

Fig. 4. Physical model of offshore wind turbine. Fig. 6. Usefulness of scaling law.

Fig. 5. (a) Model set-up and instrumentation used in the experimental investigation. (b) Impedance of the actuator.
170 D. Lombardi et al. / Soil Dynamics and Earthquake Engineering 49 (2013) 165180

aerodynamic damping to the system. Precise measurements of turbines: these involve complex dynamic windwavefoundation
displacements and accelerations were carried out using LVDTs structure interaction and no physical modelling technique can
(linearly varying differential transformers) and piezoelectric simultaneously satisfy all the interactions at a single scale. Ideally,
accelerometers at two locations (pile head and at the top of the a wind tunnel combined with a wave tank on a geotechnical
tower). The acceleration responses of the physical model were centrifuge would serve the purpose but this is unfortunately not
recorded in two orthogonal directions along the tower. Fig. 5a feasible. Special consideration is required when interpreting the
illustrates the model set-up. In order to investigate the dynamic test results. As dynamic soil structure interaction of wind turbines
response of the model, tests were performed at selected excita- are being studied, stiffness of the system is a top priority.
tion frequencies and amplitudes (Table 5). Every physical process may be expressed in terms of dimen-
The monopile was installed by controlled jacking into the clay. sionless groups and the fundamental aspects of the governing
When a pile is driven in soft clay, the soil around the pile is physical processes as encapsulated in these dimensionless groups
remoulded and undergoes high shear strains generating positive must be preserved in the design of model tests. Derivations of
excess pore pressure around the pile. In order to allow the some aspects of the scaling laws for these experiments can be
dissipation of this pore water pressure, initial measurements of found in Bhattacharya et al. [6]. In this work the following
the dynamic properties of the wind turbine model were carried physical mechanisms and parameters are considered important:
out after a pause of 40 min. The assessment of this time required
for the dissipation was based on the method suggested by
Randolph [46] using the theory of radial consolidation. (a) Pile geometry. Because of their low length to diameter ratio
The initial dynamic properties of the model (i.e. frequency and (L/D), prototype monopiles are likely to exhibit rigid
damping of the overall system) were measured using a free behaviour.
vibration test, also known in the literature as a snap back test. (b) Repeated cyclic shear strain. The strains in the soil around a
The snap back motion was generated by the impact of an impulse laterally loaded pile control the degradation of soil stiffness.
force hammer and the response signal was recorded in the time (c) Cyclic stress ratio, (CSR). The strains in the soil depend on the
domain by accelerometers. ratio of the shear stress to the vertical effective stress at a
Following the initial measurement of dynamic properties, the particular depth. It is shown in Appendix A and Bhattacharya
electro-dynamic actuator was connected to the model. The model et al. [6] that CSR can be represented by the non-dimensional
turbine was then subjected to cyclic loading for a chosen time group (P/GD2) (see denitions in the footnote of Table 1).
interval (typically 5000 cycles) by spinning the rotor blades (1P (d) Rate of loading. Generation and dissipation of pore pressure
loading) and using the actuator (3P loading). The dynamic proper- are inuenced by several parameters. The time (t) for sig-
ties were then evaluated at the end of the chosen number of cycles nicant pore water pressure dissipation will be inversely
through another snap back test. During these measurements the proportional to the permeability of the soil (kh) and directly
actuator was detached from the model and the DC motor control- proportional to a characteristic length, for example monopile
ling the 1P loading was also stopped. The actuator spring otherwise diameter (D). Time (t) is inversely proportional to the forcing
provides unwanted impedance to the system (Fig. 5b). frequency (ff), so that (kh/ffD) is a simple relevant non-
dimensional group providing rst order similarity.
(e) Frequencies of loading and system response. The dynamic
4. Similitude relationships response is strongly inuenced by the relationship between
these frequencies.
Derivation of the correct scaling laws constitutes the rst step
in an experimental study. The similitude relationships are essential
for interpretation of the experimental data and for scaling up the Table 2 presents a set of dimensionless groups, along with
results to real prototypes. There are two ways to scale up the their physical meaning, considered in this study to model the
model test results as shown in Fig. 6. The rst is to use standard dynamics of the system. Other dimensionless groups pertinent to
tables for scaling and multiply the model observations by the scale study different aspects of the problem can be derived, see for
factor to predict the prototype response, see for example Muir example [6]. The model can be compared with two prototypes:
Wood [39]. The alternative is to study the underlying mechanics/ Sheringham Shoal and Kentish Flat. The properties of the founda-
physics of the problem based on the model tests, recognising that tions are given in Table 3 while the dimensionless groups for
not all the interactions can be scaled accurately in a particular test. model and prototype are presented in Table 4.
Once the mechanics/physics of the problem are understood, the A few points may be noted:
prototype response can be predicted through analytical and/or
numerical modelling in which the physics/mechanics discovered
will be implemented in a suitable way. The second method is 1. The soil strength (Su) does not enter in any of the dimension-
particularly useful for study of the dynamics of offshore wind less groups. This is because of the fact that the present

Table 2
Dimensionless groups considered in this study.

Physical mechanism Dimensionless group


 
Strain eld in the soil next to the pile, cyclic stress ratio (CSR) in the soil next to the pile, for details see Bhattacharya et al. [6] P
GD2
 
Rate of loading kh
f D
 f
System dynamics (relative spacing of forcing to the natural frequency of the system) ff
fn

P represents the total equivalent horizontal load acting on the wind turbine at a distance y from the foundation level. Essentially, P is also the net shear acting on the
monopile. D is the diameter of the monopile. G, kh are the shear modulus and horizontal coefcient of soil permeability respectively. ff is the forcing frequency, fn is the
natural frequency of the system.
D. Lombardi et al. / Soil Dynamics and Earthquake Engineering 49 (2013) 165180 171

Table 3
Properties of monopile and soil prole for physical model and prototypes.

