Sunteți pe pagina 1din 86

Seismic Analysis on

Monopile Based Offshore


Wind Turbines Including
Aero-elasticity and
Soil-Structure Interaction

Master of Science Thesis

Faculty of Civil Engineering and Geosciences


Delft University of Technology

Author: Maria Luisa Rosales Gonzalez

Committee: Prof.Dr. Andrei Metrikine TU Delft - Chairman


Dr.ir. Apostolos Tsouvalas TU Delft
Dr.ir. Karel van Dalen TU Delft
Dr.ir. Paul van der Valk Siemens Wind Power
ir. Jan Mariens Siemens Wind Power

July 13, 2016


”Do not weep; do not wax indignant. Understand.”
– Baruch Spinoza

Cover Picture: Diego Mazo (Gracias Chu ♥)


Summary

With the depletion of traditional energy sources, the usage of renewable energy alternatives has
increased over the past years. Among them, wind energy is considered as one of the most econom-
ical and environmentally friendly alternatives, and is seen as one of the main sources of energy in
the future.
A very promising field in the area of wind energy is offshore wind energy. Its currently high
costs in comparison with onshore wind energy are forecasted to decrease in the coming years.
Moreover, an acceleration in the deployment of the offshore wind industry is expected not only in
Europe, but also in other regions such as Asia and America.
With such development, which is also linked to an increase in the capacity of the turbines, one
topic that has only been recently addressed is the effect of earthquakes on offshore wind turbines
(OWT). There exists no clear guideline on how a seismic analysis should be performed in the design
of OWTs as their dynamic behavior is far more complex when compared to regular buildings.
Several authors have studied this topic and have found that aspects such as the aerodynamic
damping of the turbine, the higher order mode influence or the soil-structure interaction should
be taken into account, as they have a significant influence on the response of the turbine due
to earthquake loading. However, there is a general lack of knowledge on how to include these
aspects in the analysis, especially in combination with regular OWT design loads. Moreover, it is
still unknown if earthquake loading could actually become design-driving for OWT design, which
is currently governed by wind and wave load cases. This thesis aims to fill this knowledge gap by
addressing all the aforementioned questions, up to a certain extent.
As a first approach into getting an insight on the effect of seismic loading on OWT design,
a linear finite element model of an OWT monopile was developed. This model accounts for the
soil-structure interaction by using a set of distributed linear springs and dashpots to model the
soil stiffness and damping, respectively. Moreover, these springs and dashpots are used to apply
a set of ground motions (as displacements and velocities), which were selected and matched to
a target response spectrum, and were subsequently calculated for different depths by using shear
wave one-dimensional ground response analysis.
The developed embedded monopile model was coupled with 4.0 MW wind turbine model using
BHawC; the existing Siemens Wind Power in-house code for wind turbine nonlinear analysis in the
time domain. This code combines elasticity and aerodynamics in order to calculate the loads
induced on the turbine components due to different load cases, while accounting for its different
operational states and the changes on the turbine dynamic properties because of them. Moreover, it
also models features and systems that are present in the actual Siemens wind turbines to optimize
its power production while minimizing the structural damages.
Initially, a flexible soil-pile model was assumed along with relatively stiff soil conditions. It was
found that the response is dominated by higher order eigenmodes during the earthquake event,
and first order tower bending modes after it has passed. Moreover, a large influence of blade
modes was observed along the whole height of the support structure, but particularly at the upper
part (tower top and blades), where significant differences in the response were noted between
both tower directions (fore-aft and side-side) and different modeled operational states. When
comparing the results with the actual turbine design loads, it was found that the tower top and the
blades (in the edgewise direction) could become design-driving if earthquake loading is present.

i
Due to the influence of the soil properties in the overall system, two additional soil cases
(medium and soft) were subsequently considered. It was found that the soft soil led to the high-
est load due to an amplification of the response in the frequency range in which the turbine was
mostly excited. The effect of the ground response analysis could also be observed, as the response
was further amplified for frequencies close to the fundamental frequencies of the soil deposit.
Furthermore, new analyses were performed with variations in the model such as the soil damp-
ing, the soil-pile model and the ground response model. It was found that the way the soil is
modeled can have a large impact in the design loads, particularly regarding the choice model-
ing approach taken for the soil-structure interaction (the spring and dashpot coefficients), albeit
the choice of the dashpot coefficients alone did not have a significant influence on the results.
Moreover, it was found that modeling the monopile as a single discrete spring and dashpot system
through its equivalent dynamic impedances can lead to an important underestimation of the loads,
which can be compensated by applying the actual monopile displacements and rotations induced
by the earthquake loading, to account for the kinematic soil-structure interaction.
As a main outcome of this study, it was concluded that earthquake loading can in fact become
design-driving, particularly at the upper portion of the turbine due to the influence of the higher
modes. In this context, it was highlighted that aspects such as the blades and the different turbine
operational states should be included in the analysis, as they have a significant influence on the
results. The importance of an accurate soil identification and modeling, along with a proper choice
of ground motion time series were emphasized. Finally, recommendations for future research
were given , particularly regarding the inclusion of aspects such as non-local soil-pile interfaces,
soil nonlinearities, wave loading and frequency domain methods in the analysis.
Acknowledgements

First and foremost, I would like to thank my parents, John and Adriana. Thanks to their effort
I came to The Netherlands to pursue my Master’s degree, and thanks to their support and love I
am graduating, two years later. Making you both proud and happy is my biggest accomplishment :)

Thanks to my thesis chairman, Mr Metrikine, for introducing me into the awesome world of struc-
tural dynamics and for always pushing me into looking deeper and challenging my own results,
ever since the structural dynamics course. I can still remember my friends telling me how crazy
I was for doing my thesis on the hardest course of my master. Maybe I am, but I managed to do
it somehow! Also, thanks to Apostolos, my TU Delft supervisor, for his support, guidance, and
constructive criticism, especially at the moments where I felt lost during this whole thesis process.

Thanks to Paul and Jan, my supervisors from Siemens Wind Power. I think we were a great team,
I learned a lot from you guys and I hope I have managed to teach you a couple of things as well.
Thanks for being so patient with me and my complete ignorance on wind energy matters. Also,
thanks to the whole Siemens Wind Power Department. You made me feel welcome since the very
first day (which is probably the reason why I was always at the office?), supported me through the
different stages of my thesis and introduced me to very interesting engineering concepts such as
the physics of football or the “Matlab peak”. I had a lot of fun with you guys!

Thanks to my master friends, and in general, to all the nice and lovely people that I have met
during these two years at TU Delft. Somehow, you made me feel at home, and helped me maintain
my emotional stability on normal levels. I made friends for the rest of my life, and memories that
will last forever. Overall, it has been an amazing experience, and the best is yet to come!

Last but not least, I would like to thank coffee. Dear coffee, without you none of this would
be happening.

iii
iv
Contents

Summary i

Acknowledgements iii

List of Figures vii

List of Tables ix

Nomenclature xi

1 Introduction 1
1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Explanation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3 Scope of the Thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.4 Goal of the thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.5 Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

2 Existing Literature 5
2.1 Codes and Regulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.1.1 European Regulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.1.2 Japanese Regulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.2 Existing Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.2.1 Aeroelastic Codes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.3 Soil Structure Interaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

3 Formulation of Model for Seismic Analysis 11


3.1 Equations of Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
3.1.1 Introduction: Single Degree of Freedom systems . . . . . . . . . . . . . . . 11
3.1.2 General Formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3.2 Soil Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3.2.1 Equivalent Spring and Dashpot coefficients . . . . . . . . . . . . . . . . . . 13
3.2.2 Continuous vs. Discrete Modeling . . . . . . . . . . . . . . . . . . . . . . . . 14
3.3 Ground Response Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.4 Finite Element Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.4.1 Embedded Monopile Model . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3.4.2 Continuous Springs and Dashpots modeling . . . . . . . . . . . . . . . . . . 20
3.4.3 Integration into Aeroelastic Code . . . . . . . . . . . . . . . . . . . . . . . . 22

4 Selection and Adjustement of Ground Motions 23


4.1 Introduction to Japanese Seismicity . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
4.2 Selection Criteria . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
4.2.1 Rotation of Time Records . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
4.3 Spectral Matching . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

v
5 Selection of Soil Parameters 35
5.1 Soil Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
5.2 Soil Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
5.2.1 Pile (flexible) model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
5.2.2 Caisson (rigid) Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
5.2.3 Feasibility of Soil Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
5.2.4 Sensitivity to Soil Coefficients . . . . . . . . . . . . . . . . . . . . . . . . . . 41

6 Preamble of Simulations 43
6.1 Wind Turbine Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
6.2 Operational States . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
6.2.1 Control Systems of Normal Operation . . . . . . . . . . . . . . . . . . . . . 44
6.3 Load Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
6.3.1 Wind Loads . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
6.3.2 Wave Loads . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
6.3.3 Seismic Loads . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
6.4 Reference Model and Parametric Study . . . . . . . . . . . . . . . . . . . . . . . . . 47

10 Final Conclusions and Recommendations 51


10.1 General Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
10.2 Answers to Research Questions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
10.2.1 Sub-Questions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
10.2.2 Main Question . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
10.3 Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

A Element properties for Finite Element Model 55


A.1 Degrees of freedom . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
A.2 Stiffness Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
A.3 Shape Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

B BHawC Load Case Table 59

Bibliography 65
List of Figures

2.1 Integrated BHawC monopile model. . . . . . . . . . . . . . . . . . . . . . . . . . . 9

3.1 1 degree of freedom mass-spring-dashpot system with a prescribed ground motion. 11


3.2 2 degrees of freedom mass-spring-dashpot system with a prescribed ground motion,
including soil-structure interaction. . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3.3 Portion of embedded monopile subjected to a force. . . . . . . . . . . . . . . . . . . 14
3.4 Continuous vs. Discrete modeling of soil-pile system. . . . . . . . . . . . . . . . . . 15
3.5 One Dimensional ground response analysis. . . . . . . . . . . . . . . . . . . . . . . 17
3.6 Explanation of outcropping bedrock motion . . . . . . . . . . . . . . . . . . . . . . 18
3.7 General components of a wind turbine. . . . . . . . . . . . . . . . . . . . . . . . . . 19
3.8 Components of full finite element model. . . . . . . . . . . . . . . . . . . . . . . . . 19
3.9 Degrees of freedom of selected finite element. . . . . . . . . . . . . . . . . . . . . . 20

4.1 Map of plates and trenches in Japan . . . . . . . . . . . . . . . . . . . . . . . . . . 23


4.2 Hypocenters of earthquakes in Japan with magnitude MW greater than 5 for the
period 1964-2011. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
4.3 Maximum design Response Spectrum for a level 2 seismic motion and a damping
ratio of 5 %, as specified in the Japanese building code. . . . . . . . . . . . . . . . . 25
4.4 Graphical explanation of seismic and engineering bedrock. . . . . . . . . . . . . . . 26
4.5 Chosen events SRSS response spectrum, compared with Target Spectrum. . . . . . 27
4.6 Response Spectrum for both directions of selected records, after spectral matching. 29
4.7 Unmatched vs. Matched ground motion time series - Seismic Seed 1 (Christchurch
record). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
4.8 Unmatched vs. Matched ground motion time series - Seismic Seed 2 (Chi-Chi record). 31
4.9 Unmatched vs. Matched ground motion time series - Seismic Seed 3 (Chuetsu-Oku
record). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
4.10 Power Spectral Density of each Seismic Seed acceleration record. . . . . . . . . . . 33

5.1 Frequency dependent dashpot coefficient from pile model and its linearization for
time domain analysis. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
5.2 Comparison of dashpot coefficient between simple and elaborated model for the
frequency range of interest. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
5.3 Active length / Pile diameter for different soil stiffnesses. . . . . . . . . . . . . . . . 38
5.4 Depiction of selected soil-pile models. . . . . . . . . . . . . . . . . . . . . . . . . . 39
5.5 Static Timoshenko beam model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
5.6 Normalized displacements and rotations over pile length - Pile model. . . . . . . . 40
5.7 Normalized displacements and rotations over pile length - Caisson model. . . . . . 41

