You are on page 1of 12

Engineering and Computational Mechanics Proceedings of the Institution of Civil Engineers

http://dx.doi.org/10.1680/jencm.16.00006
Soil–monopile interactions for offshore Paper 1600006
Received 08/02/2016 Accepted 13/07/2016
wind turbines
Keywords: foundations/granular materials/offshore engineering
Cui and Bhattacharya

ICE Publishing: All rights reserved

Soil–monopile interactions for


offshore wind turbines
Liang Cui BE, PhD Subhamoy Bhattacharya PhD
Lecturer, Department of Civil and Environmental Engineering, University of Professor, Department of Civil and Environmental Engineering, University
Surrey, Guildford, UK (corresponding author: l.cui@surrey.ac.uk) of Surrey, Guildford, UK

Many offshore wind turbines are supported by large-diameter piles (known as monopiles) and are subjected to a
large number of cyclic and dynamic loads. There are evidences suggesting that foundation stiffness are changing
with cycles of loading and this may lead to changes in the natural frequency of the system with the potential for
unplanned system resonances. There are other consequences such as excessive tilt leading to expensive repair or
even complete shutdown. Therefore, it is vital to understand the long-term response of wind turbine foundation so
that a method to predict the change in frequency and long-term tilt could be established. This paper aims to present
the experimental work of small-scale physical modelling and discrete element modelling of the interactions between
a monopile and the surrounding soil. Changes in soil stiffness under cyclic loading of various strain amplitudes were
examined for both physical modelling and discrete element modelling. Micromechanics of soils underlying the soil
stiffness change was investigated using discrete element method. Variation of force distribution along the monopile
under cyclic loading was analysed to show the influence of monopile stability.

Notation for designing monopiles for OWTs (e.g. the approach suggested
A area of hysteresis loop by DNV, 2007) are based on the methods originally developed
AΔ area of the triangle representing elastic energy for the offshore oil and gas industry (API, 1993). Lombardi
D pile diameter et al. (2013) explain the obvious differences between offshore
G shear modulus of soil platform piles and monopoles, which are summarised below:
P lateral load
α hysteresis damping ratio Piles for offshore structures are typically 60–110 m long and
1·8–2·7 m diameter; monopiles for OWTs are commonly
30–40 m long and 3·5–6 m diameter. Degradation in the upper
1. Introduction
soil layers resulting from cyclic loading is less severe for off-
1.1 Background shore piles which are significantly restrained from pile head
Offshore wind turbines (OWTs) are relatively new structures rotation, whereas monopiles are free headed and more vulner-
and are providing increasing proportion of energy. This is due able to tilt. In any case, the non-linear behaviour of the struc-
to the fact that offshore sites are characterised by stronger and ture could be taken into account using an approach similar to
more stable wind conditions than the corresponding land sites that suggested by Comodromos et al. (2009) and Conte et al.
and thus have a higher capacity factor when compared with (2013, 2015). A design method using a beam on non-linear
equivalent onshore turbines. Winkler springs (‘p − y’ method in API (1993) code or DNV
(2002, 2007) code) may be used to obtain pile head deflection
The design and construction of foundations for offshore tur- under cyclic loading, but its use is limited for wind turbines
bines are challenging due to the harsh environmental con- due to the following.
ditions. Different types of foundations have been proposed:
monopile, gravity base, jacket, suction caisson and floating & The widely used API (1993) model is calibrated
anchoring systems. However, most of the offshore turbines cur- against response to a small number of cycles
rently in operation (UK Round 1 development) are supported (maximum 200 cycles) for offshore fixed platform
on driven monopiles with diameters ranging between 3·5 and applications – for example Matlock (1970), Reese et al.
6 m. The choice of monopiles results from their simplicity of (1974), Reese et al. (1975). In contrast, for a real OWT,
installation and the proven success of driven piles in support- 107–108 cycles of loading are expected over a lifetime of
ing offshore oil and gas infrastructures. The available methods 20–25 years.

Downloaded by [ UNIVERSITY OF CAMBRIDGE] on [23/09/16]. Copyright © ICE Publishing, all rights reserved.
Engineering and Computational Mechanics Soil–monopile interactions for offshore
wind turbines
Cui and Bhattacharya