Property Model Prototype (Sheringham Shoal) Prototype (Kentish Flat)

Monopile Material Dural alloy Steel Steel


diameter 22 mm 4.75.7 m 4.3 m
Length 500 mm 2337 m 28 m
wall thickness 1.3 mm 55 mmnn 45 mmnn
Youngs modulus 70 GPa 210 GPa 210 GPa
Mass 61.5 g 145283 t 132 t
Yield stress 250 MPa 550 MPa 550 MPa

Soil prole Soil type Soft clay Firm to stiff clayn London clay
Plasticity Index 31% 30%n 41%
Shear modulus 6 MPa 448 MPan (60 MPa)
Horizontal permeability 10  9 m/s 10  9 m/s 10  9 m/s

n
Hamre et al. [18].
nn
Approximate value.

Table 4
Dimensionless groups for model and prototype.

Dimensionless Physical model Prototype (Sheringham Shoal) Prototype (Kentish Flat)


group
 
P 1.52  10  4 to 3.44  10  3 8.9  10  3 (P 9.5 MN, D 4.7 m, G 48 MPa) 1.05  10  3 (P 1.16 MN,
GD2
D 4.3 m, G 60 MPa)
 
kh 2.3  10  8 (ff 2 Hz) 2.3  10  9 (ff 20 Hz) Varied between 2.7  10  9 (ff 0.08 Hz, D 4.7 m) and 1.66  10  9 (ff 0.14 Hz)
ff D
3.6  10  10 (ff 125 Hz) 2.7  10  10 (ff 0.66 Hz, D 5.7 m) 2.58  10  10 (ff 0.90 Hz)
 
ff 0.644 from 0.09-0.69 Varied between 0.37 to 2.37
fn

research focus is on the long term performance i.e. service-


ability limit state (SLS) and fatigue limit state (FLS) which is
concerned with stiffness and variation of stiffness. There may
well be plastic deformations occurring but not signicant
regions of failure.
2. Fig. 7 shows a simple structural model of the overall system in
which the foundation is replaced by springs. The stiffness of
the foundation (pile) is represented by rotational stiffness (KR),
lateral stiffness (KL) and rotationallateral coupled stiffness
(KLR) which are complex functions of soil stiffness, length and
bending stiffness of the pile. The natural frequency of the
system (fn) is a function of KR, KL and KLR and therefore pile
length (L) and bending stiffness pile are incorporated in the
group (ff/fn).
3. Moment loading will also cause strain in the soil next to the
pile. For a specic turbine (e.g. Vestas V80, hub height of 80 m
and blade length of 50 m) and under a certain environmental
condition (for example, moderate wind and extreme sea state),
the net lateral load (P) acting on the foundation is a summa-
tion of various loads which can be assumed to be relatively
constant. Fig. 8(a) shows a schematic diagram showing the
main loads (P1, P2, P3) along with their point of application.
Fig. 8(b) shows the equivalent load acting on the wind turbine
model where the different loads are replaced by a single load
(P) acting at yc.

The moment (M) acting on the foundation is MP  yc (Fig. 1).


Following the similar argument as given in Appendix A, the
pile deection (d) at a depth is a function of the moment (M), the
shear modulus of the soil (G) and the pile diameter (D). The strain
in the soil next to the pile is a function of pile deection and pile
diameter D (see Eq. A.1). Therefore, the average strain eld in the
soil around a pile due to the moment (M) can be expressed as a Fig. 7. Simplied mathematical model of the whole system.
function of three parameters: es f(M,D,G)
172 D. Lombardi et al. / Soil Dynamics and Earthquake Engineering 49 (2013) 165180

Table 5
Details of experiments and loading parameters.

Loading conditions Derived parameters

ID P [N] ff [Hz] N [cycles] fn-nal/fn-initial P/GD2 [%] ff/fn Remarks

C-1 10 2 Up to 32,400 0.63 0.34 0.61.0 (1) Tests C-1, C-2 and C-3 were the initial tests which
C-2 10 20 Up to 32,400 0.91 0.34 6.77.3 suggested that drop in frequency is sensitive to
C-3 10 125 Up to 32,400 0.96 0.34 4244 2 Hz i.e. as expected ff/fn close to 1. Therefore the
C-4 7 2 Up to 32,400 0.79 0.24 0.60.8 remaining tests focussed on 2 Hz.
C-5 5 2 Up to 32,400 0.87 0.17 0.60.7 (2) The measured damping increased by 1.8 during
C-6 2 2 Up to 32,400 0.94 0.07 0.6 the tests.
C-7 0.44 2 Up to 32,400 0.98 0.02 0.6 (3) Results from the test C-9 are not considered in
C-8 1 2 Up to 172,800 0.95 0.03 0.50.6 this paper.
C-9 50 125 Up to 32,400 0.83 1.72 4244

fn-nal is the nal natural frequency of the system following the maximum number of cycles of loading.
ff-initial is the initial natural frequency of the system measured 40 min after the installation.

Fig. 8. Lateral loads acting on a wind turbine and load-equivalence for model testing.