6.1 Turbine model to be used in the analysis. . . . . . . . . . . . . . . . . . . . . . . . . 44


6.2 Typical power curve of a wind tubine. . . . . . . . . . . . . . . . . . . . . . . . . . 45
6.3 Power curve of modeled tubine, including HWRT feature. . . . . . . . . . . . . . . 45

vii
A.1 Degrees of freedom of finite element. . . . . . . . . . . . . . . . . . . . . . . . . . . 55
List of Tables

4.1 Selected ground motion records. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26


4.2 Correction parameters for selected ground motion records. . . . . . . . . . . . . . . 27
4.3 Equivalence between selected earthquake records and seismic seeds. . . . . . . . . 29

5.1 Selected soil types for the analysis. . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

6.1 Section properties on points of interest of support structure. . . . . . . . . . . . . . 43

ix
x
Nomenclature

Latin symbols
l Soil element length m
a0 Dimensionless frequency
c Damping N s/m
cx Distributed translational dashpot coefficient N s/m2
cθ Distributed rotational dashpot coefficient N s/Rad m
Cx Discrete translational dashpot coefficient N s/m
Cθ Discrete rotational dashpot coefficient N s/Rad
d Pile Diameter m
E Young’s modulus Pa
f Frequency Hz
fn Natural Frequency Hz
F Force N
g Acceleration of gravity m/s2
G Shear modulus Pa
H Thickness of soil deposit m
k Stiffness N/m
kω Wave number rad/m
kR Rayleigh wave number rad/m
kx Distributed translational spring coefficient N/m2
kθ Distributed rotational spring coefficient N/Rad m
Kx Discrete translational spring coefficient N/m
Kθ Discrete rotational spring coefficient N/Rad
la Active length m
L Embedded monopile length m
Le Beam element length m
m Mass kg
MW Moment Magnitude Scale
M Bending Moment kN m
N Shape function
Rjb Horizontal distance to the surface projection of the rupture plane km
Rrup Closest distance to the rupture plane km
s Generalized coordinate
t Time s
T Transfer function
Tr Return period years
Tn Natural period s
u Displacement m
u̇ Velocity m/s
ü Acceleration m/s2
vs Shear wave velocity m/s
vs30 Average shear wave velocity between 0 and 30 meters depth m/s
vP Primary wave velocity m/s
V Shear Force kN

xi
Greek symbols
α Complex impedance ratio
γ Shear strain
η Viscosity Pa s
θ Rotation Rad
ν Poisson’s ratio
ξ Damping ratio %
ρ Density kg/m3
τ Shear stress Pa
ϕ Rotation angle for ground motions Rad
ω Angular frequency rad/s

Subscripts
b Interface
e Element
ex External
f Free
g Ground
r Relative
s Soil

Superscripts
∗ Complex-valued
r Rotated
T Transposed

Abbreviations
AWEA American Wind Energy Association
ASCE American Society of Civil Engineers
BEM Blade Element Momentum
DLC Design Load Case
FE Finite Element
FEM Finite Element Method
FFT Fast Fourier Transform
DNV Det Norske Veritas
GL Germanischer Lloyd
GWEC Global Wind Energy Council
HWRT High Wind Ride Through
IEC International Electrotechnical Comission
IISEE International Institute of Seismology and Earthquake Engineering
JMA Japan Meteorological Agency
JSCE Japanese Society of Civil Engineers
MDOF Multiple Degree Of Freedom
MSE Mean Square Error
NREL National Renewable Energy Laboratory
OWT Offshore Wind Turbine
PEER Pacific Earthquake Engineering Research Center
PGA Peak Ground Acceleration
PSD Power Spectral Density
RNA Rotor Nacelle Assembly
RSN Record Sequence Number
SSI Soil-Structure Interaction
SWT Siemens Wind Turbine
SLS Service Limit State
SR Sway-Rocking
SDOF Single Degree of Freedom
SRSS Square Root of the Sum of the Squares
ULS Ultimate Limit State
Chapter 1

Introduction

The depletion of traditional energy sources has led to an increase in the usage of renewable energy
sources to supply the societal needs. Wind energy stands out as one of the most economical and
clean energy sources, with prices per kWh that can be even lower than traditional energy sources
such as coal and gas. When adding this to other advantages such as stability and inexhaustibility,
wind power can be considered as a very competitive energy source.
A very promising field in the area of wind energy is offshore wind energy. Although it is
currently more expensive and than onshore wind energy, it also comes with many more benefits.
Wind speeds at the sea are much higher than inland, which translates into larger power generation
capacities. Moreover, the space available offshore makes it possible to install larger turbines, while
avoiding the social costs of placing a wind turbine near a populated area.
Offshore wind energy is a sector that has been growing steadily over the past years, despite of
the economic crisis that the world has gone through. According to BTM consult [1], offshore wind
power capacity will to reach a total of 75 GW worldwide by 2020, with large contributions from
China and the United states. Due to its environmental advantages, combined with the predicted
reduction of its costs [2], it is seen by a majority of people as one of the main sources of energy in
the future.
One topic related to this subject that has only been recently addressed is the vulnerability of
offshore wind turbines to earthquakes. This topic could become of increasing importance in the
future years, due to the forcasted high developement of the Offshore Wind market in countries
with high seismic activity such as China and Japan [3], combined with the introduction of higher
capacity turbines which could lead to an increase on its seismic loads.
It has been found that, due to the particular structural characteristics of offshore wind turbines,
they cannot be seismically analyzed in the same way as regular buildings. Several authors have
attempted to model earthquakes on onshore and offshore wind turbines by taking into account
the characteristics of their dynamic behavior. However, tasks such as 1) including the soil or
2) considering the different operational states of the turbine in the models cannot be done in a
straight-forward way. Additionally, there are no measurements of the behavior of OWT (Offshore
Wind Turbines) during an earthquake, that can be used to validate such models. There is, overall,
a general lack of knowledge on how the turbine behaves during an earthquake and which aspects
are more relevant for the analysis.

1.1 Motivation
Siemens Wind Power is one of the major turbine supplier of offshore wind energy in the world,
with a combined capacity of 4.8 GW and more than 1500 turbines installed not only in Europe, but
all around the globe [4]. As a part of its constant market expansion, they are providing turbines
for asian offshore wind farms. It is known that such areas are prone to considerable seismic events,

1
2 1. INTRODUCTION

which could affect the turbine design.


Siemens uses its in-house code BHawC to perform non-linear dynamic analysis for its different
OWT projects. Such code can perform batch simulations in the time domain under different condi-
tions of wind speeds, wave loads and operational states. However, it currently has no capabilities
for seismic analysis. Moreover, there is a general lack of knowledge on how necessary it is to in-
clude earthquake loading into the OWT dynamic analyses, how it compares to “normal” loads for
an OWT (such as wind or waves) or which aspects are more important to consider if such analyses
are to be undertaken.
As a first step into implementing seismic loads into the integral analysis and design process of
an OWT, it is necessary to get a general idea on the behavior of a turbine during an earthquake.
Aspects such as which component of the turbine is more sensitive to earthquake loads, which
operational state is the most critical or which is the most realistic way to apply the earthquake
load could play a major role into developing more advanced tools, as well as new BHawC features
to include seismic loading into the turbine analysis.

1.2 Explanation
The first part of this thesis consists of the development of an Offshore Wind Turbine monopile-
based finite element model. The model is constructed by taking into account past conclusions and
experiences from different authors in the field of seismic analysis of wind turbines, in aspects such
as including aeroelastic interaction, including soil structure interaction, full RNA (Rotor Nacelle
Assembly) modelling and preference in time domain over frequency domain for the dynamic anal-
ysis. This model couples a nonlinear model developed with the existing in-house Siemens Wind
Power aeroelastic code BHawC with a linear monopile model which includes the earthquake loads.
In the second part of the thesis, this developed model, further referred to as “reference model”,
is used as a starting point to gain a better understanding on the general effect of seismic loading on
an OWT, while considering the changing dynamic properties (eigenfrequencies and their respective
modal damping ratios) of the OWT due to its own operation. An analysis on the effect of the
combined earthquake and wind loading on the dynamic response is also included, and the results
are compared with the case in which the seismic load is not present.
Several authors have pointed out the importance of including soil-structure interaction into
wind turbine seismic analysis. Moreover, as earthquake loads are transmitted to the structure
through the ground, the way in which soil the is modeled has a significant impact on the anal-
ysis results. In order understand how sensitive the dynamic response can be to soil modeling, a
parametric study is conducted. Starting from the reference model, the soil stiffness, as well as the
description of the soil and the pile, are varied by using different soil-structure interaction models.
The differences in the response of the system due to these variations are subsequently analyzed.

1.3 Scope of the Thesis


When constructing the model for this study, special care was taken into making it as realistic as
possible. However, some aspects were overly simplified in order to reduce the complexity of the
model and to be able to implement it along with the aeroelastic interaction. There is a large un-
certainty on the accuracy of the soil modeling that was used in this thesis, as it was not possible to
validate it with experimental results. Additionally, some parameters, such as the soil profile, were
defined in the simplest possible way, in order to make the analysis of the results feasible. More-
over, due to the lack of concise regulations regarding this topic, a “building code approach” was
followed to perform the time domain analysis and to choose the earthquake time series that were
applied. This contradicts the fact that wind turbines are much more complex dynamic structures,
which cannot be evaluated in a similar way as buildings.
1.4. GOAL OF THE THESIS 3

Due to these reasons, among others, the results of this thesis should be considered as purely
academic and can not be used directly in a real turbine design, as the values of the results greatly
depend on factors of the model that have a great level of uncertainty (such as the applied earth-
quake motions or the way the soil was modeled).
However, these results can be used as a guideline for future research on the area of seismic
analysis of OWT. Although the results may not be entirely accurate in terms of actual values,
they can give a useful insight on several aspects that should properly be taken into account when
performing an actual design of a turbine against seismic loading.

1.4 Goal of the thesis


The main goal of this thesis is to answer the following research question:

Can earthquake loading become design driving for Offshore Wind Turbines?

This question cannot be answered by, for example, looking into load values and comparing it to the
turbine structural capacity, as such load values are very sensitive to several aspects of the model
that are not entirely accurate (such as the soil modeling). This is why a more global approach is
followed, by analyzing the global behavior of the turbine in both time and frequency domain.
In order to be able to answer the proposed research questions, several specific sub-questions
are addressed:

1. What is the influence of the operational state on the seismic response of an OWT?
2. Which is the most critical part of an OWT when seismic events occur?

3. How do the soil properties influence the dynamic response of an OWT?


4. How sensitive is the dynamic response of an OWT on the way the soil is modelled?

1.5 Outline
The structure of the thesis is be as follows:
Chapter 2 shows a review on the existing literature and regulations regarding seismic analysis
of wind turbines. Based on the conclusions drawn by previous authors, a FE model of an OWT
capable of taking into account seismic loading was developed, as described in Chapter 3. The
seismic loads to be applied to this model were calculated by using time series of actual earthquake
events, which were selected and adjusted as described in Chapter 4. The selected soil parameters
for the reference model and the subsequent parametric study are described in Chapter 5. A brief
preamble to the results, including information about the wind turbine model and its assumptions
is shown in Chapter Chapter 6. An extensive analysis of the seismic response of the reference
model is shown in Chapter 7 and an overview on its variation with the soil properties is presented
in Chapter 8. Furthermore, the variation of the response with other relevant modeling parameters
is described in Chapter 9. Finally, the main conclusions of the thesis, and recommendations for
future studies are presented in Chapter 10.
For this report, Chapters 7 to 9 were omitted for confidentiality reasons.
4
Chapter 2

Existing Literature

Currently, there is still a knowledge gap in the area of seismic analysis of wind turbines. The first
approaches which considered the turbines as regular civil structures, have proven to be inaccu-
rate, as they do not take into account the dynamic characteristics of the turbine behavior itself.
Furthermore, it has been also found that the interaction with the soil plays a major role in the dy-
namic response of the turbine, in a much greater extent than what is usually the case for a regular
building.
This chapter will offer a quick overview of the existing literature related to seismic analysis
of wind turbines. First, a summary of the existing codes and regulations specifically related to
earthquakes and wind turbines is presented. Furthermore, several models developed by various
authors are discussed. Finally, a brief description of what has been achieved related to seismic
analysis and soil structure interaction is shown.