& Under cyclic loading, the API (1993) or DNV (2002, 2007) & The mass and aerodynamic imbalances of the rotor
model always predicts degradation of foundation stiffness (blades) generate vibration at the hub level and apply
in sandy soil. However, recent work by Bhattacharya and lateral load. This load has a frequency equal to the
Adhikari (2011), Cuéllar et al. (2012), LeBlanc (2009) rotational frequency of the rotor (referred to as 1P loading
suggested that the foundation stiffness for a monopile in in the literature). Since most of the industrial wind turbines
sandy soil will actually increase as a result of densification are variable speed machines, 1P is not a single frequency
of the soil next to the pile. but a frequency band between the frequencies associated
& The ratio of horizontal load to vertical load is very high with the lowest and the highest revolutions per minute
in OWTs when compared with fixed jacket structures. (rpm).
Therefore, monopiles can be considered as moment & The blade shadowing effects (referred to as 2P/3P in the
resisting foundations. literature) also applied loads in the tower. This is a
dynamic load having a frequency equal to three times the
rotational frequency of the turbine (3P) for three-bladed
1.2 Complexity of loads on the monopile wind turbines and two times (2P) the rotational frequency
The loads acting on the wind turbine tower are ultimately of the turbine for two-bladed turbines. The 2P/3P loading
transferred to the foundation and can be classified into two is also a frequency band like 1P and is simply obtained by
types: static or dead load due to the self-weight of the com- multiplying the limits of the 1P band by the number of
ponents and the cyclic/dynamic loads arising from the wind, turbine blades.
wave, 1P and 3P loads (for further details see Arany et al.,
2014). However, the challenging part is the dynamic loads A calculation procedure is developed by Arany et al. (2014)
acting on the wind turbine and the salient points are discussed which can be easily carried out in a spreadsheet program. The
below. output of such a calculation will be relative wind and the wave
loads and an example is shown in Figure 1, where it is
& The rotating blades apply a cyclic/dynamic lateral load at assumed that the wind and the wave are perfectly aligned
the hub level (top of the tower) and are a function of the which is a fair assumption for deeper water further offshore
turbulence of the wind. The magnitude of the dynamic projects (i.e. fetch distance is high).
component depends on the turbulent wind speed
component. From the Kaimal spectrum as suggested in the DNV (2002,
& The waves crashing against the substructure apply a lateral 2007) code, the peak frequency of wind turbulence can be
load very close to the foundation. The magnitude of this obtained theoretically. In the absence of site-specific data,
load depends on the wave height and wave period, as well and for the purpose of foundation design, the time period
as the water depth. for wind can be conservatively assumed to be acting at the

Simplified cyclic load scenario Peak


frequency/period
of wave spectrum
~10 s
Peak frequency/period of turbulence spectrum
Mudline bending moment

~100 s
Mmax

Mmean

Mmin

0 20 40 60 80 100 120 140 160 180 200


Time: s
Wind + wave load Wind load

Figure 1. Mudline moment (wind and wave load) acting on the


wind turbine tower

Downloaded by [ UNIVERSITY OF CAMBRIDGE] on [23/09/16]. Copyright © ICE Publishing, all rights reserved.
Engineering and Computational Mechanics Soil–monopile interactions for offshore
wind turbines
Cui and Bhattacharya

hub level having a time period at about 100 s as shown in as each blade passes through the shadow of the tower.
Figure 1. Figure 2 shows the blade passing frequency for the
3·6 MW wind turbine generator.
Therefore, one of the objectives of the paper is to study theor-
etically the soil–structure interaction on a monopile foundation From Figure 2 it may be observed that in order to avoid the
for an asymmetrical loading as shown in Figure 1. resonance of the system, the designed frequency of the overall
system must be kept away from the frequency content of
applied loads. Specifically, DNV (2002) suggests that the
1.3 Dynamic issues in OWT design natural frequency of the wind turbine should be at least ±10%
OWTs are characterised by a unique set of dynamic loading away from the 1P and 2P/3P frequencies. Bearing these con-
conditions and are summarised as follows. siderations in mind, there are three possible slots where the
natural frequency of the system may lie. They correspond to
& Environmental dynamic loads arising from the wind and three different design approaches namely: soft–soft (natural
waves. Figure 2 shows the plot of power spectral density frequency <1P), soft–stiff (natural frequency between 1P and
of wind and wave loading around the UK coastline 2P or 3P) and stiff–stiff (natural frequency >2P or 3P).
(particularly in the North Sea). The predominant wave The most common design, used, for example, in the Round 1
frequency is 0·1 Hz, which corresponds to 10 s wave UK development, is ‘soft–stiff’, which implies that the natural
period. frequency lies between 1P and 3P. Few points may be noted as
& Rotor loading at a frequency which is commonly referred follows.
to as 1P. Figure 2 shows the rotor frequency for a 3·6 MW
wind turbine having an operational range between 5 and & The natural frequency of three-bladed wind turbines in
13 rpm – that is, 0·08–0·22 Hz. In the power spectral soft–stiff design is very close to the forcing frequency due
density plot, the 1P frequency appears as a band. to 1P and 3P.
& The blade passing frequency (3P or 2P for a three-bladed & Any change in natural frequency either an increase or a
or two-bladed turbine, respectively) is a forced loading decrease will have adverse consequences on fatigue life of
generated from the effect of wind deficiency that occurs the wind turbine system.

Typical frequency diagram


1P ±10% safety region
3P ±10% safety region
Kaimal
spectrum
(wind)

Rotor
frequency 1P Blade passing frequency – 3P
Power spectral density

JONSWAP
spectrum
(wave)

0 0·1 0·2 0·3 0·4 0·5 0·6 0·7 0·8


Frequency: Hz

Soft–soft Soft–stiff Stiff–stiff

Figure 2. Forcing simplified power spectral density of the forcing


frequencies applied to typical three-bladed 3·6 MW OWT with
an operational interval in the range of 0·08–0·22 Hz (5–13 rpm)

Downloaded by [ UNIVERSITY OF CAMBRIDGE] on [23/09/16]. Copyright © ICE Publishing, all rights reserved.
Engineering and Computational Mechanics Soil–monopile interactions for offshore
wind turbines
Cui and Bhattacharya