A dimensionless group can then be obtained: cycles of loading N, and the forcing frequency ff) were varied during
  the tests. Tests C-1, C-4, C-5, C-6, C-7 and C-8 were carried out at a
M FL
es f 3
1 constant forcing frequency of 2 Hz, varying the magnitude of
GD FL2 L3
strains in the soil by using different P/GD2 values. These tests
Or alternatively: applied 32,400 cycles, apart from test C-8 which applied 172,800
      cycles. Tests C-2 and C-3 were performed at different forcing
M Pyc P yc  P
3
3
2
b 2
2 frequencies but with the same initial value of P/GD2 (i.e. 0.34%).
GD GD GD D GD
Which suggests that the strain induced in the soil next to the
pile due to the moment loading M can be expressed as a scalar 5.1. General features of the vibration
multiplier (b) of (P/GD2). As the diameter of monopiles ranges
between 3 m and 6 m and the hub height above the foundation The assessment of natural frequency was carried out in the
level can be up to 100 m, the maximum value of (b) can be 33 but frequency domain and the fast Fourier transform (FFT) was
the value depends on many factors, including direction of wind evaluated using the method suggested by Welch [53]. This
and wave loading, water depth (which will control the force due function estimates the power spectral density (PSD) of the input
to current and wave), environmental loading at that point in time (i.e. acceleration record) using Welchs averaged modied period-
(i.e. wind and the sea state), type of turbines, geometry of the ogram method of spectral estimation. The rst natural frequency
tower (uniform of tapered), geometry and orientation of the of the model corresponds to the frequency having the highest
blades. b will vary throughout the life time of the wind turbine. power spectral density value. The damping was assessed in the
However, in the model tests, the lateral load was applied at a time domain using the logarithmic decrement method.
xed point i.e. 600 mm above the foundation level giving b E27. Figs. 9 and 10 show typical results obtained from the free
vibration tests. The acceleration time history response of the
physical model installed in the clay sample is shown in Fig. 9
5. Summary of the test resuls using the continuous line.
In the same plot, the frequency response of the xed base model
Table 5 summarises the main characteristics of the tests carried is shown by a dotted line. The xed base condition was obtained
out. Three parameters (i.e. magnitude of the force P, number of by clamping the bottom part of the tower to a heavy steel bench.
D. Lombardi et al. / Soil Dynamics and Earthquake Engineering 49 (2013) 165180 173

(b) From the frequency domain analysis (Fig. 10), the xed base
natural frequency of the turbine (10.27 Hz) is reduced by
about 68% (3.3 Hz) in the presence of the foundation. The
foundation provides exibility to a wind turbine.

It may be concluded that the assessment of the dynamic


properties of the offshore wind turbines is strongly dependent
on the foundation stiffness.

5.2. Change in natural frequency and damping of the system due to


cyclic loading

Table 5 shows the change in natural frequency of the wind turbine


system as a function of the number of cycles. The change is
represented by the ratio fn-nal/fn-initial i.e. the ratio between the natural
Fig. 9. Time history of acceleration records from free vibration tests carried out on frequency measured after the maximum number of cycles have been
the physical model with or without the foundation. applied and the natural frequency in the rst cycle. This ratio provides
an indication of the level of degradation caused by the cyclic loading.
Fig. 11 plots the change in natural frequency of the overall system
with the number of cycles for different values of P/GD2 which controls
the strain level in the soil next to the pile. In all cases the ratio
representing the system dynamics, ff/fn, changed between 0.6 and 44.
Fig. 12 illustrates the change of natural frequency of the system
for different ratios of ff/fn but for a particular strain level P/
GD2 0.34%. The drop in natural frequency is higher if the forcing
frequency is close to the natural frequency of the system, which is
basically a resonance type mechanism. The natural frequency of real
wind turbines is close to the forcing frequency and therefore the
experiments demonstrate the importance of such considerations.
Fig. 13 plots the results from the test carried out up to 172,000
numbers of cycles of loading. The test showed that the overall
stiffness change is minimal, possibly due to the very low strain
level in the soil i.e. P/GD2. For values of ff/fn close to 1, this change
in frequency is strongly dependent on the dimensionless group
P/GD2. On the other hand, for high values of ff/fn (6.77.3 and 42
44 in Fig. 12) relatively high amplitudes of P/GD2 (0.34%) seem to
have little effect on the dynamic response of the model. This
Fig. 10. Frequency response from free vibration tests carried out on the physical
observation is consistent with dynamic analysis which indicates
model with or without the foundation.
that when the excitation frequency is far from the natural
frequency of the system the dynamic interaction is negligible
and the response becomes inertia dominated.
Fig. 14 plots the normalised damping values of the system for
four tests. As expected, the damping of the model increased with
the number of cycles and higher damping variations are recorded
for larger strain amplitudes in the soil.

5.3. Curve tting of the data based on the dimensionless groups

One of the aims of this paper is to link the change in natural


frequency of the wind turbine system based on the dimensionless
groups. The test results suggest that the change in natural frequency
of the physical model with number of cycles is dependent on the
strain level in the soil which is controlled by P/GD2. The data shown
in Fig. 11 can be tted by a logarithmic regression line of the form:
f N-cycles
1a lnN 3
f initial
Fig. 11. Change in frequency of the physical model with number of cycles for
different amplitudes of P/GD2. The tted line is shown by Eq. (5). Fig. 15 plots the variation of the coefcient a with P/GD2: The
value of a varies linearly with P/GD2.
Fig. 8 plots the same data in the frequency domain. Two important  
P
observations may be made: a 0:0915  4
GD2
A high value of P/GD2 produces a higher value of a, which
(a) From the pattern of decay of acceleration in Fig. 9, the presence corresponds to a much greater change in frequency. A similar form
of the soil-foundation system which typically increases damping of equation is also used to t the degradation of the Secant Young0 s
by a factor of 8 to 12. modulus of soil with number of cycles [19,20].
174 D. Lombardi et al. / Soil Dynamics and Earthquake Engineering 49 (2013) 165180

Fig. 12. Change in natural frequency fN-cycles/finitial for constant P/GD2 of 0.34%.

Fig. 13. Result from test C-8: 172,800 cycles of loading applied at constant
frequency 2 Hz and constant force amplitude 1 N.

Fig. 15. Parameter a as a function of P/GD2. The line shows the result of linear
regression tting.

Fig. 14. Change in damping with number of cycles (tests C-1, C-5, C-6, C-7).