2.1 Codes and Regulations


2.1.1 European Regulations
Probably the most worldwide known wind turbine regulations and standards are the Design Re-
quirements for Wind Turbines from the International Electrotechnical Commission IEC [5], the
Guideline for Certification of Offshore wind Turbines from Germanischer Lloyd [6] and the Guide-
lines for Design of Wind Turbines by Det Norske Veritas and Risø National Laboratory [7].
The IEC requirements, which are probably the most widely known and referenced, state that
earthquake effects should only be taken into account on sites where earthquake assessment analy-
sis is required according to local building codes. For such cases, the required return period for the
design event is 475 years. The analysis (that can be done in the time or the frequency domain)
should be performed according to the local seismic code, or by taking into account a modal mass of
85% of the total turbine mass if such code is missing. A simplified analysis approach is suggested,
where the load for the system should be calculated using a concentrated mass at the tower head
and an acceleration from the design response spectrum (calculated with a damping ratio of 1%)
of the first eigenfrequency of the tower. This load should be added the highest load between both
the normal power production and emergency stop load cases.
The GL guideline establishes that the earthquake calculation is only needed on sites with a
design peak ground acceleration exceeding 0.05 g. In this case, both a response spectrum analysis
(considering at least the first 3 modes of the turbine) or a time history analysis (applying at least
6 ground motion time series) are accepted. The required return period of the seismic event is 500
years. It is stated that, in case of earthquake analysis, the effect of the earthquake loading on the
foundation has to be evaluated, as well as the effect of soil.
The guidelines from Risø and Det Norske Veritas state that earthquakes should be considered in

5
6 2. EXISTING LITERATURE

seismically active areas. They suggest an analysis using the response spectrum method including
one vertical and one horizontal component of the seismic action (due to symmetry) where the
wind turbine can be represented by a concentrated mass on top of a vertical beam.
Most recently, the American Wind Energy Association AWEA, along with the American Society
of Civil Engineers ASCE, developed the Recommended Practice for Compliance of Large Land-
Based Wind Turbine Support Structures [8]. It states that seismic analysis should be done accord-
ing to the local seismic code or the IEC 61400-1. Also, it is indicated that the spectral response
acceleration (used in the response spectrum method) should be amplified to account for lower
system damping (taking into account that the given response spectrum is calculated for a damping
ratio of 5%). These damping ratio values should be taken as 1% for load combinations without
operational loads, and of 5% for load combinations with operational loads, to take into account the
aerodynamic damping effects. Also, load combinations for simultaneous wind turbine operational
loads and earthquake loads are presented. For the sum of both loads, a multiplication factor of
0.75 is suggested.
As for the analysis, this standard suggests that horizontal forces should be applied at the tur-
bine’s center of gravity and at the tower joints, in case an equivalent horizontal force method
(which consists on replacing the effect of the earthquake by a set of equivalent horizontal forces in
equilibrium with the seismic base shear) is chosen. Also, a response modification factor R (equiv-
alent to the Eurocode q factor, which accounts for the structure ductility in terms of the internal
forces of an elastic system compared to those of the real inelastic structure) of 1.5 is suggested, to
account for topics such as conservatism and soil-structure interaction. As for time history analysis,
it is suggested to be done according to the ASCE 7-05 requirements.

2.1.2 Japanese Regulations

In a more specific context, the Japanese Society of Civil Engineers JSCE has its own guidelines for
design of wind turbines support structures and foundations, where a much more detailed explana-
tion on how to include seismic loading is given. Here, a time-history analysis is assumed, to take
into account the predominant response higher dynamic modes.
It is stated that seismic loads on wind turbines have to be evaluated for both operational and
non-operational states. This is done by adding the seismic loads and the loads corresponding to the
annual mean wind load. In case of aeroelastic models (like the BHawC code), seismic loads should
be determined in a time-history aeroelastic analysis that assumes a wind speed corresponding to
the annual mean wind load.
A time-history analysis should be performed along with a SR (sway-rocking) model in order to
take into account the effect of soil structure interaction. A target response spectrum is defined for
Level 1 (reference peak ground acceleration of 0.16 g and return period of 50 years) and Level
2 (reference peak ground acceleration of 0.32 g and return period of 500 years) seismic vertical
and horizontal motions, for a damping ratio of 5%. The input for the seismic analysis can be from
either artificial motion matching this predefined response spectra, or from observed motions with
frequency content modified to match the same target spectra.
Furthermore, the Japanese code gives guidelines on how to calculate the response of the soil
and the respective spring and dashpot coefficients to use in the SR model. The horizontal force
resulting from the time domain analysis of the SR model has to be applied to a decoupled model of
the pile analyzed using the force-displacement method, in order to study the foundation behaviour
from both intertial and kinematic effects. Alternatively, an aggregated pile model can also be done,
in which the turbine superstructure is modeled along with the pile foundation, and the ground
displacements are applied directly on the distributed soil springs.
2.2. EXISTING MODELS 7

2.2 Existing Models


The problem of earthquakes and wind turbines has gained significant importance over the past
few years. As the guidelines given by current codes and regulations are not very specific and
mostly based on guidelines for seismic analysis of buildings, numerous authors have tried different
approaches to get a better understanding of this issue.
As early as in 2002, the problem of earthquake loadings on wind turbines started being anal-
ysed. The simplest way to do so was by analyzing the tower with a lumped mass to account for the
RNA. Bazeos et al.[9] compared detailed finite element models with simplified lumped mass mod-
els of a 450 kW turbine under combined seismic and quasi-static wind loads, and determined that
earthquake loading did not lead to a critical response. Kiyomiya et al.[10] also found that earth-
quake loading was not critical for an existing 1.65 MW turbine and a combined earthquake and
static wind case, although emphazizing that dynamic wind loads and other kinds of loads should
also be taken into account, as its combined action could lead to higher moment capacity demands.
Lavassas et al [11] also found no critical response due to earthquakes, but pointed out how they
may become critical for turbines built on earthquake prone areas. Other studies ([12, 13, 14]) have
come to this same conclusion, especially for the current situation of growing size of the turbines.
Apart from lumped mass models, several full models have been proposed, apart from the al-
ready existing aeroelastic analysis tools, which will be discussed later. Zhao and Maißer [15]
presented a multi-body system model to account for the different components of the turbine. Ishi-
hara and Sawar [16] compared the results from a simplified model and a full finite element model
including blades and determined that simplified models do not take into account the influence of
higher modes, which could be decisive for medium to large turbines. Diaz Nevarez and Suarez
[14] evaluated a V82 Vestas turbine using a model with beam elements to account for the tower
combined with rigid elements linked with hinges and springs representing the RNA. For the applied
time histories (used according to local guidelines) the turbine was significantly excited on its first
mode of vibration, which did not lead to and exceedance of the capacity on the tower (the most
critical part). Additionally, they found that the blades were the less critical part (in terms of seismic
loading) as they had been designed for much higher fatigue loads. However, it was recommended
to evaluate the combined action of seismic and wind loading, as this could lead to governing loads
on the top parts of the tower (which agrees with other authors using full models such as Jin et al
[17]). Indu and Binu [18] modelled a 65 kW turbine using different levels of simplification and
compared the results to the experimental results obtained by Prowell et al. [19]. It was concluded
that simplified models show a considerable deviation from more refined models (which agree with
the numerical results) and might not be reliable for larger turbines.
Probably one of the most complete studies performed in the topic of wind turbines and earth-
quakes was done by Prowell [20] who dedicated his whole PhD to study different aspects of this
topic. Along with the University of California, he performed several shake table experiments on
a full scale 65 kW turbine, whose results are publicly available and are proven useful to calibrate
numerical models. By using this, he found that the first mode was mostly prominent in the seismic
response of small turbines, whereas for increasing size, higher order modes play a more important
role in the overall response. Furthermore, he performed several dynamic ambient vibration tests
on existing turbines. This allowed to verify several phenomena concluded from models of differ-
ent authors, such as the assumed independence between the main fore-aft and side-side modes,
the variation of the damping ratios for higher modes with respect to the first modes and the in-
crease in damping due to aerodynamic interaction. Also, he introduced a seismic module into the
FAST code, for which he showed that parked-state simulations agree with analysis performed by
any structural analyses software. Moreover, he concluded that with increasing size of turbines,
earthquake loading could probably become design driving due to tower moment demand (and
in a similar way, showed that even for very strong earthquakes is very unlikely that blades can
become design-driving). Additionally, as simulations using FAST are performed assuming a fixed
base, he also highlighted the importance of soil structure interaction as it could change this kind
of behavior.
8 2. EXISTING LITERATURE

From a very early stage, it was pointed out that time domain simulations were preferred over
frequency domain and modal analysis. The major difference between both approaches is the as-
sumption of a system damping in the case of the frequency domain method. However, in the case
of wind turbines, the system damping can be greatly affected by the aerodynamic effects, which
leads to different damping ratios for different directions or operational states [21]. Ritschel et
al. [22] compared the results from a simplified modal analysis and a full time domain simulation
using the aeroelastic code FLEX5, and determined that the modal approach led to conservative
results near the tower base, while under estimating the blade loads. It was concluded that these
results were due to the “oversimplification” of the model in the modal approach, by not taking
into account higher modes where tower and blades degrees of freedom interact with each other.
Similar analysis done by Witcher [21] using GH Bladed determined a large increase of tower base
demands for time domain analysis, specially in the parked state case. It was concluded that time
domain analysis is important to account for aeroelastic interaction. Alati et al. [23] also used GH
Bladed to analyze turbines with different foundation models and found differences in the location
of the highest moment demand, which highlights the importance of full model simulations in time
domain to properly capture the earthquake effect. Several other studies have concluded that exist-
ing response spectrum methods from building codes cannot capture the acceleration characteristics
for wind turbine behaviour [16, 24, 25]. Some attempts [26, 16] have been done account for the
aerodynamic damping by means of an spectral amplification factor, in order to be able to use the
response spectra as given in such codes.

2.2.1 Aeroelastic Codes


Over the past few years, different full system models have been developed to analyze offshore
wind turbines by modeling its complex behavior, combining aerodynamics, hydrodynamics and
structural dynamics, along with features such as pitch control and soil structure interaction. This
combined behavior is commonly known as aero-elasticity, and several codes and tools are available
to analyze wind turbines by this approach. Accounting for the recent need to take into account
seismic loading into wind turbine analysis and design, seismic modules have been developed in
order to include earthquake effects into such tools. Here, some aeroelastic codes with seismic ca-
pabilities will be briefly discussed.

GH Bladed [27], produced by DNV GL (former Garrad Hassan), uses a modal model with a fi-
nite number of degrees of freedom. GH Bladed runs modal calculations in the time domain for
which the structural loads are calculated for each one of the turbine’s main components (rotor,
power train and tower). A seismic module for GH Bladed was implemented by Witcher [21] in
which a response spectrum can be defined or generated. This way, earthquake forces (calculated
from the spectral accelerations) are included in the equation of motion for each one of the modal
degrees of freedom.
The FAST (Fatigue, Aerodynamics, Structures and Turbulence) code [28], developed and main-
tained by the United States National Renewable Energy Laboratory (NREL), allows time domain
simulation of a wide range of wind turbine configurations, by means of a combined modal and
multibody dynamics formulation. It is worth mentioning that FAST has been validated by DNV
GL for calculating design loads of wind turbines [29]. Prowell and Asareh [30, 31] developed a
seismic module for FAST, in which a damped oscillator with a certain natural frequency is used to
reproduce the desired ground motion time histories at the base of the turbine.
Another wide known tool is FLEX5, developed by Stig Oye from the Department of Fluid Me-
chanics of the Technical University of Denmark. FLEX5 models wind turbines (using a modal
superposition method) in the time domain as a mechanical model with up to 28 degrees of free-
dom, in which the first two tower modes in each direction are taken into account. Ritschel et al.
[22] included earthquake loading in the code by applying the ground acceleration to each mass
element of the model.
It is noteworthy that none of these aeroelastic codes is capable to take soil structure interaction
2.3. SOIL STRUCTURE INTERACTION 9

into account.