& It is well known that under the action of cyclic/dynamic created a database of change of frequency and damping of the
loads most soils change its properties. Depending on the wind turbine system for different values of
frequency and amplitude of loading, the soil may stiffen or
soften. Stiffness of a monopile to large extent is dependent & strain field in the soil next to the pile. This is expressed in
on the stiffness of the soil and therefore the impact of a Figure 3 by a non-dimensional group (P/GD 2), where P is
large number of cycles on the monopiles needs to be the lateral load, G is the shear modulus of the soil and D is
evaluated. the pile diameter. Further details on the scaling relations
can be found in Bhattacharya et al. (2011) and Lombardi
The next section of the paper collates the research carried out et al. (2013)
on dynamic soil–structure interaction along with field evidences. & forcing frequency imposed by the different dynamic loads
& number of cycles of loading.

Figure 3 shows one such graph obtained from scaled model


2. Dynamic soil–structure interaction
test studies by Bhattacharya et al. (2013a) and Yu et al. (2015)
in OWTs
on monopiles where the observed change in natural frequency
2.1 Summary based on scaled model tests is plotted with the number of cycles for various levels of strains
Extensive research has been carried out by Bhattacharya et al. in the soil next to the pile. The main observations from the
(2013a, 2013b), Lombardi et al. (2013) and Yu et al. (2015) to tests are as follows.
study the effects of cyclic loads on the first natural frequency
of wind turbines. Tests were carried out on different types of & For strain-hardening sites (e.g. loose to medium dense
foundations: monopiles, jackets, multiple pods. A typical test sand) where the stiffness of the soil increases with cycles of
consists of the application of cyclic loading for a particular loading, the natural frequency of the overall system will
time interval (or for a certain number of cycles) and then increase possibly due to densification.
measuring the frequency and damping of the system by a free & For strain-softening sites (clay sites) where the stiffness of
vibration test. The cyclic loading was applied through an the soil may decrease with cycles of loading, the natural
actuator. However, during the free vibration test (also known frequency of the overall system will also decrease
as a ‘snap back’ test in the literature), the actuator was discon- correspondingly. Of course, this depends on the strain level
nected from the tower and the tower was given a small ampli- in the soil next to the pile and the number of cycles.
tude vibration and the acceleration of the system was recorded.
The cyclic lateral loading was applied at different frequencies Lombardi et al. (2013) established a link between the strain of
and for different lateral load magnitudes. This set of tests the soil next to the pile as observed in the mechanism (P/GD 2)

10

0
% change in natural frequency

0 10 000 20 000 30 000 40 000 50 000 60 000 70 000 80 000


–5
Saturated sand (strain 5·0 × 10–4)
–10
Dry sand (strain 5·0 × 10–4)
–15 Soft clay (strain 2·0 × 10–4)
–20 Soft clay (strain 3·4 × 10–3)
Dry sand (strain 1·945 × 10–4)
–25
Dry sand (strain 1·876 × 10–4)
–30 Dry sand (strain 1·798 × 10–4)
–35 Dry sand (strain 0·243 × 10–4)

–40
Number of cycles

Figure 3. The observed change in natural frequency with the


number of cycles for different strain levels in the soil around
the pile

Downloaded by [ UNIVERSITY OF CAMBRIDGE] on [23/09/16]. Copyright © ICE Publishing, all rights reserved.
Engineering and Computational Mechanics Soil–monopile interactions for offshore
wind turbines
Cui and Bhattacharya

with ‘threshold strains’ from element testing. This allows (Cui, 2006; Cui et al., 2007; O’Sullivan et al., 2008) was used
designers to choose pile diameter to avoid stiffness degradation. to perform the presented study.