Using Eq. (3), the degradation is disproportionately higher for


the rst few cycles, with the subsequent degradation becoming
slower as loading continues. This is in line with the observations
reported by Long and Vanneste [36] where 34 full-scale cyclic tests
on pile foundations were analysed. However, these results do not Fig. 16. Experimental results tted with Eq. (10).
take into account dynamic considerations, such as the change of
the ratio between the frequency of the excitations and the natural
frequency of the system (ff/fn) with number of cycles. As the ratio obtained using an alternative tting Eq. (7) shown in Fig. 16.
ff/fn approaches 1, the degradation is amplied. This particular
f N-cycles 1
behaviour was observed clearly in test C-1 (solid square data 1D 5
points in Fig. 11). An improved match to the experimental data is f initial 1 lnN=al
D. Lombardi et al. / Soil Dynamics and Earthquake Engineering 49 (2013) 165180 175

where the parameters D and l are related to P/GD2. Their values are observed to increase considerably due to the cycling. It was also
plotted in Fig. 17a and b and they may be described by Eqs. (6) and noticed that for the tests performed at higher P/GD2 values (i.e.
(7). 0.34% and 0.24%) and for values of ff/fn in the range 0.61.0, the
  moisture content increased from its initial value of 50% to a
P
D 11:0371  6 maximum of 100%. However, much smaller increments of moist-
GD2
ure content were observed for lower values of P/GD2 and higher
  values of ff/fn. The increase in moisture content evidently implies
P
l 72:083  2
7 reduction of the undrained strength and decrease of the stiffness
GD
of the clay. Fig. 18 shows photographs of the monopile following
a is a parameter which seems to lie in the small range from 7.8 to tests in which the entire structure tilted considerably.
8.9. An average value of 8.35 has been assumed in Fig. 16.
The experimental work suggests that the change in frequency
5.5. Rigidity of the model pile
over a wide range of strain level and also number of cycles can be
predicted by the following equation where D and l are related to
It may also be observed from Fig. 18 that the soil surrounding
P/GD2.
the aluminium alloy model pile failed and there were no evidence
f N-cycles 1 of plastic strain in the pile following the tests which ensured that
1D 8
f initial 1 lnN=8:35l rigid behaviour of the model pile.

Comparing Figs. 11 and 16, it may be observed that Eq. (8)


gives a better overall t to the entire range of the data. In 6. Discussion: choice of monopile diameter
particular, Eq. (8) may represent the model behaviour for higher
values of P/GD2 (e.g. the data indicated with the solid square) and 6.1. Comparison with other published results
for values of ff/fn close to 1.
Offshore wind turbines are dynamically sensitive structures
5.4. Moisture content of the clay next to the pile because the natural frequency of the system is very close to the
forcing frequency. The study showed the variation of the normal-
Measurements of moisture content of the clay sample next to ised frequency of the system with respect to the number of cycles
the foundation were carried out before and after the tests: It was of loading for various strain levels in the soil imposed by the cyclic

Fig. 17. Parameters for tting the degradation as function of P/GD2, Eq. (3); (a) parameter D; (b) parameter l.

Fig. 18. Details of the monopile after the tests.


176 D. Lombardi et al. / Soil Dynamics and Earthquake Engineering 49 (2013) 165180

loading. As expected, higher strain levels led to higher reductions soil will reduce. The value of gtl can be estimated from the secant
in natural frequency of the model. For a low value of (P/ shear modulus reduction curve (schematically shown in Fig. 19)
GD2 0.02%) there is practically no degradation in the natural assuming that linear threshold shear strain corresponding to a
frequency even after 105 cycles of loading. Based on the trends ratio Gsec/Gmax of 0.99. There exists another value of threshold
and in the absence of any other data, these results can be shear strain, known as volumetric threshold shear strain level
extrapolated to cycles corresponding to a fatigue limit state (FLS) [52], denoted by gtv, beyond which permanent microstructural
of 107 cycles, which is typically the number of cycles likely to be change of the fabric do occur. In other words, beyond gtv the soil is
experienced by an offshore wind turbine. degradable possibly due to pore pressure build up. In the triaxial
The test results presented in the paper are consistent with the stress space this state corresponds to the boundary between the
centrifuge tests reported by Jeanjean [23] and Doyle et al. [16] recoverable and the plastic zone [22].gtv can be assessed from
where cyclic loading was applied on pile foundations embedded the secant shear modulus reduction curve. However two values of
in kaolin clay and the effects of degradation were monitored. Gsec/Gmax ratios (namely 0.85 and 0.60) are often used depending
Jeanjean [23] proposed Eq. (9) based on observations made over on the application. The value of 0.85 represents a degradation of
1000 cycles. secant shear modulus beyond which permanent deformation is
MN 0:9 expected. On the other hand 0.6 is dened on the basis of the
9 accumulation of excess pore pressures. These two values can
M1 0:9 2:5 tanh0:7  logN
represent upper and lower bound values of gtv for the problem in
where MN and M1 are secant modulus of the py curve i.e. hand. The upper bound value will give the highest diameter that
factored pile-soil stiffness after Nth and 1st cycle respectively. may be required. On the other hand, the lower bound will give the
Overconsolidated clays (more representative of those found at minimum diameter necessary.
locations of the current generation of offshore wind farms) are While Fig. 19 shows the schematic representation of threshold
more difcult to model in a centrifuge, as in-ight consolidation is strains, Fig. 20 plots typical experimental data obtained from cyclic
slow and the clay would need to be prepared off the centrifuge triaxial tests carried out on 10 samples of Turkish clay having a
and manipulated to size. Over-consolidated clays respond by plasticity index (PI) ranging from 9 to 40 following Okur and Ansal
predominantly undrained plastic displacement which should be [41]. Fig. 20 shows an upper and lower bound for gtv which
independent of effective stress and capable of study in 1 g corresponds to Gsec/Gmax ratio of 0.85 and 0.60 respectively. The
experiments which can be used to apply large numbers of cycles experimental results clearly show that samples with higher PI tend to
in a realistic time frame. It is also easier to isolate the test rig from
the effects of external vibration so that parasitic displacements
generated by the rotation of the centrifuge itself can be avoided.
It is suggested that in the absence of other data, the present data
may be used to extrapolate to the long term performance.

6.2. P/GD2 and the determination of acceptable values based on soil


element tests

Fig. 19 shows the variation of normalised secant stiffness with


strain for fully saturated cohesive soil adapted from Vucetic [52].
It seems that there exists a strain level, the threshold linear shear
strain, gtl, [22], for which there is no stiffness degradation and the
behaviour of the soil is practically linear elastic, so that there is no
permanent microstructural changes of the fabric of the soil with
cyclic loading. As the strain level is increased beyond gtl, the soil
behaves non-linearly and as a result the secant shear modulus of
Fig. 20. Experimental secant shear modulus reduction curve for samples of
Turkish clay with different plasticity index (after [41]).