Siemens in-house code BHawC (Bonus Energy Horizontal axis wind turbine Code) has been contin-
uosly in developement by Bonus Energy and Siemens Wind Power since 2003. The code represents
the wind turbine structure by a geometrically non-linear finite element model, in which the aerody-
namic effects are combined with extensions to take dynamic inflow behavior into account. BHawC
is able to model a wind turbine monopile foundation as flexible beam elements with non linear soil
(p-y) springs to take the soil structure interaction into account. This is illustrated in Figure 2.1.

Figure 2.1: Integrated BHawC monopile model.

However, BHawC currently has no capabilities for seismic analysis. An initial approach that has
been implemented within the code is by applying a force to a very large mass at the mudline, such
that the resulting acceleration according to F = ma is equal to the desired ground acceleration
time history to be applied. Interestingly, this was also the first approach by Prowell et al. [30] to
include earthquake loading into the FAST code.

2.3 Soil Structure Interaction


The interaction of the structure with the soil can dramatically affect the response of both over-
ground and underground structures against earthquake loading, especially in the case of soft soils
[32]. In the case of wind turbines, the guidelines from Risø and Det Norske Veritas [7] note the
importance of analyzing the soil structure interaction by including springs in the base of the tur-
bine to account for soil stiffness and its effect on the turbine. Such spring coefficients are given for
different kinds of foundations. In the case of earthquake loading, such coefficients should also ac-
count for a dynamic variation due to the frequency. On the other hand, for the analysis of monopile
foundations of wind turbines, it is recommended that the soil reaction is accounted by means of
distributed p-y springs along the length of the monopile.
Several authors have followed similar approaches as the one described by Risø and DNV to
account for soil structure interaction in their seismically loaded wind turbine models [9, 15, 13,
33]. Probably one of the first studies on the area of wind turbines and earthquake, done by Bazeos
et al. [9], included soil structure interaction as a set of discrete springs and dashpots at the base
of the turbine, along with an added mass due to the portion of soil that moves in phase with the
foundation. It was concluded that the interaction of the soil affected all, but mainly the higher
order modes of the turbine. Therefore, it was concluded that soil-structure interaction played an
important role on seismic analysis of OWT. Zhao and Maißer [15] arrived to a similar conclusion
when accounting for SSI on their multi-body model. Hongwang [13] also reported significant
10 2. EXISTING LITERATURE

influences on the tower top acceleration and tower base moment for 1.65 MW and 3 MW turbines
according to the ground motions suggested by the Chinese building code. It was also found that SSI
had much less influence on the vertical response of the turbine, when compared to the horizontal
response.
Another approach to model soil in order to account for SSI is by modeling a large part of the soil
as a continuum using 3D finite elements. This approach may avoid large errors that could occurr
in the soil-structure system at resonance, specially when using frequency-independent coefficients
[34]. Prowell [20] compared the response of a turbine with a fixed base against one supported
in such kind of soil model. A significant influence of the soil in the response was observed in the
case of soft soils. Althought the variations of natural frequencies lied within the safety margins,
maximum moment and shear demand changes were significant and could lead to the need of a re-
design. It was reccommended to conduct further SSI research as a component of seismic response
studies of wind turbines. Kjørlaug et al. [35] evaluated the response of his turbine model on a soil
continuum for different kinds of soils, and found that some soils lead to very large amplifications
of the accelerations and demands on the turbine. He pointed out the importance of these kind of
analyses, specially in the case where the tower eigenfrequencies coincide with the frequencies of
the soil deposit. Also, he evaluated the response of the turbine under wind load and combined
wind and earthquake load, and concluded that SSI was not as important for wind loads alone as it
was for wind and earthquake combined loads. Hacıefendioğlu [36] performed a parametric study
in which he analyzed a turbine model under earthquake loading for different soil conditions, and
found considerable increases in the displacement of the turbine in softer soils.
In the case of monopile foundations, there are discrete spring and dashpots coefficients avali-
able in the literature to account for the dynamic impedances of monopiles embedded in soils [37].
However, another option is to model the entire monopile with distributed springs and dashpots
along its embedded length, to account for the reaction of the soil. This approach, although as-
suming uncoupling between the different layers of the soil, is very convenient because it accounts
for the kinematic interaction between the soil and the structure as well. One common practice is
to use the p-y curves used in turbine foundation design, as it was done by Kim et al. [38], who
compared different methods of imputting the earthquake motion into these soil springs and deter-
mined that, when using this approach, ground motions have to be calculated for each soil layer
using ground response analysis. Alati et al. [23] also used p-y curves to compare different founda-
tion models (tripod and jacket) with flexible and fixed bases, deriving significant differences in the
responses for a flexible foundation compared with a fixed foundation, specially in terms of tower
base and blade root demand. Apart from the p-y curves, frequency dependent dynamic coefficients
for distributed springs and dashpots have been derived in the literature for piles embedded in soil
[39]. Based on these approximations, Kampitsis et al. [40] derived an assembly of frequency in-
dependant springs and dashpots to capture the characteristics of the soil reaction. By using this,
Sapountzakis et al. [41] modeled the nonlinear response of a 5 MW turbine under combined earth-
quake and wind loading by means of a beam formulation using the BEM method. It was found
that the monopile foundation led to a flexible structure, which translated into smaller stresses on
the structure. Also, the importance of soil-structure interaction and geometrical nonlinearities was
emphasized.
Chapter 3

Formulation of Model for Seismic Analysis

3.1 Equations of Motion


3.1.1 Introduction: Single Degree of Freedom systems
In order to get a better understanding on the effects of soil structure interaction in the seismic
behavior of a dynamic system, a single degree of freedom (SDOF) mass-spring-dashpot system
with a prescribed ground motion will be analyzed.

Figure 3.1: 1 degree of freedom mass-spring-dashpot system with a prescribed ground motion.

The system consists of a mass m, and a spring and a dashpot with coefficients k and c, respec-
tively. The mass is subjected to an external force Fex , and a ground displacement ug is prescribed
at the other side of the spring and dashpot.
By means of Newton’s second law, and taking into account that the spring and dashpot forces
depend respectively on the relative displacement ur = u − ug and velocity u̇r = u̇ − u̇g of the mass,
the equation of motion of the system can be derived. This derivation can be found in any structural
dynamics book [42].
mü + cu̇ + ku = Fex + cu˙g + kug (3.1)
This equation can also be written in terms of relative motions as:

mu¨r + cu˙r + kur = Fex − mu¨g (3.2)

The problem becomes more complex when the soil-structure interaction is introduced. A mass mb
is introduced to represent the “interface” degree of freedom of the system between the soil and
the structure. Therefore, this mass is attached to a soil spring and dashpot to represent the soil
stiffness ks and damping cs respectively. The ground motion is then prescribed at the end of these
soil spring and dashpot, in order to take into account the soil effect in the seismic response of the
system. A graphical representation of the system is shown in Figure 3.2.
By Newton’s second law, the equations of motion of the system are derived. For the mass mf :

mf ü + cu̇ + ku − cu¨b − kub = Fex (3.3)

11
12 3. FORMULATION OF MODEL FOR SEISMIC ANALYSIS

Figure 3.2: 2 degrees of freedom mass-spring-dashpot system with a prescribed ground motion, in-
cluding soil-structure interaction.

For the mass mb :

− ku − cu̇ + mb u¨b + (c + cs ) u˙b + (k + ks ) ub = −cs u˙g − ks ug (3.4)

Rewriting equations 3.3 and 3.4 in a matrix form:

Mü + (C + Cs ) u̇ + (K + Ks ) u = f + Cs u̇s + Ks us (3.5)

where:
       
mf 0 c −c 0 0 k −k
M= C= Cs = K=
0 mb −c c 0 cs −k k
       
0 0 u 0 Fex
Ks = u= f us = f=
0 ks ub ug 0
The forces exerted by the soil to the system due to the ground motion can be written as:

fs = Cs (u̇s − u̇) + Ks (us − u) (3.6)

Replacing equation 3.6 in equation 3.5 the following, more general expression is obtained:

Mü + Cu̇ + Ku = f − fs (3.7)

Some important remarks:

• The depicted dynamic system consists of two components: The structure itself, that has two
degrees on freedom (one “free” and one “interface” which is in contact with the soil), with
a mass M, a damping C and a stiffness K and the soil contribution, represented by a soil
stiffness Ks and damping Cs .

• The mass mb does NOT correspond to the mass of the soil, but to the mass of the ”interface”
degree of freedom of the structure.

• If the mass mb was equal to zero, the system cannot be depicted as a single degree of freedom
system like the one of figure 3.1 with an equivalent stifness k + ks and damping c + cs . The
reason for this is because the “interface” degree of freedom cannot be eliminated, as it is
the one that is in contact with the soil. Moreover, by doing this the structure stiffness k and
damping c would be directly related to the ground motions ug and u̇g , which is not correct, as
if would be equivalent to apply the ground motion directly on the structure and not through
the soil.

3.1.2 General Formulation


Equation 3.7 is a general equation that can be applied for larger systems, as long as the “free” and
“interface” degrees of freedom are still properly differentiated. In this formulation, the subscript f
is used for the ”free” degrees of freedom, and the subscript b is used for the ”interface” degrees of
3.2. SOIL MODELING 13

freedom. This means that the mass, damping and stiffness matrices of the structure may be written
as:      
Mff Mfb Cff Cfb Kff Kfb
M= C= K=
Mbf Mbb Cbf Cbb Kbf Kbb
And the stiffness and damping matrices of the soil may be written as:
   
0 0 0 0
Cs = Ks =
0 Cs,bb 0 Ks,bb
The displacement, velocity and acceleration vectors are, respectively:
     
uf u̇f üf
u= u̇ = ü =
ub u̇b üb
The ground motion vectors, which are nonzero at the ”interface” nodes, are then:
   
0 0
us = u̇s =
ug u̇g
And the external force vector, as the external forces are only applied on the ”free” nodes, is then:
 
f
f= e
0
On the other hand, the vector of interaction forces with the soil fs , which is, evidently, applied only
on the “interface” nodes, depends on the relative displacement ub − ug and velocity u̇b − u̇g of
the structure with respect to the ground. A way to visualize this is by assuming that ub = ug and
u̇b = u̇g , that is, that the structure motion (in the “interface” degrees of freedom) is exactly the
same as the ground motion. In this case, the structure moves together with the ground and hence
there are no interaction forces between the soil and the structure (fs = 0).
If the soil behaves linearly, the vector of interaction forces with the soil can be calculated by
means of equation 3.6. Therefore, equation 3.5 can then be used to model the dynamic behaviour
of the system. An important remark about this equation is its similarity to the basic equation of
motion for soil-structure interaction formulated by Wolf [43], using the assumption that the soil
with excavation behaves just as the soil without excavation (free-field motion).

Equation 3.5 is written in terms of absolute motions u. In some cases, the equation of motion
for earthquake loading is written in terms of the motions relative to the ground ur = u − ug . This
gives the advantage that the earthquake is then represented as an inertial force on each node Mur ,
(in a similar way as in equation (3.2)) eliminating the need of working with ground displacements
and velocities. This is much more convenient as displacements and velocities are usually not mea-
sured directly and have to be obtained through integration of acceleration records. However, for
soil-structure interaction problems this transformation is not that useful. Rewriting equation 3.5
in terms of relative displacements:
Mür + (C + Cs ) u̇r + (K + Ks ) ur = f − Müs − Cu̇s − Kus (3.8)
As it can be seen, if the equation of motion is rewritten in terms of relative displacements, the
ground displacements and velocities are still needed. Additionally, if the excitation in the interface
nodes is not the same for all of them (as it will be shown in next sections), this equation is not
valid anymore. This is why the formulation in terms of absolute displacements is used.