2.2 Field evidences of dynamic soil–structure To simulate the soil stiffness changes, a DEM model of a soil
interaction tank (100 mm  100 mm  50 mm) was first created. The soil
There are some field evidences of dynamic soil–structure inter- tank was filled by about 13 000 spherical particles with radii in
action and they are briefly discussed below. the range of 1·1–2·2 mm. The particles were deposited under
gravity. The pile is 20 mm in diameter and was embedded to a
& Natural frequency was measured for a wind turbine depth of 40 mm by removing particles located in the space
structure in the Hornsea wind farm where it was observed which was to be occupied by the pile. Particles were allowed to
that the natural frequency decreased from 1·23 to 1·13 Hz settle down again following the installation of the pile. The
after 3 months of operation (Lowe, 2012). current work aims to obtain qualitative characteristics of soil
& It has been observed that there is an increasing scatter in behaviours for investigations of micromechanics, not to repro-
the observed natural frequency at higher wind speeds. duce quantitatively the model tests. Therefore, this study used
Higher wind speed will lead to higher strain in the soil next large particle sizes, which may cause size effect. Once the soil
to the pile and therefore higher P/GD 2 leading to higher particles were settled down in the soil tank, cyclically horizon-
soil–structure interaction. tal movements were assigned to the pile to simulate the cyclic
& Kuhn (2000, 2002) reported that the target design movements of OWT monopile due to the cyclic loadings.
frequency for Lely wind farm of 0·4 increased to 0·63 Hz Translational movements rather than rotational movements
after 6 years of service. were assigned to the pile at this stage. Rotational movement
will be simulated in the future study to differentiate the effects.
While the field studies provides anecdotal evidence of the Three different strain amplitudes, 0·1, 0·01 and 0·001%, were
change in system frequency and small-scale tests provides the chosen to examine the effects of strain levels. For each strain
overall understanding, discrete element modelling of the prob- amplitude, two types of cyclic loading were applied: symmetric
lem is carried out to better understand the monopile–soil inter- cyclic loading with stains in the range of (−0·1, 0·1%), (−0·01,
action. The next section of the paper describes the modelling 0·01%) and (−0·001, 0·001%) and asymmetric cyclic loading
technique and the results obtained from the simulation. with stains in the range of (0, 0·2%), (0, 0·02%) and
(0, 0·002%). As constrained by the computational costs,
500 cycles were simulated for strain amplitude of 0·1% and
3. Numerical analyses using discrete
1000 cycles were simulated for other strain amplitudes. The
element method modelling
current simulation for 0·1% strain amplitude required about
3.1 Description of discrete element method model 1 month in computational time. The simulation parameters are
Numerical simulations were performed to investigate the listed in Table 1.
underlying mechanism for soil stiffness changes surrounding
the wind turbine monopile. The discrete element method
(DEM) was found to be more appropriate than other numeri- 3.2 Stress–strain response and damping ratio
cal methods (e.g. finite-element method) as it allows direct The resultant horizontal stress applied on the pile against the
monitoring of change in soil stiffness, and more importantly it horizontal strain of soil is illustrated in Figure 4. As shown
offers a method to analyse the micromechanics, which in Figure 4, the stress–strain curves form hysteresis loops,
underlies the stiffness changes. Originally proposed by Cundall
and Strack (1979), DEM simulates granular materials as
assemblies of individual particles which respond to given load Parameters Value
conditions. The interactions between particles are simulated by Soil particle density ρs: kg/m3 2650
contact laws, where the normal and tangential contact forces Particle sizes: mm 1·1, 1·376, 1·651,
are dependent on the overlap and relative displacement 1·926, 2·2
between two contact particles. In this study, the elastic Hertz– Inter-particle frictional coefficient, μ 0·3
Mindlin contact model (Mindlin and Deresiewicz, 1953) is Particle-boundary frictional coefficient, μ 0·1
adopted. The contact forces, accelerations, velocities and dis- Gs (Hertz–Mindlin contact model): Pa 2·868  107
placements of all particles are updated in each small time step Poisson’s ratio 0·22
using the central difference time integration method. Stresses Initial void ratio, e 0·539
and strains are then calculated from the contact forces within a
representative volume element or along a boundary. An open- Table 1. Input parameters for DEM simulation
source DEM code modified and validated in previous studies

Downloaded by [ UNIVERSITY OF CAMBRIDGE] on [23/09/16]. Copyright © ICE Publishing, all rights reserved.
Engineering and Computational Mechanics Soil–monopile interactions for offshore
wind turbines
Cui and Bhattacharya

600 800
First cycle First cycle
Tenth cycle 600 Tenth cycle
400 100th cycle 100th cycle
500th cycle 500th cycle
400
Horizontal stress: Pa

Horizontal stress: Pa
200
200

0 0

−200
−200
−400
−400
−600

−600 −800
−0·10 −0·05 0 0·05 0·10 0 0·05 0·10 0·15 0·20
Horizontal strain: % Horizontal strain: %
(a) (b)

150 150
First cycle First cycle
Tenth cycle Tenth cycle
100 100th cycle 100 100th cycle
1000th cycle 1000th cycle
Horizontal stress: Pa

Horizontal stress: Pa

50 50

0 0

−50 −50

−100 −100

−150 −150
−0·010 −0·005 0 0·005 0·010 0 0·005 0·010 0·015 0·020
Horizontal strain: % Horizontal strain: %
(c) (d)

15 25
Firstcycle First cycle
Tenth cycle 20 Tenth cycle
10 100th cycle 100th cycle
1000th cycle 1000th cycle
15
5
Horizontal stress: Pa

Horizontal stress: Pa

10
0
5
−5
0
−10
−5

−15 −10

−20 −15
−1·0 −0·5 0 0·5 1·0 0 0·5 1·0 1·5 2·0
Horizontal strain: % x 10−3 Horizontal strain: % x 10−3
(e) (f)

Figure 4. Hysteresis loops formed by the stress–strain curves


during cyclic loadings: (a) strain (−0·1, 0·1%); (b) strain (0, 0·2%);
(c) strain (−0·01, 0·01%); (d) strain (0, 0·02%); (e) strain (−0·001,
0·001%); (f) strain (0, 0·002%)

Downloaded by [ UNIVERSITY OF CAMBRIDGE] on [23/09/16]. Copyright © ICE Publishing, all rights reserved.
Engineering and Computational Mechanics Soil–monopile interactions for offshore
wind turbines
Cui and Bhattacharya

indicating the energy dissipations during cyclic loading. It is damping ratio for the strain amplitude of 0·01% oscillated
also observed that the areas of the hysteresis loops increase about a constant value. However, it is obvious that the
with strain amplitude, indicating greater energy dissipation. damping ratio for the strain amplitude of 0·1% decreased dra-
The hysteresis damping ratio, α, can be determined by the matically in the first 30 cycles and then oscillated at an
expression (Karg, 2007) approximately constant value. The damping ratios for asym-
metric cyclic loading are much lower than those for the corre-
A sponding symmetric cyclic loading, due to the higher elastic
1: α¼
4πAΔ energy stored in each cycle.