Fig. 19. Secant shear modulus reduction curve for fully saturated clay subjected to
cyclic undrained loading. Fig. 21. Volumetric shear strain threshold values for different PI.
D. Lombardi et al. / Soil Dynamics and Earthquake Engineering 49 (2013) 165180 177

have a more linear cyclic stressstrain response at small strains and 6.3. Choosing diameter of monopile for design
to degrade less at larger strain than soils with lower PI. Therefore
clays with higher PI are characterised by higher values of volumetric The non-dimensional group (P/GD2) suggests that higher the
threshold shear strain. For example, considering the lower bound, gtv diameter of the monopile, the lower is the average strain in the
increases from 0.025 to 0.15 when PI increases from 9 to 40. surrounding soil and therefore lower is its degradation. For the two
Fig. 21 collates the volumetric threshold shear strain value (gtv) prototype turbines considered in this paper (Sheringham Shoal and
for 18 types of clays having plasticity index (PI) ranging from 9 to Kentish Flat, see Table 4), (P/GD2) is in the range of 0.001 to 0.008
100. The experimental data were collected from research published which is indeed in the range of gtv for clayey soils, see Fig. 20. For a
by Kim and Novak [24], [17,26,31,37,41,50,51,56]. In the same particular wind turbine (i.e. known P) to be located at a particular
gure, two linear correlations for evaluating gtv are also suggested site (representative shear modulus of the soil G is also known) the
for the investigated range of PI (9oPIo100). The experimental allowable value of (P/GD2) may be chosen based on volumetric
data used for deriving the plot in Fig. 21 are given in Appendix B. threshold shear strain (gtv) as illustrated earlier. However, if the
The experimental results presented in Figs. 11 and 16 indicate prediction of change in frequency, based on the value of gtv, may
that the dimensionless group (P/GD2) representing the strain level have an adverse effect on the dynamic performance of WTG (Wind
in the soil next to the monopile i.e. the strain in the mobilised turbine Generator), a more conservative approach may assume the
shear zone can be mapped to the volumetric threshold strain, linear shear strain (gtl) as threshold value.
gtv which is a fundamental property of a soil and can be obtained While valuable insights on the mechanisms that affect the long
from soil testing. The reason being, for (P/GD2 0.7%) or less, there term performance can be obtained from a relative simple test,
is less than 5% reduction in natural frequency. For kaolin clay used more work is necessary if detailed guidelines are to be prepared.
in this experimental investigation, the value of the threshold The next section shows the applicability of the above concepts for
volumetric shear strain (gtv) was estimated by Vucetic [52] to be deciding a monopile diameter in clayey soils.
in the range of 0.0800.100%, based on the work carried out by
Ohara and Matsuda [40]. Vucetic [52] states that the value of gtv 6.4. Example
increases with the plasticity index (PI) of the soil but that the
inuence of the overconsolidation ratio may be neglected. An Consider a stiff monopile of length 20 m, required for vertical
attempt has been made to link the value of gtv with the value of bearing capacity to support a 70 m high tower of a 3.5 MW
(P/GD2) for which negligible change of natural frequency was turbine. Based on dynamic considerations, in order to be able to
observed. Figs. 11 and 16 suggest that for (P/GD2) of 0.07%, the generate power over a wide range of wind speed, it is necessary to
change in natural frequency is within 5% (fN-cycles/finitial 0.95), and be condent that any change in natural system frequency is no
from an engineering point of view, the behaviour can be con- greater than 5%. The maximum shear acting in the pile is 2 MN.
sidered practically non-degradable. If an average value of The soil at the site is stiff overconsolidated high plasticity clay
gtv 0.09% is taken for koalin clay, the ratio of threshold strain (PI 74) which has a Shear modulus of 100 MPa.
and (P/GD2) becomes 1.3, see Eq. (10). Using the tting equation given in Fig. 21, and for PI of 74, the
lower bound volumetric threshold strain can be taken as 0.2%.
gtv 0:09
The allowable (P/GD2) is given by
1:3 10
P=GD2 0:07
P gtv 0:2
) 0:0008 12
It is interesting to note that, the average strain in the soil (gav) GD2 2:5 2:5  100
around a laterally loaded pile, which can be expressed by Eq. (11), and the outer diameter required is
suggested by Klar [25], is of the same order as that given by (10). r
2 MN
  D 5:0 m 13
d 100 MPa  0:0008
gav 2:6 11
D Therefore the minimum diameter required is 5 m. This
research suggests that if the diameter is less than 5 m, the strain
where d and D are the displacement and the outer diameter of the developed in the soil next to the pile may lead to progressive
pile, respectively. degradation of the foundation stiffness leading to lower of natural
Eqs. (10 and 11) are similar in the sense that the terms (P/GD2) frequency of the overall system.
and (d/D) represent factored average strain in the soil (but not the
actual strain) in the deformation mechanism (pile-soil interac-
tion) or in the mobilised strain zone. Further details of derivation 7. Conclusion
of these two terms can be found in Appendix A and Bhattacharya
et al. [6]. It is therefore reasonable to make the allowable value The natural frequency and the long-term performance of a
of (P/GD2) a multiple of the volumetric threshold strain gtv. wind turbine model founded on clay soil have been studied using
While (P/GD2) represents factored average shear strain in the soil a number of 1-g tests. It has been shown that small scale
(having shear modulus of G) in the mobilised zone next to the pile experimental studies can be carried out to study complex
(having diameter D) due to a lateral load P, gtv represents the dynamic soilstructure interaction problems.
limiting strain level in the same beyond which progressive Based on the experimental results reported here, the following
degradation in the soil will occur. conclusions may be drawn:
This concept is quite similar to the Mobilisable Strength
Design (MSD) concept pioneered by Bolton [9], Osman and Bolton
[43] where the average strain in a mechanism (deformed zone or (a) The dynamic response of the physical model is very sensitive
sheared zone) is linked with an element test in a soil. In the to the exibility of the foundation. The presence of the
absence of more data and in order to ensure a change in natural foundation provides increased exibility and increased damp-
frequency of the wind turbine within 5%, the ratio expressed in ing of the system.
Eq. (10) can be taken equal to 2.5. This provides a factor of safety (b) The natural frequency of a monopile supported wind turbine
of about 2. founded on clayey soil may change with number of cycles of
178 D. Lombardi et al. / Soil Dynamics and Earthquake Engineering 49 (2013) 165180