3.2 Soil Modeling


3.2.1 Equivalent Spring and Dashpot coefficients
As it was seen in Chapter 2, probably the two most common ways to model monopiles embedded
in soil for SSI are by using 3D continuum elements, or by using an equivalent system of springs and
14 3. FORMULATION OF MODEL FOR SEISMIC ANALYSIS

dashpots, which are usually frequency-dependent. For simplicity of implementation in the finite
element code and coupling with the existing aeroelastic code, this last approach is adopted in the
analysis.

Figure 3.3: Portion of embedded monopile subjected to a force.

In order to understand the concept of equivalent springs and dashpots, consider an infinitely
long vertical monopile embedded in soil. An infinitely small portion of the monopile is assumed to
be excited with an harmonic force (See figure 3.3). This force can be written as:

F (t) = F eiωt (3.9)


Assuming a single (traslational) degree of freedom for both the force and the response, then
the out of phase response of the monopile portion to this harmonic force can be written as:

u(t) = ueiωt (3.10)


Where F and u are complex amplitudes that include the phase difference between the excita-
tion and the response.
The impedance function between the force and the response, which can be interpreted as a
“complex dynamic stiffness” can be written as:

F (t)
k∗ = = k + iωc (3.11)
u(t)
Where k(ω) and c(ω) represent the equivalent distributed ”spring” and ”dashpot” coefficients,
which correspond to the in-phase and out-of-phase components of the response, respectively. Of
course, such coeffients need to be discretized over a certain length in order to be used for a finite
element model.
If u(t) is considered to be the relative displacement between the monopile and the ground (as
discussed in section 3.1) and f (t) is interpreted as the force exerted from the soil to the monopile
due to this motion, this principle can also be used for seismic SSI problems.
The main restriction of this approach is that it assumes that the force exerted by the soil to
a point of the monopile is independent to such force on any other point. This rather simplistic
assumption could make this model inaccurate when comparing to the real behavior of the soil.

3.2.2 Continuous vs. Discrete Modeling


By using the principle discussed in previous section, the embedded monopile can be modelled in
two different ways.

In the first one, the monopile is modelled entirely, with the distributed springs and dashpots
over its entire embedded length to represent the soil behavior. Therefore, the soil exerts a force
over the entire length of the monopile, which depends on the relative motion of the monopile with
respect to the ground. Consequently, it is required to know the earthquake motion continuously
over the length of the monopile.
3.2. SOIL MODELING 15

Figure 3.4: Continuous vs. Discrete modeling of soil-pile system.

On the other hand, the monopile can also be modeled discretely at the ground surface. In this
way, the entire foundation system, including the continuous springs and dashpots, is replaced by a
reduced, 1-node system with stiffness and damping matrices that represent the discrete “springs”
and “dashpots”.
Considering a SDOF system subjected to an harmonic force F (t) = |F |eiωt :

mü + cu̇ + ku = F (3.12)

Assuming a response of the form u(t) = |u|eiωt , the complex dynamic stiffness can be written
as:

F (t)
k∗ = = k − mω 2 + iωc (3.13)
u(t)

Comparing previous expression with equation 3.11, it can be seen that the equivalent ”spring”
of the system includes the stiffness and the inertial contribution of the original system, while the
equivalent ”dashpot” contains the damping contribution.
Under this same principle, the monopile system can be reduced by finding its complex dy-
namic stiffness for the mudline node. This is done by applying a unit harmonic load on each of
the mudline degrees of freedom, and then inverting the flexibility matrix composed by the com-
plex responses due to these unit harmonic loads. The real and imaginary parts of the resulting
frequency-dependent matrix (which includes also non-diagonal terms to account for coupling be-
tween the translational and rotational degrees of freedom) represent the ”stiffness” and ”damping”
of the condensed system, and can be interpreted as a set of springs and dashpots at the mudline
node. These spring and dashpot coefficients are also known as impedance functions, and have
been derived by different authors for a wide range of foundation systems.
An advantage of discretizing the foundation into a single node system is that the earthquake
motion only needs to be known and applied at the ground surface, instead of at all the depths of
the monopile. However, this also implies that the kinematic interaction between the soil and the
structure is not taken into account (due to the difference between the response of the pile and
the ground motion itself), and therefore the response can be assumed to be generated only by the
inertial interaction (from the mass of the superstructure).
16 3. FORMULATION OF MODEL FOR SEISMIC ANALYSIS

3.3 Ground Response Analysis


When modeling the monopile foundation as a continuous structure embedded in soil, it is not
correct to assume that the ground motion is only applied at a certain point (such as the pile tip).
Moreover, it is also incorrect to assume that each point of the monopile is subjected to the same
ground motion. Both of these assumptions can lead to unrealistic responses of the structure [38],
which is why it is preferred to perform free-field ground response analysis.
Ideally, a full ground response analysis would model the rupture mechanism and evaluate how
the waves propagate until reaching the site of interest. However, this mechanism is so complicated
and unpredictable that such approach would not be practical. This is why different simplified and
empirical 1D, 2D and 3D methods of ground response analysis have been developed in order to
approximate the ground motion given an existing earthquake record, and they mostly depend on
the soil properties and the frequency characteristics of the ground motion.
For this case, a One-Dimensional shear wave propagation model was chosen, due to the simplic-
ity to implement in the finite element code. This is the same model as the one used in the software
SHAKE91 [44], STRATA [45], and described in detail by Kramer [46]. The main assumptions of
such model are the following:

• The response of the soil deposit is predominantly caused by SH waves (shear waves polarized
in the horizontal plane) propagating vertically from the underlying bedrock.
• The boundaries between the different layers of the soil, the bedrock and the surface are
horizontal.
• The soil deposit and the bedrock extend infinitely in the horizontal direction.
• The soil behaves in a linear way.
• The soil material allows dissipation of energy (material damping).

Consider a soil deposit with different soil layers on top of an underlying bedrock, as shown in
figure 3.5. For layer n with thickness hn , the unit weight ρn , shear modulus Gn and damping ratio
ξ are known.
The displacement due to a shear wave propagating in the z-direction on viscous media can be
represented with the following wave equation:

∂2u ∂2u ∂3u


ρ 2
=G 2 +η 2 (3.14)
∂t ∂z ∂z ∂t
Where η is the viscosity, which can be related to the damping ratio and shear modulus as:
2Gξ
η= (3.15)
ω
The solution to equation 3.14 may be written as:
∗ ∗
u(z, t) = Aei(ωt+k z)
+ Bei(ωt−kω z) (3.16)

Where A and B correspond to the amplitudes of the incident and reflected waves, respectively, and
kω∗ is the complex wave number defined as:
ω
kω∗ = (3.17)
vs∗
Where vs∗ is the shear wave velocity, which is related to the complex shear modulus G∗ as:
s s
∗ G∗ G(1 + 2iξ)
vs = = (3.18)
ρ ρ
3.3. GROUND RESPONSE ANALYSIS 17

Figure 3.5: One Dimensional ground response analysis.

For this particular problem, the wave equation is defined for each layer m (from 1 to n) such that
m = 0 at the layer top and m = hm at the layer bottom.

∂ 2 um ∂ 2 um ∂ 3 um
ρm = Gm + η m (3.19)
∂t2 ∂z 2 ∂z 2 ∂t
The solution for each equation may then be written as:
∗ ∗
um (z, t) = Am ei(ωt+kω,m z) + Bm ei(ωt−kω,m z) (3.20)

Due to compatibility of displacements and stresses, for m = 1...n − 1 the following conditions must
hold:

um (zm = hm , t) = um+1 (zm+1 = 0, t) (3.21)


τm (zm = hm , t) = τm+1 (zm+1 = 0, t) (3.22)

Where τm = G∗m ∂u
∂zm .
m

An additional boundary condition is that at the ground surface the stresses must be equal to zero.

τ1 (z1 = 0, t) = ikω,1 G∗1 (A1 − B1 )eiωt → A1 = B1 (3.23)

From equations 3.21 and 3.22 the following recursive formulas can be derived:
1 ∗ ∗ 1 ∗ ∗
Am+1 = Am (1 + αm )eikω,m hm + Bm (1 − αm )e−ikω,m hm (3.24)
2 2
1 ∗ ∗ 1 ∗ ∗
Bm+1 = Am (1 − αm )e ikω,m hm
+ Bm (1 + αm )e−ikω,m hm (3.25)
2 2

Where αm is the complex impedance ratio defined as:


kω,m G∗m
αm = ∗ (3.26)
kω,m+1 G∗m+1
18 3. FORMULATION OF MODEL FOR SEISMIC ANALYSIS

Starting with the surface layer (m = 1), repeated use of the recursive formulas 3.24 and 3.25 lead
to the following expression that relates the amplitudes at the surface with those at each layer m:
Am = am (ω)A1 (3.27)
Bm = bm (ω)B1 (3.28)
The functions am (ω) and bm (ω) can simply be derived by assuming A1 = B1 = 1 (in accordance
to the boundary condition stated by 3.23).
The transfer function between the displacements at two layers m1 and m2 can be consequently
defined as:
um1 (ω) Am1 + Bm1
Tm1,m2 (ω) = = (3.29)
um2 (ω) Am2 + Bm2
This transfer function can be used to compute the response at any layer given the motion at any
other layer.
um1 (ω) = Tm1,m2 (ω)um2 (ω) (3.30)
Because üm = iω u̇m = −ω 2 um for an harmonic motion, the transfer function T (ω) can be applied
for displacements and velocities as well. As the input motion is usually taken from records in the
time domain, a Fast Fourier Transform is applied to obtain its amplitude spectrum.
One important issue has to do with the fact that the input ground motion is usually recorded
on rock at the ground surface, where the amplitudes of the upward and downward waves are
equal (An = Bn ), rather than on the rock at the base of the soil deposit. These motions are
known as outcropping motions, and can be taken into account by considering the change in the
boundary conditions in the transfer function itself. This is done by means of transforming the input

Figure 3.6: Explanation of outcropping bedrock motion

motion into a “within” motion, for which the previously found transfer function can be applied.
Consequently, the Transfer function for an outcropping bedrock motion is:
An + Bn Am + Bm Am + B m
Tm,n (ω) = = (3.31)
2An An + Bn 2An
With this transfer function, the motion at any layer m can be found from a recorded surface bedrock
motion.

3.4 Finite Element Modeling


In a general way, wind turbines consist on two substructures: the rotor-nacelle-assembly and the
support structure, which can usually be divided in two components: the foundation and the tower.
In a similar way, a finite element model of a wind turbine can also be divided into these three
parts.
In order to take into account the aeroelastic and geometrically nonlinear effects that are already
calculated by using the Siemens in-house code BHawC, only the embedded part of the monopile
will be modeled separately. This model, which also includes the soil, will be coupled with the
existing monopile (above mudline), tower and RNA models to provide a full model capable of
modeling earthquake loading including aeroelastic interaction.
3.4. FINITE ELEMENT MODELING 19

Figure 3.7: General components of a wind turbine.

Figure 3.8: Components of full finite element model.

3.4.1 Embedded Monopile Model


The monopile FE model was built using Matlab, by assuming the following:

• The monopile and the soil behave in a purely linear way, and its behaviour is described by
the equation 3.5.

• The ground motions (velocities and displacements) are applied on the system through equiv-
alent springs and dashpots.

• The ground motions to be applied on each equivalent spring and dashpot are calculated using
free field 1-Dimensional ground response analysis.

The monopile is modeled using one-dimensional beam elements that take into account axial,
shear and bending deformation. In order to do this, the beam element proposed by Przemieniecki
[47] is used. This element, as proven by Friedman[48], has a high accuracy even for a large
20 3. FORMULATION OF MODEL FOR SEISMIC ANALYSIS

element length L, and is shear-locking free (as the bending and shear terms tend to the ones of an
Euler-Bernoulli beam for long, slender beams).
The element shape functions for shear and bending degrees of freedom are derived from the
exact displacement field as described by Friedman [48]. Moreover, linear shape functions are used
for the axial and torsional degrees of freedom. The full expressions for the element degrees of
freedom, stiffness matrix and shape functions, as used in the FE code, can be found in Appendix A.

3.4.2 Continuous Springs and Dashpots modeling

Figure 3.9: Degrees of freedom of selected finite element.