where A is the area of the hysteresis loop, representing the It is also interesting to observe that although the stresses in
energy dissipated and AΔ is the area of the triangle as indicated the first half cycle for the asymmetric cyclic loading is positive,
in Figures 4(a) and 4(b), representing the elastic energy stored it reduced to negative when the strain goes to zero. Following
in the soil during one load cycle. The variation of the damping a few cycles, the minimum negative stress approaches the
ratio during cyclic loading is illustrated in Figure 5. The same magnitude as the maximum positive stress. Moreover,
the magnitude of stresses and the shape of the hysteresis loops
for both symmetric cyclic loading and asymmetric cyclic
0·40 loading are almost identical after many cycles. The system
Strain (–0·1, 0·1%)
0·35 Strain (0, 0·2%) under asymmetric cyclic loading behaves the same as the sym-
Hysteresis damping ratio, α

Strain (–0·01, 0·01%)


metric cyclic loading with the same strain amplitude after
0·30
many cycles, indicating that the strain amplitude, rather than
Strain (0, 0·02%)
0·25 the maximum strain, dominates the long-term cyclic
0·20 behaviour.

0·15
3.3 Evolution of stiffness
0·10
The secant Young’s modulus of soil in each cycle was calcu-
0·05 lated by determining the slope of a line connecting the
0 maximum and minimum points of each full loop. It is evident
0 200 400 600 800 1000 from Figure 4 that the secant Young’s modulus of soil
Number of cycles increased during cyclic loading. A clearer evolution of secant
Young’s modulus is shown in Figure 6. At a strain amplitude
Figure 5. Hysteresis damping ratio at the end of each cycle of 0·1%, Young’s modulus increases dramatically from 250 to
around 600 kPa for symmetric cyclic loading and from 130 to

1600 1600
Secant Young’s modulus: kPa

Secant Young’s modulus: kPa

1400 1400
1200 1200
1000 1000
800 800
600 600
400 Strain (–0·1, 0·1%) 400 Strain (0, 0·2%)
Strain (–0·01, 0·01%) Strain (0, 0·02%)
200 200
Strain (–0·001, 0·001%) Strain (0, 0·002%)
0 0
0 250 500 750 1000 0 250 500 750 1000
Number of cycles Number of cycles
(a) (b)

Figure 6. Secant Young’s modulus of soil at the end of each


cycle: (a) symmetric cyclic loading and (b) asymmetric cyclic
loading

Downloaded by [ UNIVERSITY OF CAMBRIDGE] on [23/09/16]. Copyright © ICE Publishing, all rights reserved.
Engineering and Computational Mechanics Soil–monopile interactions for offshore
wind turbines
Cui and Bhattacharya

around 600 kPa for asymmetric cyclic loading. At a strain increase in soil stiffness in the first 50 cycles. The soil particle
amplitude of 0·01%, the Young’s modulus increases quickly in displacements for a strain amplitude of 0·01% are less remark-
the first few cycles and then only increases slightly to about able, corresponding to less reduction in soil stiffness.
1100 kPa. At a strain amplitude of 0·001%, Young’s modulus
only increases initially then mobilises at a constant value at It is also evident that only soil particles in an inverted triangu-
1500 kPa. The initial stiffness of asymmetric cyclic loading is lar region enclosing the monopile have noticeable convective
lower than that of symmetric cyclic loading with the same displacements. This is because that the vertical effective stress
strain amplitude due to a higher maximum strain applied. in a soil body increases linearly against depth, which imposes
However, following a few cycles, the stiffness for both types of
cyclic loading approaches the same values, confirming the
50
observations from the stress–strain responses.
40

50 mm in height
The Young’s modulus against horizontal strain in the first half
cycle and in the 500th cycle for a strain amplitude of 0·1% are 30
shown in Figure 7. The stiffness–strain curve in the first cycle
20
displays the similar ‘S’ shape as expected for the shear
modulus–shear strain curve. Following cyclic loadings, stiffness 10
at different strain levels all increases significantly. At a strain
amplitude of 0·01 and 0·001%, stiffness also increases but at a 0
0 20 40 60 80 100
much smaller scale as indicated in Figure 6.
Soil tank − 100 mm in width
(a)
3.4 Convective granular flow
In the model test, ground settlements around the monopile 50
could be observed. To illustrate the ground settlement, plots
40
50 mm in height

of incremental soil particle displacements in the symmetric


cyclic loading with a strain amplitude of 0·1% in the first 30
50 cycles and in the next 50 cycles are given in Figure 8. The
soil particle displacements in the symmetric cyclic loading with 20
a strain amplitude of 0·01% were also illustrated for compari-
10
son. Each arrow in the plot starts from the original centre of a
particle and ends at the new centre at the end of a given cycle. 0
It is evident that the soil particles surrounding the pile moved 0 20 40 60 80 100
downwards, causing ground settlement. Soil densification Soil tank − 100 mm in width
around the pile is the main reason causing the increase of soil (b)
stiffness. It is also clear that the soil particle displacements are 50
only significant in the first 50 cycles, underlying the significant
40
50 mm in height