repeated loading. For clayey soil, a decrease in natural with the strain distribution in the soil and the stresses in the soil
frequency is expected depending on the strain level in the must be in equilibrium with the reaction force from the pile.
soil next to the pile and the ratio of system frequency to the The pile deection will cause strain in the soil and the strain eld
forcing frequency. in the soil around a pile is very complex, see Fig. A1. A soil element in
(c) Non-dimensionless groups permit scaling the model test the front of the pile will have pure compression and a corresponding
results to a prototype through conceptual understanding soil element in the orthogonal direction will have pure shear.
and knowledge gained from element tests on soils. It has Similarly a soil element behind the pile will have pure extension.
been shown that the dimensionless group (P/GD2) (P being is Any other soil element will have a combination of either shear and
the net shear force in the foundation, D diameter of the compression or shear and extension. However, a spatial average
monopile and G is the representative shear modulus of the shear strain in whole deformable zone (mobilised) may be obtained.
soil) also captures the strain level in the soil due to moment The average shear strain in the soil can be expressed as a function of
loading and the cyclic stress ratio (CSR) in the soil adjacent to pile deection (d) and pile outer diameter (D) given by Eq. (A.1).
the pile. Tests having a high value of (P/GD2) showed a higher
reduction of natural frequency with long-term cyclic loading.
d
es p A:1
On the other hand, for low values of (P/GD2) (e.g. 0.02% or D
lower), the change of natural frequency was negligible. This Klar [25] suggested a value of 2.6 for the coefcient of
non-dimensional group also justies the use of large diameter proportionality between the average strain in the soil and the
monopiles as foundations for modern offshore wind turbines. ratio of head deection and pile diameter. Similar concept is used
(d) Furthermore, (P/GD2) is calibrated against a well-known in API code to relate e50(strain at 50% yield stress) to displace-
element test parameter threshold volumetric strain (gtv) of ments in the pile.
the soil for providing practical design guidelines. The pile deection at a particular depth is a function of the
(e) Finally, practical guidance for choosing the diameter of external load, P, the shear modulus of the soil, G, and the pile
monopile foundations has been proposed. diameter, D. Therefore, the average strain eld in the soil around a
pile can be expressed as a function of three parameters:

es f P,D,G A:2

Appendix A. Derivation of the group CSR (cyclic stress ratio)


and (P/GD2)

8.1. Strain eld in the soil around the laterally loaded pile

Repeated shear strain may reduce the stiffness of saturated


soils. As discussed in the paper, the changes in soil stiffness may
drive the long term performance and as a result, the average
strain next to a pile is a governing criterion that must be
preserved in order to ensure similar stiffness degradation in both
model and prototype. The relevant non-dimensional group can be
derived by considering that the average shear strain eld around
a laterally moving pile.
Fig. A1 schematically shows the plane strain idealisation of
pilesoil interaction at a particular depth of consideration (repre-
sented by section X-X at a depth d). The deection of the pile at a
depth d is represented by (d). The deection must be compatible Fig. A2. Geometry of the characteristic mesh, Randolph and Houslby [45] .

Fig. A1. Schematic diagram showing the average shear strain concept in the soil around the pile .
D. Lombardi et al. / Soil Dynamics and Earthquake Engineering 49 (2013) 165180 179

The parameters in Eq. (A.2) can be used to obtain a dimension- The vertical effective stress can be related to the shear
less group as follows: modulus (G) of the soil:
 
P F F
es f 2
A:3 s0v pG A:7
GD FL2 L2 L2
Eq. (A.3) describes the non-dimensional group that takes into It is usually found that G is proportional to s0n v where the value
account a measure of strain eld in the soil generated by a lateral of n depends on the type of soil. The value of n varies between
loaded pile. Eq. (A.3) shows that the strain in the soil is directly 0.435 to 0.765 for sandy soil [55] but a value of 0.5 is commonly
proportional to the horizontal load applied at the pile head, used. For clayey soil, the value of n is generally taken as 1.
inversely proportional to the soil stiffness and inversely propor- Combining Eqs. (A.4) to (A.7), one can see that the non-
tional to the square of the pile diameter. dimensional group expressed by Eq. (A.3) can also guarantee
similarity of cyclic stress ratio. This leads us to a non-dimensional
group (Eq. (A.8)) that must be satised.
8.2. Cyclic stress ratio (CSR) in the soil in the shear zone
   
P P
In geotechnical earthquake engineering, it is well established A:8
GD2 mod el GD2 prototype
that degradation of a soil due to liquefaction-type failure is a
function of cyclic stress ratio (CSR) which is dened as the ratio of It is interesting to note that using two different approaches
the shear stress to the effective vertical stress at a particular based on average strain in the soil and on the cyclic stress ratio
depth, dened by Eq. (A.4) (see, for example, [49]). The cyclic (CSR) dimensional analysis leads to a unique non-dimensional
stress ratio (CSR) can be expressed by Eq. (A.4): group given by Eq. (A.8).
tcyc
CSR A:4
s0v
where tcyc is the cyclic shear stress imposed by the pile on the soil Appendix B
at a particular depth; s0v is the effective vertical stress on the soil
at the same depth. Fig. 18 in the paper is based on collation of the following data.