In order to model the embedded part of the monopile, the soil has to be modeled by using a set
of distributed springs and dashpots over the of the pile (plus adding discrete springs and dashpots
at the base for the case of the caisson model). The procedure to include this in the FE code will be
presented below.
For the sake of simplifying the finite element calculation of the monopile, each one of the ele-
ments of the embedded length of the monopile is discretized into ”soil elements” of a considerably
small length with respect to the total element length. For each one of these soil elements, the
distributed springs and dashpots are replaced with discrete springs and dashpots by multiplying
the distributed coefficients by the soil element length l. These discrete springs and dashpots are
considered to be acting at the midpoint of each soil element.
As it was shown in Section 3.1, the interaction forces with the soil depend on the relative
motions of the monopile with respect to the ground Fs = f (u − ug ). Therefore, this motion is
calculated for each one of the points where the discrete springs and dashpots are located. This is
done by means of the beam element shape functions.
The displacement q at any point of one element going from node 1 to node 2, using a gen-
eralized coordinate s = x/Le (where Le is the lenght of the beam element) can be calculated
as:
 
u
q=N 1 (3.32)
u2

Where N is the shape function matrix, as described in Appendix A and u1,2 is the vector of degrees
of freedom at nodes 1 and 2.
The stiffness and damping matrices for each one of the discrete springs and dashpots is then
3.4. FINITE ELEMENT MODELING 21

calculated as:
   
0 0 0 0 0 0 0 0 0 0 0 0
0 kx l 0 0 0 0 0 cx l 0 0 0 0
   
0 0 kx l 0 0 0 0 0 cx l 0 0 0
k̃ = 
0
 c̃ =   (3.33)
 0 0 0 0 0 
0
 0 0 0 0 0 
0 0 0 0 kθ l 0 0 0 0 0 cθ l 0
0 0 0 0 0 kθ l 0 0 0 0 0 cθ l

Where kx and cx are the lateral spring and dashpot soil coefficients, and kθ and cθ are the
rotational spring and dashpot coefficients, calculated according to the corresponding soil models
indicated in Section 5.2. For the case of the flexible pile-soil model, as it will be explained in
Section 5.2, kθ and cθ are equal to zero.
Therefore, the forces exerted by the discrete spring and dashpot located at a point s are calcu-
lated as:
f̃ = k̃q + c̃q̇ (3.34)
The equivalent nodal forces of the element due to this point force are then:

f = NT f̃ (3.35)

Replacing equations 3.32 and 3.34 in 3.35, the following expression is obtained:

f = NT k̃Nũe (3.36)

From equation 3.36, it can be deducted that the equivalent element stiffness matrix due to the
contribution of a discrete spring and dashpot can be calculated as:

K = NT k̃N (3.37)

With equations 3.35 and 3.37 the element equivalent force and stiffness for each discrete spring
and dashpot element can be calculated. By summing up the contributions of each of the “soil
elements”, the equivalent soil force vector and soil stiffness matrix for each element can be found.
These are subsequently assempled to find the full monopile stifness and damping matrices, and
soil force vector. As the soil is assumed to behave linearly, the discrete springs and dashpots at the
base of the monopile can be added to the general monopile stiffness matrix once it is assembled.
One remark about this procedure is that, while the stiffness and damping matrices of the soil
Ks and Cs remain constant over time, the force vector is dependent on the ground motions,
which vary over time based on the ground motion time history and the ground response analysis.
Therefore, equation 3.35 would have to be applied for each time step. A more efficient approach
for this, taking advantage of the assumed linearity of the problem, will be explained here:

1. For each one of the discrete springs and dashpots, assume a unit displacement and velocity,
while setting the displacement and velocity of the rest of the springs and dashpots to zero.

2. Calculate the resulting nodal forces from each unit displacement and velocity using equation
3.35.

3. Assemble each resulting force vector as a column in the matrices Tc (for unit velocities) and
Tk (for unit displacements). Therefore, the matrices will have as many columns as discrete
springs are assumed in the model, and as many columns as the number of degrees of freedom
of the embedded monopile.

4. For each time step, the soil force vector can be calculated as:

fi = Tk ug,i + Tc u̇g,i
22 3. FORMULATION OF MODEL FOR SEISMIC ANALYSIS

3.4.3 Integration into Aeroelastic Code


The Siemens in-house code BHawC offers several features to analyze OWT models with different
kinds of foundations, ranging from monopiles to more complex foundations such as jackets. While
modeling a monopile is fairly simple, as it can be constructed with beam elements, jacket founda-
tions usually consist on many degrees of freedom, which would make the full OWT model too big.
This is why complex foundations are introduced into the BHawC code as superelements, which
are in essence a set of mass, stiffness and damping matrices that represent the complex foundation
structure with a reduced number of degrees freedom. These reduced matrices are usually obtained
using matrix reduction methods such as the Craig-Bampton or Guyan methods. These superele-
ments are accompanied by a set of wave load time series to be applied in each one of its reduced
degrees of freedom.
The superelement feature was used in order to include the linear embedded monopile model
into the aeroelastic calculation. Instead of using a set of reduced mass, stiffness and damping
matrices, the actual matrices of the embedded monopile foundation were included. The calculated
seismic loading time series were inserted as “wave loads”, distributed over the full length of the
embedded monopile.
Chapter 4

Selection and Adjustement of Ground Mo-


tions

4.1 Introduction to Japanese Seismicity


In order to be able to select a proper ground motion to apply to the model, a good starting point
is to understand the seismicity of the zone where the model is going to be used. In this case,
Siemens Wind Power major interest for this research is to become prepared for the Asian market,
particularly for its new developing projects in Japan. Therefore, the focus here is placed on the
seismicity of Japan.
Japan is situated in the collision zone of four major lithospheric plates: the Pacific plate, the
Phillipine plate, the Eurasian Plate and the North American Plate. Such plates form a so-called
subduction zone, which is an area where convergent boundaries (where one plate sinks under-
neath the mantle of the other plate) exist between the plates. In this movement, the plate with
more density (usually oceanic crust) bends and sinks into the mantle underneath the lighter plate
(usually continental crust). During this movement, energy is stored and then released as an earth-
quake. The very large scale of subduction zones is the reason why they are related to the largest
earthquakes in the world, mainly in the boundaries of the pacific ocean, commonly known as ”the
ring of fire” [49].

Figure 4.1: Map of plates and trenches in Japan. Source: GLGArcs [50].

23
24 4. SELECTION AND ADJUSTEMENT OF GROUND MOTIONS

Due to the subduction zone, many of the earthquakes that happen in Japan are very deep, and
are known as subduction-type eathquakes. However, shallow earthquakes can also occur, and they
tend to be stronger than deep focus earthquakes, as it was the case with the Tohoku Earthquake
in 2011 (30 km), the Kobe earthquake in 1995 (17 km) and the great Kanto earthquake in 1923
(23 km), which have been the largest and deadliest earthquakes that have hit Japan over the last
100 years, measuring a 7 on the Japan Meteorological Agency (JMA) seismic intensity scale. On
the other hand, as deep focus earthquakes generate minimal surface waves, they are not usually
likely to generate great damages. Additionally, recent studies have found that deep earthquakes
may be more efficient at dissipating energy (in terms of sub-event triggering) than shallow ones
[51], hence making them less damaging.

Figure 4.2: Hypocenters of earthquakes in Japan with magnitude MW greater than 5 for the period
1964-2011. Source: Earth Observatory of Singapore [52].

4.2 Selection Criteria


A common practice to select earthquake motions for using in time domain seismic analysis is by
using a target spectrum (usually, a defined 5 % damped acceleration response spectrum at the
location of interest). By doing this, the acceleration time record is transformed to the frequency
domain and then suitably modified to match this target acceleration response spectrum. Several
target spectra can be used for this, such as the Uniform Hazard Spectrum (UHS), the Conditional
Mean Spectrum (CMS) or the Conditional Spectrum (CS) [53]. A more simplistic approach, which
will be used here, is simply to use the Design Response Spectrum from the applicable seismic code
as the target spectrum for the derivation of different time histories.
As discussed in Chapter 2, Japan has very well explained recommendations for seismic analysis
of wind turbines, included in its Guidelines For Design of Wind Turbine Support Structures and
Foundations [54]. In this case, a simplified procedure based on these regulations is applied, where
existing ground motions following certain characteristics were matched to the design response
spectrum given by this design code. This response spectrum corresponds to a Level 2 Seismic
motion, which is defined as an event with a 500-year return period Tr , for a system with a damping
ratio ξ of 5 % of critical damping. This covers the requirements specified by IEC [5] and GL
[6]. By multiplying this spectrum by the maximum value of seismic zoning factor (Z = 1.0),
the maximum acceleration design spectrum for the country of Japan is generated, referred as the
“Target Spectrum” hereafter.
4.2. SELECTION CRITERIA 25

Target Response Spectrum


1

0.9

0.8

0.7

0.6

Sa [g]
0.5

0.4

0.3

0.2

0.1

0
0 0.5 1 1.5 2 2.5 3 3.5 4
Period [s]

Figure 4.3: Maximum design Response Spectrum for a level 2 seismic motion and a damping ratio of
5 %, as specified in the Japanese building code.

In order to find ground motions with a response spectrum that resembles the Target Spectrum,
the PEER NGA-West2 Database [55] is used. This database currently includes over 8000 corrected
three-component records of accelerations, displacements and velocities from over 300 shallow
crustal events from all around the world for moment magnitude scales MW ranging from 3.4 to
7.9. Also, it contains site information for the stations where these motions were recorded, which
is very useful when searching for ground motions for different purposes. Furthermore, a free
webtool is available which can automatically search and scale ground motions to match a certain
given Target Spectrum, providing a number of parameters to narrow the search results.
It is worth mentioning that most of the records of the NGA-West2 database were corrected us-
ing the PEER Record Processing Methodology which consists generally of a baseline correction and
a low-pass and high-pass acausal Butterworth filters applied in the frequency domain, for which
the corner frequencies were selected by visual examination of the Fourier amplitude spectrum of
accelerations. This methodology was compared by Boore et al. [56] to other processing method-
ologies, concluding that such post-processing did not affect the bulk of the results and hence the
PEER NGA records can be used with confidence.
The largest restriction of this database is the fact that it is limited to shallow crustal earthquakes
recorded in “active tectonic regimes”. Therefore, it does not include records from subduction and
deep earthquakes, or records in less active tectonic regimes. However, as shallow earthquakes
are usually more damaging for the the structures than deep-focus earthquakes, the contents of
NGA-West2 are still valid to use under the scope of this study.
The design spectrum as given in the Japanese code is for the engineering bedrock (not to
be confused with seismic bedrock), which is defined as a shallow interface of which underlying
stratum has a shear wave velocity of 300 to 700 m/s, according to Japan’s International Institute
of Seismology and Earthquake Engineering IISEE [57]. Figure 4.4 shows the difference between
the concepts of Seismic and Engineering bedrock. All of the records from the NGA-West2 database
were recorded at the surface, therefore, it would important to use only records on sites with soil
properties similar to the ones of the studied model. However, since a spectral matching algorithm
will be applied to the chosen ground motions, and such algorithm modifies the frequency content
of the record, the effect of soil amplification is ruled out, thus eliminating the need to only work
with stiff soil records. This is why no restrictions were chosen for the vs30 (average shear wave
velocity between 0 and 30 meters depth).
26 4. SELECTION AND ADJUSTEMENT OF GROUND MOTIONS

Figure 4.4: Graphical explanation of seismic and engineering bedrock. Source: IISEE [57].

On the other hand, the parameters Rjb (Horizontal distance to the surface projection of the
rupture plane) and Rrup (closest distance to the rupture plane) are selected in order to select only
records corresponding to near-field events (in this case, Rjb < 15 km). This was chosen in order
to reduce the number of parameters to investigate, although the differences in the responses due
to near-field and far-field motions can be significant, as it has been pointed out by several authors
[58, 33], and could be a matter of research for future studies in the subject.
Moreover, the rest of the search parameters are selected as follows:

Magnitude: (Also known as moment magnitude scale MW ) Since the level 2 seismic event corre-
sponds to ”extremely rare seismic motions”, a minimum magnitude of 6.0 is selected.