20 000 30
Secant Young’s modulus: kPa

First cycle
500th cycle strain (–0·1, 0·1%)
20
15 000 500th cycle strain (0, 0·2%)
10
10 000
0
0 20 40 60 80 100
5000 Soil tank − 100 mm in width
(c)
0
1 × 10–8 0·000001 0·0001 0·01 1
Figure 8. Incremental soil displacements at the end of the given
Horizontal strain
cycle (unit: mm): (a) strain (−0·1, 0·1%), 1st–50th cycles; (b) strain
(−0·1, 0·1%), 51th–100th cycles; (c) strain (−0·01, 0·01%),
Figure 7. Secant Young’s modulus against horizontal strain 1st–1000th cycles

Downloaded by [ UNIVERSITY OF CAMBRIDGE] on [23/09/16]. Copyright © ICE Publishing, all rights reserved.
Engineering and Computational Mechanics Soil–monopile interactions for offshore
wind turbines
Cui and Bhattacharya

an increasing constraint on particle movement against depth. 650


First cycle
Cuéllar et al. (2012) performed physical tests of cyclic rotation 600 Tenth cycle
of a pile and observed the granular convective flow and soil 100th cycle
densification around the pile. The shape of the convective soil 550 500th cycle

volume observed by them is similar to the region where the 500


soil particle displacements were concentrated in the current

Stress: Pa
450
DEM simulations. In the current DEM simulations, samples
with same initial void ratio of 0·539 (medium dense sample) 400
all showed densification behaviour and stiffness increase under 350
cyclic loading. It would be interesting to investigate soil beha-
300
viours and stiffness evolutions for a wide range of initial void
ratios in the future study. 250

200
−0·10 −0·05 0 0·05 0·10
3.5 Contact stresses and forces Strain: %
(a)
3.5.1 Average radial stresses
The evolution of average radial stress on the pile at the end of 650
First cycle
each cycle is depicted in Figure 9. Due to soil densification, Tenth cycle
600
the radial stress increased significantly for a strain amplitude 100th cycle
500th cycle
of 0·1% under cyclic loading, reached a peak value and then 550
mobilised at about a constant value. Increased radial stress
Radial stress: Pa

500
would increase side friction and would improve shaft resist-
ance. The increase in radial stress is less remarkable for smaller 450
strain amplitudes due to smaller particle displacements. The
400
evolutions of average radial stress in representative cycles are
illustrated in Figure 10. The shapes of the radial stress–strain 350
curves are quite different for different strain amplitudes.
300
For symmetric cyclic loading with 0·01% strain amplitude
(Figure 10(c)), the radial stress increases at positive strain 250
0 0·05 0·10 0·15 0·20
values, but decreases slightly at negative strain values.
Strain: %
However, for 0·1% strain amplitude (Figure 10(a)), the radial
(b)
stress increases at both positive and negative strains, forming a
‘butterfly’ shaped curve. It is also interesting to observe that, 295
for asymmetric cyclic loading (Figure 10(b)), the stress–strain First cycle
290 Tenth cycle
100th cycle
285 1000th cycle
280
Radial stress: Pa

700
Average radial stress on pile: kPa

275
600
270
500 265
400 260
300 255

200 Strain (–0·1, 0·1%) Strain (0, 0·2%) 250


Strain (–0·01, 0·01%) Strain (0, 0·02%) 245
100 −0·010 −0·005 0 0·005 0·010
Strain (–0·001, 0·001%) Strain (0, 0·002%)
0 Horizontal strain: %
0 250 500 750 1000 (c)
Number of cycles

Figure 10. Evolution of average radial stresses in representative


Figure 9. Evolution of average radial stress on the pile at the end cycles: (a) strain (−0·1, 0·1%); (b) strain (0, 0·2%); (c) strain
of each cycle (−0·01, 0·01%)

Downloaded by [ UNIVERSITY OF CAMBRIDGE] on [23/09/16]. Copyright © ICE Publishing, all rights reserved.
Engineering and Computational Mechanics Soil–monopile interactions for offshore
wind turbines
Cui and Bhattacharya