Soil name PI OCR Test apparatus Reference

Windsor silty clay 30 2.7 Resonant Column [24]


Wallaceburg silty clay 25 5.1 Resonant Column [24]
Hamilton clayey silt 12 5.8 Resonant Column [24]
Sarnia silty clay 14 1.8 Resonant Column [24]
Alluvianal clay at Teganuma 44, 57, 58, 83, 96 NC Cyclic triaxial Kokusho, 1982
laboratory-made clays 15, 30, 50, 100 NC Not specied [51]
Vallericca clay 31 OC Resonant Column [17]
Pietratta clay 30 OC Resonant Column [17]
Todi clay 28 OC Resonant Column [17]
laboratory-made clay 22 3 Direct simple shear [31]
Fat clay 36 Not specied Not specied [50]
Not specied 10, 20, 40, 60, 80 Not specied Not specied [37];
Turkish clay 9,10,12,15,18,20,25,27,35,40 NC Cyclic triaxial [41]
Ariake clay 44 NC Hollow Cylinder [56]a
Onada clay 50 NC Hollow Cylinder [56]a

Fig. A2 shows the deformation mechanism for a quarter of a pile, References


following Randolph and Houslby [45], where D (the angle that denes
the characteristics) is a function of roughness of the pile. For a rough [1] Achmus M, Kuo YS, Abdel-Rahman K. Behavior of monopile founda-
pile, D p/2 and therefore angle CQE will be 90 degrees. It can be tions under cyclic lateral load. Computers and Geotechnics 2009;36(5):
72535.
shown that the area of the deformed zone (Adef) is approximately [2] Adhikari S, Bhattacharya S. Dynamic analysis of wind turbine towers on
given by Eq. (A.5). exible foundations. Shock and Vibration 2011;19:3756, http://dx.doi.org/
10.3233/SAV-2012-0615 IOS press.
[3] Adhikari S, Bhattacharya S. Vibrations of wind-turbines considering soil-
D2 3 p
Adef  p 8 2 2:644D2 A:5 structure interaction. Wind and Structures 2011;14(2):85112.
16 [4] API. Recommended practice for planning, designing, and constructing xed
offshore platforms: working stress design. 20th edn. Washington, DC:
As the deformed zone is proportional to D2, cyclic shear stress American Petroleum Institute; 1993 RP2A-WSD.
[5] Bhattacharya S, Cox, J, Lombardi, D Muir Wood, D. (2012). Dynamics of
can be expressed by Eq. (A.6). offshore wind turbines on two types of foundations. In: Proceedings of
ICEgeotechnical engineering. Ahead of print.
P F [6] Bhattacharya S, Lombardi D, Muir Wood D. Similitude relationships of
tcyc p A:6
D2 L2 physical modelling of monopile-supported offshore wind turbines. Interna-
tional Journal of Physical Modelling in Geotechnics 2011;11(2):2868.
180 D. Lombardi et al. / Soil Dynamics and Earthquake Engineering 49 (2013) 165180