D5-95: (Effective duration of motion / Time interval between the 5% and the 95% Arias intensity
is reached.) A value of D595 between 15 and 25 seconds is selected to account for similar
records in the time domain, with a considerable duration with respect to the total time
analysis duration.

Fault: Being a subduction zone, the most common type of faults in Japan are reverse faults. There-
fore, only reverse-type faults are considered.

With these defined parameters, the response spectrum (for ξ = 5 %) of the SRSS of the two
horizontal components of several records is compared with the target spectrum. Each record is
amplified in order to minimize the Mean Square Error (MSE) by means of a constant weight
function (period-dependent weight functions are also possible, but were not used in this case). It
was also taken into account to select motions with not so large values of scale factors, in order
to not use an inaccurately amplified record, which could lead to unreal values of velocities and
displacements.
According to the described criteria, three records were selected, as shown in Table 4.1. For
each record, the corresponding RSN (Record Sequence Number) from the NGA-West2 database is
listed. For reference purposes, information about the correction performed to each one of these
records (for each one of the two horizontal components H1 and H2) is shown in Table 4.2.

RSN Event Mw Rjb [km] Vs30 [m/s] D5-95 [s] Scale Factor MSE
1495 Chi-Chi, Taiwan 7.62 6.34 359.13 26.8 0.9252 0.036
4847 Chuetso-oki, Japan 6.8 9.43 383.43 20.3 0.7785 0.1115
8134 Christchurch, NZ 6.2 11.25 247.5 14.5 1.4776 0.0563

Table 4.1: Selected ground motion records.

The response spectrum of the chosen events, compared to the target response spectrum, are
shown in figure 4.5.
4.2. SELECTION CRITERIA 27

RSN Event Component Type of Number of High-Pass Corner Low-Pass Corner Lowest Usable
Filter* Passes of Filter Frequency [Hz] Frequency [Hz] Frequency [Hz]
1495 Chi-Chi, Taiwan H1 A 1 0.02 50 0.250
H2 A 1 0.04 50 0.050
4847 Chuetso-oki, Japan H1 A 1 0.07 50 0.088
H2 A 1 0.06 50 0.075
8134 Christchurch, NZ H1 A 1 0.04 40 0.050
H2 A 1 0.02 45 0.025

* Type of Filter A: Acausal Butterworth filter

Table 4.2: Correction parameters for selected ground motion records.

Response Spectrum of Selected Records


SRSS
1.2

Target Spectra
Seed 1 - Christchurch
1 Seed 2 - Chi-Chi
Seed 3 - Chuetsu-Oku
Aritmetic Mean
0.8
Sa [g]

0.6

0.4

0.2

0
0 0.5 1 1.5 2 2.5 3 3.5 4
Period [s]

Figure 4.5: Chosen events SRSS response spectrum, compared with Target Spectrum.

4.2.1 Rotation of Time Records


As an input for time history analysis, strong motion records databases (such as the NGA West-
2) usually provide ground accelerations recorded in 3 orthogonal directions: two horizontal and
one vertical. The accelerometers are often (but not always) aligned to match the North-South
and East-West directions. Since the records depend on how the accelerometers is oriented, they
are commonly known as ”as-recorded” time histories, and linear transformations can be applied
on them to find the records as it would have been if the accelerometers had been oriented in a
different way.
It is a common practice to rotate the recorded ground motions in order to reach the maximum
possible response of the structure. Certain codes such as the ASCE [59] require that records that
will be used on earthquake time history analysis should be rotated in their fault-normal and fault-
paralel directions (for sites within 5 km of the active fault),as it is assumed that this approach will
lead to two sets of responses that envelope all the range of possible responses of the structure.
Formulas have been proposed to calculate the critical angle ϕcr on which the records should be
rotated in order to reach the maximum response [60, 61]. However, it has been proved [62] that
rotating earthquake records to either its Fault-Normal or Fault-Parallel directions or to the critical
angle ϕcr can guarantee the highest possible structural response.
This is why, for the academic purposes of this thesis, it was chosen to follow a simpler ap-
proach, initially proposed by Penzien amd Watabe [63], which is to rotate the ground motions to
28 4. SELECTION AND ADJUSTEMENT OF GROUND MOTIONS

its principal directions. Such directions are defined as the directions for which the two horizon-
tal components of the ground motion are mostly uncorrelated, this is, for which the covariance
between the two horizontal components is minimized.
In order to do this, the given acceleration record in both horizontal directions (ax and ay ) is
rotated to an angle ϕ using the following transformation equations:

arx,i = ax,i cos(ϕ) + ay,i sin(ϕ) (4.1)


ary,i = −ax,i sin(ϕ) + ay,i cos(ϕ) (4.2)

The covariance of the rotated record of a certain angle ϕ is then calculated as:
N
1 X r
cov(arx , ary ) = ax,i − a¯rx ary,i − a¯y r
 
(4.3)
N − 1 i=0

Where a¯rx and a¯ry are the mean values of the rotated records in both horizontal directions arx
and ary , respectively, and N is the number of points of each record.
Each pair of time histories (acceleration, velocities, displacements) was rotated to the direction
for which the calculated covariance of the acceleration record was minimized.

4.3 Spectral Matching


Usually, when performing time history seismic analysis, it is required by the building codes to
apply a certain number of earthquake records to the structure, such that the the mean response
spectrum of such events is close to the given target response spectrum. Due to the fact that response
spectrum from real earthquake events are not usually smooth but have large peaks and troughs, a
large number of time histories may be needed to get a mean response spectrum that approaches
the target spectrum with a certain accuracy.
One common engineering practice to deal with this is by using Spectral Matching Algorithms.
With this, the frequency content of the original time series is modified to match the target response
spectrum at a certain range of spectral periods. This brings the advantage that less records are
needed to get a mean spectrum that is closer to the Target Spectrum, hence reducing the number
of analysis needed.
Such approach is, nevertheless, still controversial, as the response spectrum of the matched
earthquake will end up resembling more a mean spectrum of several events than the one of a
single event (in the cases where a uniform hazard spectrum is used as a target spectrum). This
could lead to an overestimation of the the structural response in some cases [64]. However, with
the purpose of getting a good representation of the given Target Spectrum, without having to apply
more earthquake events and thus do more simulations, it was chosen to apply a spectral matching
algorithm to the time histories that were previously selected.
For this study, the spectral matching was done using the software SeismoMatch [65], which is
free for educational purposes. SeismoMatch uses the wavelets algorithm proposed by Abrahamson
[66] and Hancock et al.[67], which consists on iteratively adding wavelets to the acceleration time
series so that the corresponding response spectrum of such record matches a given target response
spectrum (on a certain period range) within a certain tolerance. One advantage of this algorithm
is that, while the acceleration time series is modified to match the target spectrum, the resulting
velocity and displacement time series are not corrupted with long-period drifts, avoiding the need
to use further baseline corrections after each iteration.
The 3 pairs of time histories were matched to the Target Spectrum over a range of periods
from 0.1 to 1.4 seconds. This range was chosen in order to get a better match towards the higher
modes, which are expected to be more relevant in the seismic response. For this case, the chosen
range would correspond to a frequency range of 0.7 Hz to 10 Hz. This criteria was chosen instead
of the more common criteria of matching for a range between 0.2Tn and 1.5Tn (where Tn is the
4.3. SPECTRAL MATCHING 29

natural period of the structure) which is adopted in building codes such as the ASCE standard [59],
due to the fact that the same earthquake events were going to be used for different models with
different natural periods. Additionally, this criteria does not consider the very unique behavior of
wind turbines with respect to its higher order modes.
After applying the spectral matching algorithm, a maximum error of 7.23% and 10.97% were
obtained for X and Y direction, respectively. The resulting response spectrum of the matched
records for each direction, compared to the initial Target Spectrum, are shown in Figure 4.6.

Response Spectrum of Matched Records Response Spectrum of Matched Records


Direction 0 Direction 90
1.2 1.2

Target Spectra Target Spectra


Seed 1 - Christchurch Seed 1 - Christchurch
1 Seed 2 - Chi-Chi 1 Seed 2 - Chi-Chi
Seed 3 - Chuetsu-Oku Seed 3 - Chuetsu-Oku
Aritmetic Mean Aritmetic Mean
0.8 0.8
Sa [g]

0.6 0.6

0.4 0.4

0.2 0.2

0 0
0 0.5 1 1.5 2 2.5 3 3.5 4 0 0.5 1 1.5 2 2.5 3 3.5 4
Period [s] Period [s]

Figure 4.6: Response Spectrum for both directions of selected records, after spectral matching.

For reference purposes, both the initial and the matched ground motions for the three selected
records are shown in figures 4.7, 4.8 and 4.9. For practical reasons, each one of the selected
records will be referred as a “seismic seed” (in a similar way as the “wave seeds” that are used for
the wave loads time series of OWT). The equivalence between each shown seismic event and its
”seismic seed” is shown in table 4.3.

Earthquake Record Seismic Seed


Christchurch 1
Chi-Chi 2
Chuetsu-Oku 3

Table 4.3: Equivalence between selected earthquake records and seismic seeds.

The FFT of the acceleration time history of each one of the selected seismic seeds is shown
in figure 4.10. It is interesting to notice that, despite the fact that the response spectrum of the
matched earthquakes is very similar, the time histories still vary significantly between earthquakes.
Factors such as the difference in the duration of the events (D5-95 varies in 10 seconds between
Seeds 1 and 3), how the energy is spread over this duration or the presence of velocity pulses
(more noticeable on Seismic Seed 2) can have an influence on the time histories of the earth-
quake acceleration and its corresponding structural responses, and therefore, should be taken into
account when analyzing the results.
30 4. SELECTION AND ADJUSTEMENT OF GROUND MOTIONS

Seed 1 - 0 Seed 1 - 90
4

3 Matched Matched
Unmatched Unmatched
2
Acceleration [m/s2 ]

-1

-2

-3

-4
0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
Time [s] Time [s]

Matched Matched
0.4 Unmatched Unmatched

0.2
Velocity [m/s]

-0.2

-0.4

-0.6
0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
Time [s] Time [s]

Matched Matched
0.2 Unmatched Unmatched
Displacement [m]

0.1

-0.1

-0.2

-0.3
0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
Time [s] Time [s]

Figure 4.7: Unmatched vs. Matched ground motion time series - Seismic Seed 1 (Christchurch
record).
4.3. SPECTRAL MATCHING 31

Seed 2 - 0 Seed 2 - 90
4

3 Matched Matched
Unmatched Unmatched
2
Acceleration [m/s2 ]

-1

-2

-3

-4
0 10 20 30 40 50 0 10 20 30 40 50
Time [s] Time [s]

Matched Matched
0.4 Unmatched Unmatched

0.2
Velocity [m/s]

-0.2

-0.4

-0.6
0 10 20 30 40 50 0 10 20 30 40 50
Time [s] Time [s]

Matched Matched
0.2 Unmatched Unmatched
Displacement [m]

0.1

-0.1

-0.2

-0.3
0 10 20 30 40 50 0 10 20 30 40 50
Time [s] Time [s]

Figure 4.8: Unmatched vs. Matched ground motion time series - Seismic Seed 2 (Chi-Chi record).
32 4. SELECTION AND ADJUSTEMENT OF GROUND MOTIONS

Seed 3 - 0 Seed 3 - 90
4

3 Matched Matched
Unmatched Unmatched
2
Acceleration [m/s2 ]

-1

-2

-3

-4
0 10 20 30 40 50 0 10 20 30 40 50
Time [s] Time [s]

Matched Matched
0.4 Unmatched Unmatched

0.2
Velocity [m/s]

-0.2

-0.4

-0.6
0 10 20 30 40 50 0 10 20 30 40 50
Time [s] Time [s]

Matched Matched
0.2 Unmatched Unmatched
Displacement [m]

0.1

-0.1

-0.2

-0.3
0 10 20 30 40 50 0 10 20 30 40 50
Time [s] Time [s]

Figure 4.9: Unmatched vs. Matched ground motion time series - Seismic Seed 3 (Chuetsu-Oku
record).
4.3. SPECTRAL MATCHING 33

×10 -3 Seed 1 - 0 Seed 1 - 90


4.5

3.5

3
PSD [g2 /Hz]

2.5

1.5

0.5

0 2 4 6 8 10 0 2 4 6 8 10
Frequency [Hz] Frequency [Hz]
×10 -3 Seed 2 - 0 Seed 2 - 90
4.5

3.5

3
PSD [g2 /Hz]

2.5

1.5

0.5

0 2 4 6 8 10 0 2 4 6 8 10
Frequency [Hz] Frequency [Hz]
×10 -3 Seed 3 - 0 Seed 3 - 90
4.5

3.5

3
PSD [g2 /Hz]

2.5

1.5

0.5

0 2 4 6 8 10 0 2 4 6 8 10
Frequency [Hz] Frequency [Hz]

Figure 4.10: Power Spectral Density of each Seismic Seed acceleration record.
34
Chapter 5

Selection of Soil Parameters

A large part of this study deals with the influence of soil-structure interaction on the seismic anal-
ysis of the OWT. Since the soil modeling has a large level of uncertainty in this case, a reference
model will be defined using certain base parameters. Afterwards, these parameters will be varied
to study the sensitivity on the dynamic response to such changes.