response in the first few cycles is different from that in the cor- the pile at the end of each cycle is oriented to the left.
related symmetric cyclic loading, However, it eventually However, for the symmetric cyclic loading, the pile compressed
evolves to the same ‘butterfly’ shape after many cycles. It con- the soils on both sides to the same strain level. At the end of a
firms again that the influence of different maximum strains of full cycle, the residual horizontal force is oriented to the right.
cyclic loading can be eliminated after many cycles and the By referring to Figures 4(a) and 4(b), it can be understood
dominant factor for cyclic soil response is the strain amplitude. clearly that the end of a symmetric loading cycle is in the
A comparison between the patterns of convective particle middle of a hysteresis loop, but the end of an asymmetric
flow was made; however, no noticeable difference can be ident- loading cycle is at the left corner of the hysteresis loop, even
ified between the particle flow patterns. In other words, par- though the stress magnitudes for both hysteresis loops are
ticle flow is not the main reason contributing to the change approximately the same. Therefore, although the asymmetric
in the shape of the radial stress–strain curves for symmetric cyclic loading does not cause different dynamic soil responses
cyclic loading. Further study will seek the underlying (stress–strain curves) from the symmetric cyclic loading during
mechanism from the perspective of microstructure/fabric of the cyclic loading, it does build up greater unbalanced hori-
particles. zontal force, which will cause excessive tilt and undermine the
stability of the monopile more significantly in the long term.
Asymmetric response of radial stresses may suggest unba- In this study, monopile was driven to move by a predefined
lanced radial stresses at both sides of the pile. Further investi- constant velocity. It is more realistic to simulate the free
gation on the unbalanced horizontal forces is provided as motions of monopile under the action of resultant external for-
follows. ce/moment, including the interaction force/pressure between
the monopile and the soil. It will be the next objective of the
3.5.2 Unbalanced horizontal force authors’ future work.
As shown in the hysteresis loops formed by the cyclic stress–
strain curves in Figure 4, at the end of each cycle, the stress
4. Conclusions
does not return to zero but builds up gradually. The evolution
The DEM simulations and small-scale tests provide a good
of unbalanced horizontal force at the end of each cycle is illus-
understanding on the soil–structure interaction of foundations
trated in Figure 11. It is evident that the unbalanced horizon-
of OWTs. Various features observed in model tests could also
tal force is more significant with increasing strain amplitude. It
be replicated in DEM studies and thus provides confidence in
can also be seen that the unbalanced force for asymmetric
small-scale physical model tests. The following conclusions
cyclic loading is much larger in magnitude and is oriented in
could be drawn from the study.
the opposite direction compared with that for symmetric cyclic
loading. It is because that for asymmetric cyclic loading, the
& Stiffness of granular soils increases under cyclic loading.
pile only moves to the right side and compresses the soils on
Therefore, the stiffness of monopiles founded on
the right side significantly; therefore, the horizontal force on
granular material is expected to increase with cycles
of loading and this increase may cause a change in
natural frequency of the wind turbine system. It may be
0·3 concluded that a ‘soft–stiff’ will move towards the 3P
0·2 frequency. It is necessary for the designers to predict the
Unbalanced horizontal force: N

change in frequency which is essential to predict the


0·1
fatigue life.
0 & The convective soil flow and the soil densification
0 250 500 750 1000
–0·1 surrounding the monopile is the main reason underlying
Strain (–0·1, 0·1%)
–0·2 the increase in stiffness. These phenomena are more
Strain (0, 0·2%)
significant with the increase in strain amplitude.
–0·3 Strain (–0·01, 0·01%)
& Due to soil densification, the average radial stress on the
Strain (0, 0·02%)
–0·4 pile, and thus the shaft resistance of the pile increases
Strain (–0·001, 0·001%)
–0·5 Strain (0, 0·002%) under cyclic loading.
& Asymmetric cyclic loading applies larger maximum strain
–0·6
Number of cycles to soil, which results in higher stress in the first cycle;
however, the difference in the magnitude of stresses and the
Figure 11. Evolution of unbalanced horizontal force on the pile shape of the stress–strain curves between asymmetric cyclic
at the end of each cycle loading and symmetric cyclic loading eliminates quickly
under cyclic loading. The long-term governing factor for

10

Downloaded by [ UNIVERSITY OF CAMBRIDGE] on [23/09/16]. Copyright © ICE Publishing, all rights reserved.
Engineering and Computational Mechanics Soil–monopile interactions for offshore
wind turbines
Cui and Bhattacharya