[7] Bhattacharya S, Adhikari S. Experimental validation of soilstructure inter- [32] LeBlanc, C. (2009). Design of offshore wind turbine support structuresselected
action of offshore wind turbines. Soil Dynamics and Earthquake Engineering topics in the eld of geotechnical engineering. PhD Thesis. Aalborg University.
2011;31:80516. [33] Li Z, Haigh SK, Bolton MD. The response of pile groups under cyclic lateral
[8] Bhattacharya, S, Carrington, TM and Aldridge, TR. (2006) Design of FPSO piles loads. International Journal of Physical Modelling in Geotechnics
under storm loading, OTC conference. Paper Number 17861. 2010:4757.
[9] Bolton, MD. (2012). Performance-based design in geotechnical engineering. [34] Little, RL Briaud, JL. (1988). Full scale cyclic lateral load tests on six single
In: 52nd Rankine lecture, 21 March 2012. piles in sand. Miscellaneous paper GL-88-27. Geotechnical Division, Texas
[10] Camp TR, Morris MJ, Van Rooij R, Van Der Tempel J, Zaaijer MB, Henderson A, A&M University, TX.
et al. Design method for offshore wind turbine at exposed sites. Garrad [35] Lombardi, D.. Dynamics of offshore wind turbines. MSc (by research) thesis,
Hassan and Partners Ltd.; 2003 Report. University of Bristol; 2010.
[11] Carter M. North Hoyle offshore wind farm: design and build. Proceedings of [36] Long J, Vanneste G. Effects of cyclic lateral loads on piles in sand. Journal of
the ICE: Energy 2007;160(1):219. Geotechnical Engineering 1994;120(1):22544.
[12] Chang CS, Whitman RV. Drained permanent deformation of sand due to [37] Massarsch, KR.. Deformation properties of ne-grained soils from seismic
cyclic loading. Journal of Geotechnical Engineering 1988;114(10):116480. tests. In: Keynote Lecture, Intern. Conf. Site Characterization, Porto; Septem-
[13] Cuellar P, Georgi S, Baeler M, Rucker
W. On the quasi-static granular ber 2004. p. 13346.
convective ow and sand densication around pile foundations under cyclic [38] Matlock, H. (1970). Correlations for design of laterally loaded piles in soft
lateral loading. Granular Matter 2012;14(1):1125. clay. In: Proc. 2nd Ann. Technol. Conf., Houston, TX.
[14] DNV. Guidelines for design of wind turbines. 2nd Ed.DNV/Riso; 2002. [39] Muir Wood D. Geotechnical modelling. Spon Press (Taylor and Francis
[15] DNV (2007).Offshore standard: design of offshore wind turbine structures. Group); 2004.
DNV-OS-J101, Det Norske Veritas. [40] Ohara S, Matsuda H. Study on the settlement of saturated clay layer induced
[16] Doyle, EH, Dean, ETR, Sharma, JS, Bolton, MD, Valsangkar, AJ Newlin, JA. by cyclic shear. Soils and Foundations 1988;28(3):10313.
(2004). Centrifuge model tests on anchor piles for tensionleg platforms. Proc. [41] Okur DV, Ansal A. Stiffness degradation of natural ne grained soils during
Ann. Technol. Conf., Houston, TX, Paper 16845. cyclic loading. Soil Dynamic Earthquake Engineering 2007;27:84354.
[17] Georgiannou VN, Rampello S, Silvestri F. Static and dynamic measurements [42] ONeill, MW Murchison, JM. (1983). An evaluation of py relationships in
of undrained stiffness on natural overconsolidated clays. Proceedings of 10th sands. In: Research Rep. No. GT-DF02-83, Department of Civil Engineering,
ECSMFE Florence 1991;1:916. University of Houston, TX.
[18] Hamre, L, Feizi Khankandi, S, Strom, PJ Athanasiu, C.. Lateral behaviour of [43] Osman AS, Bolton MD. Simple plasticity-based prediction of the undrained
settlement of shallow circular foundations on clay. Geotechnique
large diameter monopiles at Sheringham Shoal Wind Farm. In: Gourvenec,
2005;55(6):43547.
White (Eds.), Frontiers in offshore geotechnics II. Taylor & Francis Group,
[44] Poulos, H Hull, T.. The role of analytical geomechanics in foundation
London; 2011. p. 57580.
engineering. In: Foundation engineering: current principles and practices,
[19] Idriss, IM, Dobry, R, Doyle, EH, Singh, RD. Behaviour of soft clays under
vol 2; 1989. p. 1578606. Reston, VA: ASCE.
earthquake loading conditions. In: Proc. 8th Ann technol conf, Houston, TX;
[45] Randolph MF, Houslby GT. The limiting pressure on a circular pile loaded
1976. p. 60516.
laterally in cohesive soil. Geotechnique 1984;34(4):61323.
[20] Idriss IM, Singh RD, Dobry R. Nonlinear behavior of soft clays during cyclic
[46] Randolph MF. Science and empiricism in pile foundation design. In: 43rd
loading. Journal of Geotechnical and Geoenvironmental Engineering
Rankine Lecture. Geotechnique 2003;53(10):84776.
1978;104(12):142747.
[47] Reese, L, Cox, WR Koop, FD.. Analysis of laterally loaded piles in sand. In: Proc.
[21] IEC 61400-1, Wind turbine generator systemsPart 1: Safety requirements,
6th Ann. Technol. Conf., Houston, TX; 1974. Paper no. 2079.
2nd ed; 1999.
[48] Reese, LC, Cox, WR Koop, RD. (1975). Field testing and analysis of laterally
[22] Jardine RJ. Some observations on the kinematic nature of soil. Soils and
loaded piles in stiff clay. In: Proc. 7th Ann. Technol. Conf., Houston, TX.
Foundations 1992;32(2):11124. [49] Seed HB, Idriss IM. Simplied procedure for evaluating soil liquefaction potential.
[23] Jeanjean, P.. Re-Assessment of py curves for soft clays from centrifuge Journal of the Soil Mechanics and Foundation Engineering Division. ASCE
testing and nite element modeling. In: Proc. 6th Ann. technol. conf., 1971;97(SM9):124973.
Houston, TX; 2009. paper: 17861. [50] Stokoe, KH, Darendeli, MB, Andrus, RD Brown, LT. Dynamic soil properties:
[24] Kim TC, Novak M. Dynamic properties of some cohesive soils of Ontario. laboratory, eld and correlations studies. In: Secoe Pinto, editor. Proc. Second
Canadian Geotechnical Journal 1981;18(3):37189. Int. Conf. on earthquake geotechnical rngineering, Lisbon, vol. 3Rotterdam,
[25] Klar A. Upper bound for cylinder movement using elastic elds and its Balkena. 1999. p. 81145.
possible application to pile deformation analysis. International Journal of [51] Vucetic M, Dobry R. Degradation of marine clays under cyclic loading. Journal
Geomechanics 2008;8(2):1627. of Geotechnical Engineering ASCE 1988;117(1):89107.
[26] Kokusho T, Yoshida Y, Esashi Y. Dynamic properties of soft clay for wide [52] Vucetic M. Cyclic threshold shear strains in soils. Journal of Geotechnical
strain range. Soils and Foundations 1982;22(4):118. Engineering 1994;120(12):220828.
[27] Kramer SL, Heavey EJ. Lateral load analysis of nonlinear piles. Journal of Soil [53] Welch PD. The use of fast Fourier transform for the estimation of power
Mechanics & Foundations Division 1988;114(9):10459. spectra: a method based on time averaging over short, modied period-
[28] Kuhn, M. (2000). Dynamics of offshore wind energy converters on mono-pile ograms. IEEE Transactions on Audio and Electroacoustics 1967;15(2):703.
foundation experience from the Lely offshore wind turbine. In: OWEN [54] Wichtmann T, Niemunis A, Triantafyllidis T. Strain accumulation in sand due
Workshop. to cyclic loading: drained cyclic tests with triaxial extensions. Soil Dynamics
[29] Kuhn, M. Offshore wind farms. In: Gash, Twelve (eds.), Wind power plants: and Earthquake Engineering 2007;27(1):428.
fundamentals, design, construction and operation; 2002. p. 36584. [55] Wroth, CP, Randolph, MF, Houlsby, GT Fahey, M. 1979. A review of the
[30] Kuo Y, Achmus M, Abdel-Rahman K. Minimum embedded length of cyclic engineering properties of soils with particular reference to the shear
horizontally loaded monopiles. Journal of Geotechnical and Geoenvironmental modulus, Report CUED/D-SOILS TR75. University of Cambridge.
Engineering 2012;138(3):35763, http://dx.doi.org/10.1061/(ASCE)GT.1943- [56] Yamada S, Hyodo M, Orense RP, Dinesh SV. Initial shear modulus of remolded
5606.0000602. sand-clay mixtures. Journal of Geotechnical and Geoenviromental Engineer-
[31] Lanzo G, Vucetic M, Doroudian M. Reduction of shear modulus at small ing ASCE 2008;134(7):96071.
strains in simple shear. Journal of Geotechnical and Geoenviromental [57] Zaaijer MB. Foundation modelling to assess dynamic behaviour of offshore
Engineering ASCE 1997;123(11):103542. wind turbines. Applied Ocean Research 2006;28:4557.

Das könnte Ihnen auch gefallen