the dynamic stress–strain response is the cyclic strain Cuéllar P, Georgi S, Baeßler M and Rücker W (2012) On the
amplitude. quasi-static granular convective flow and sand densification
& The asymmetric cyclic loading builds up greater around pile foundations under cyclic lateral loading.
unbalanced horizontal force, which will undermine Granular Matter 14(1): 11–25.
the stability of the monopile more significantly in the Cui L (2006) Developing a Virtual Test Environment
long term. for Granular Materials Using Discrete Element
Modelling. PhD thesis, University College Dublin,
Acknowledgements Dublin, Ireland.
The authors acknowledge the RWE Innogy and the University Cui L, O’Sullivan C and O’Neill S (2007) An analysis of the
of Surrey for collectively funding this project. triaxial apparatus using a mixed boundary three-
dimensional discrete element model. Geotechnique
57(10): 831–844, http://dx.doi.org/10.1680/geot.2007.57.
REFERENCES 10.831.
API (American Petroleum Institute) (1993) Recommended Cundall P and Strack O (1979) A discrete numerical model for
Practice for Planning, Designing, and Constructing Fixed granular assemblies. Géotechnique 29(1): 47–65,
Offshore Platforms: Working Stress Design, RP2A-WSD, http://dx.doi.org/10.1680/geot.1979.29.1.47.
20th edn. American Petroleum Institute, Washington, DC, DNV (Det Norske Veritas) (2002) Guidelines for Design of Wind
USA. Turbines, 2nd edn. DNV/Riso, Hovik, Denmark.
Arany L, Bhattacharya S, Macdonald J and Hogan SJ (2014) DNV (2007) Offshore Standard: Design of Offshore Wind
Simplified critical mudline bending moment spectra of Turbine Structures, DNV-OS-J101. Det Norske Veritas,
offshore wind turbine support structures. Wind Hovik, Denmark.
Energy 18(12): 2171–2197. Karg C (2007) Modelling of Strain Accumulation Due to Low
Bhattacharya S and Adhikari S (2011) Experimental Level Vibrations in Granular Soils. PhD thesis, Ghent
validation of soil–structure interaction of offshore University, Ghent, Belgium.
wind turbines. Soil Dynamics and Earthquake Engineering Kuhn M (2000) Dynamics of offshore wind energy converters
31(5–6): 805–816. on mono-pile foundation experience from the Lely offshore
Bhattacharya S, Lombardi D and Muir Wood D (2011) wind turbine. OWEN Workshop, CLRC Rutherford
Similitude relationships for physical modelling of Appleton Laboratory, Swindon, UK.
monopile-supported offshore wind turbines. Kuhn M (2002) Offshore wind farms. In Wind Power Plants:
International Journal of Physical Modelling in Fundamentals, Design, Construction and Operation
Geotechnics 11(2): 28–68, http://dx.doi.org/10.1680/ijpmg. (Gasch R and Twele J (eds)). Springer, Heidelberg,
2011.11.2.58. Germany, pp. 365–384.
Bhattacharya S, Cox J, Lombardi D and Muir Wood D (2013a) LeBlanc C (2009) Design of Offshore Wind Turbine Support
Dynamics of offshore wind turbines on two types of Structures – Selected Topics in the Field of Geotechnical
foundations. Proceedings of the Institution of Civil Engineering. PhD thesis, Aalborg University, Aalborg,
Engineers – Geotechnical Engineering 166(2): 159–169, Denmark.
http://dx.doi.org/10.1680/geng.11.00015. Lombardi D, Bhattacharya S and Muir Wood D (2013) Dynamic
Bhattacharya S, Nikitas N, Garnsey J et al. (2013b) Observed soil–structure interaction of monopile supported wind
dynamic soil–structure interaction in scale testing of turbines in cohesive soil. Soil Dynamics and Earthquake
offshore wind turbine foundations. Soil Dynamics and Engineering 49: 165–180.
Earthquake Engineering 54: 47–60. Lowe J (2012) Hornsea met mast – A demonstration
Comodromos EM, Papadopoulou MC and Rentzeperis IK of the ‘twisted jacket’ design. Proceedings of the
(2009) Effect of cracking on the response of pile test Global Offshore Wind Conference, ExCel London,
under horizontal loading. Journal of Geotechnical London, UK.
and Geoenvironmental Engineering ASCE 135(9): Matlock H (1970) Correlations for design of laterally loaded
1275–1284. piles in soft clay. Proceedings of the 2nd Annual Technology
Conte E, Troncone A and Vena M (2013) Nonlinear Conference, Houston, TX, USA, pp. 577–587.
three-dimensional analysis of reinforced concrete piles Mindlin R and Deresiewicz H (1953) Elastic spheres in contact
subjected to horizontal loading. Computers and under varying oblique forces. ASME Journal of Applied
Geotechnics 49: 123–133. Mechanics 20: 327–344.
Conte E, Troncone A and Vena M (2015) Behaviour of flexible O’Sullivan C, Cui L and O’Neil S (2008) Discrete element analysis
piles subjected to inclined loads. Computers and of the response of granular materials during cyclic loading.
Geotechnics 69: 199–209. Soils and Foundations 48(4): 511–530.

11

Downloaded by [ UNIVERSITY OF CAMBRIDGE] on [23/09/16]. Copyright © ICE Publishing, all rights reserved.
Engineering and Computational Mechanics Soil–monopile interactions for offshore
wind turbines
Cui and Bhattacharya

Reese L, Cox WR and Koop FD (1974) Analysis of laterally 7th Annual Technology Conference, Houston, TX, USA,
loaded piles in sand. Proceedings of the 6th Annual OTC paper no. 2312, pp. 671–675.
Technology Conference, Houston, TX, USA, Paper no. Yu LQ, Wang LZ, Guo Z et al. (2015) Long-term dynamic
2079, pp. 473–483. behavior of monopile supported offshore wind turbines in
Reese L, Cox WR and Koop FD (1975) Field testing and analysis sand. Theoretical and Applied Mechanics Letter 5(2):
of laterally loaded piles in stiff clay. Proceedings of the 80–84.

HOW CAN YOU CONTRIBUTE?


To discuss this paper, please email up to 500 words to the
editor at journals@ice.org.uk. Your contribution will be
forwarded to the author(s) for a reply and, if considered
appropriate by the editorial board, it will be published as
discussion in a future issue of the journal.
Proceedings journals rely entirely on contributions from
the civil engineering profession (and allied disciplines).
Information about how to submit your paper online
is available at www.icevirtuallibrary.com/page/authors,
where you will also find detailed author guidelines.

12

Downloaded by [ UNIVERSITY OF CAMBRIDGE] on [23/09/16]. Copyright © ICE Publishing, all rights reserved.

You might also like