5.1 Soil Properties


In order to investigate on the differences of the dynamic response of the model for varying soil
stifness, three types of soil will be used: A stiff soil, a soft soil, and an intermediate (medium) soil.
The criterion to choose their properties was based on the soil classification reccomended by ASCE
[59], which is summarized below.

Class A: Hard Rock, vs >1500 m/s

Class B: Rock, 760 m/s <vs <1500 m/s

Class C: Very dense soil and soft rock, 360 m/s <vs <760 m/s

Class D: Stiff soil, 180 m/s <vs <360 m/s

Class E: Soft clay soil, vs <180 m/s

Since the engineering bedrock as defined in Section 4.2 falls within Class C, the soils to be
used in the analysis will correspond to Class D (for the stiff soil) and Class E (for the medium and
soft soil). These would correspond to soil types I and III respectively, according to the Japanese
Building Code [68].
By using tables for typical values of soil properties avaliable in the literature [69], the different
soil properties selected to represent the three different kinds of soil and the engineering bedrock
were selected, as shown in Table 5.1. Keeping in mind that the scope of the analysis is for offshore
turbines, a saturated condition is assumed in all cases. Moreover, a damping ratio ξ of 0.05 is
assumed for all soil materials. For simplicity of the parametric study, it is assumed that the soil
properties are constant from the surface until the depth of the engineering bedrock, which is set
as 50 meters. Therefore, a single soil column on top of the engineering bedrock was modeled.
Because of this, extra precaution was taken to not select a soil with very low stiffness conditions,
as having a very small stiffness for a depth of 50 meters would be very unrealistic to assume.
The stiff soil was selected as a starting point for the reference model.

35
36 5. SELECTION OF SOIL PARAMETERS

Number Type Es [MPa] ν ρ[kg/m3 ] G [MPa] vs [m/s] ξ


1 Hard Soil 300 0.2 2000 125 250 0.05
3 Medium Soil 150 0.25 1900 60 178 0.05
2 Soft Soil 80 0.3 1800 31 131 0.05
4 Engineering Bedrock 1200 0.3 2000 462 480 0.05

Table 5.1: Selected soil types for the analysis.

5.2 Soil Models


For this study, it was chosen to model the soil as a set of continuous springs and dashpots over the
length of the monopile. Therefore, adequate springs and dashpots coefficients must be chosen in
order to model the soil in an accurate way.

5.2.1 Pile (flexible) model


The main assumption for modeling the soil in this way, which is also the main assumption for
several models done by other authors who have tried to empirically predict the values of the
spring and dashpot coefficients, is the assumption of locally reacting points (this is, that the force
exerted by the soil on the point depends only on the motion of that point), which is linked to a
plane strain state. As explained in Section 3.2.1, for an infinitely long pile undergoing harmonic
vibration, the soil reaction at one point can be assumed independent of any other point. Therefore,
the soil is modeled as a Winkler foundation.
One of the first authors to propose the plane strain assumption was Novak [70], who found
the values of the soil dynamics stifnesses (the “spring” and “dashpot” coefficients) under different
kinds of harmonic vibrations. Novak’s model assumes that the waves propagate in all directions
on the 2D plane perpendicular to the monopile axis.
Furthermore, other authors have proposed different models to account for the problem of find-
ing these equivalent coefficients for an embedded monopile under the plane strain assumption.
One of the main differences between these models is how the wave propagation is idealized.
Gazetas and Dobry [39] assumed a much more simplified 1 dimensional propagation over the 2D
plane. This model proved to be in good agreement with more sophisticated 2-D and 3-D models.
For this study, the simplified distributed spring and dashpot coefficients proposed by Makris
and Gazetas [71] based on the simplified model developed by Gazetas and Dobry [39] were used
in the monopile model. Their values are the following:

kx = 1.2Es (5.1)
−1/4
cx = 6a0 ρs vs d (5.2)

Where Es is the soil Young’s modulus, vs is the soil shear wave velocity, ρs is the soil density, d is
the pile diameter, and a0 is the dimensionless frequency defined as:
ωd
a0 = (5.3)
vs
As the analysis is done in the time domain, the values of the damping coefficients (which are
dependent on the frequency) are averaged for a certain frequency range of interest (0.2 to 2.5 Hz).
Figure 5.1 shows the normalized values of these damping coefficients as a function of frequency
and its corresponding linearized value (which was used for the model). As it can be seen, the
chosen damping coefficient underestimates the soil damping for frequencies below approximately
1 Hz, and underestimates it for frequencies above such value.
It is worth mentioning that these damping coefficients, for the used pile diameter and selected
soil properties, underestimate the soil damping when compared to more elaborated soil damping
models as they do not take into account, for example, the dependency with the cylindrical section
5.2. SOIL MODELS 37

Stiff Soil
0.25
Makris and Gazetas
0.2

0.15
cx/Es

0.1

0.05

0
0.5 1 1.5 2 2.5
Frequency [Hz]
Soft Soil
0.25
Makris and Gazetas
0.2

0.15
cx/Es

0.1

0.05

0
0.5 1 1.5 2 2.5
Frequency [Hz]

Figure 5.1: Frequency dependent dashpot coefficient from pile model [71] and its linearization for
time domain analysis.

of the pile through Hankel Functions. As a reference, the used coefficients are compared in Figure
5.2 with the much more elaborated radiation damping coefficients proposed by Gazetas and Dobry
[72], which include this effect on its formulation.
One important remark about the mentioned coefficients is that they were derived for flexible
piles, which are defined as piles that don’t deform over their entire length L, but only up to certain
length defined as “active length” la . Gazetas and Dobry present an estimation of the active length
which depends on the soil stiffness, the pile stiffness, and the pile section. For the material of the
monopile of this study (steel), figure 5.3 shows the active length over diameter ratio (la /d) for
different soil stiffnesses according to this estimation.
As it can be seen, for the soil stiffnesses defined in section 5.1, the active length corresponds
to about 5-12 times the pile diameter. Hence, it is expected that the pile is longer than this active
length, to ensure a flexible behavior for which Gazetas and Dobry model is applicable. However,
the typical slenderness for an OWT with the dimensions of this study ranges between 4-6. This
means that the monopile might not be long enough to be considered flexible, especially for the
softer soils.
38 5. SELECTION OF SOIL PARAMETERS

Stiff Soil
0.3
Simplified Soil Model
0.25 Elaborated Soil Model

0.2
cx/Es

0.15

0.1

0.05

0
0.5 1 1.5 2 2.5
Frequency [Hz]
Soft Soil
0.3
Simplified Soil Model
0.25 Elaborated Soil Model

0.2
cx/Es

0.15

0.1

0.05

0
0.5 1 1.5 2 2.5
Frequency [Hz]

Figure 5.2: Comparison of dashpot coefficient between simple [71] and elaborated [72] model for
the frequency range of interest.

40

35

30

25
la/d

20

15

10

0
0.10 1.00 10.00 100.00 1000.00 10000.00
Es [Mpa]

Figure 5.3: Active length / Pile diameter for different soil stiffnesses.
5.2. SOIL MODELS 39

5.2.2 Caisson (rigid) Model


Because the monopile of this study cannot be considered as fully flexible, an additional soil model
will also be analysed. This model, presented by Varun et al. [73], was originally created to find
the response of large diameter caisson foundations. The main difference between this model and
the previous one, is that the caisson is assumed to move as a rigid body. Based on this, a 3-spring
and dashpot system is used to model the soil reaction:

• A distributed lateral spring to account for the soil lateral reaction.


• A distributed rotational spring to account for the vertical shear stresses along the caisson.
• A discrete lateral spring at the base to account for the soil base reaction of the caisson.

The coefficients for this model were obtained empirically by Varun et al. [73] from elaborated
continuum FE models.

5.2.3 Feasibility of Soil Models


As it was previously explained, two different models will be used to model the soil reactions in the
OWT monopile, as it is shown in Figure 5.4.

Figure 5.4: Depiction of selected soil-pile models. Left: Pile model by Makris and Gazetas [71]. Right:
Caisson model by Varun et al. [73].

In order to verify which model is more accurate for this study, the behavior of the monopile
will be analyzed by means of solving a static case of the Timoshenko beam differential equations.
The equations were written so that the distributed soil reaction is taken into account as a kind of
Winkler foundation. On the other hand, the soil reactions at the base of the monopile are taken
into account by means of the boundary conditions. A unit shear force (V = 1) was applied to the
top of the monopile as a static excitation. Figure 5.5 shows the schematics of this model.
The coefficients used to get the soil reactions were the same as the ones presented by Makris
and Gazetas [71] and Varun et al. [73], for the pile and caisson model, respectively. The chosen
pile section is hollow circular, with a thickness of 1% and made of high strength steel (E = 210 GPa,
ν = 0.3).
The resulting normalized pile displacements and rotations for both sets of spring coefficients
are shown in figures 5.6 and 5.7, for different ratios of pile and soil stiffness (E/Es ).
40 5. SELECTION OF SOIL PARAMETERS

Figure 5.5: Static Timoshenko beam model.

Displacements Rotations
0

0.1

0.2

0.3

0.4
z/L

0.5

0.6

0.7 E/E = 10
2
E/E = 10
2
s s
3 3
E/E = 10 E/E = 10
s s
0.8 E/E = 10
4
E/E = 10
4
s s
5 5
E/E = 10 E/E = 10
s s
0.9 6 6
E/E = 10 E/E = 10
s s

1
-0.5 0 0.5 1 -0.2 0 0.2 0.4 0.6 0.8 1
u/umax u'/u'max

Figure 5.6: Normalized displacements and rotations over pile length - Pile model.

When observing the curves corresponding to the ranges of soil stiffness defined for this study, it
can be concluded that the modeled monopile does not exhibit neither a flexible nor a rigid behavior.
Only for really soft soils, the pile’s rotation over the length varies insignificantly to consider the
pile moving as a rigid body. On the other hand, the pile deflections become negligible towards
the tip only on very stiff soils. For the soils in between these two extremes, a ”mixed” behavior is
observed.
Since none of the proposed soil models are proven to be fully accurate to describe the monopile
behaviour, it was decided to use the Pile (flexible) model by Makris and Gazetas [71] for most of
the simulations (including the reference model). This model was chosen over the caisson (rigid)
5.2. SOIL MODELS 41

Displacements Rotations
0

0.1

0.2

0.3

0.4
z/L

0.5

0.6

0.7 E/E = 10
2
E/E = 10
2
s s
3 3
E/E = 10 E/E = 10
s s
0.8 E/E = 10
4
E/E = 10
4
s s
5 5
E/E = 10 E/E = 10
s s
0.9 6 6
E/E = 10 E/E = 10
s s

1
-0.5 0 0.5 1 -0.2 0 0.2 0.4 0.6 0.8 1
u/umax u'/u'max

Figure 5.7: Normalized displacements and rotations over pile length - Caisson model.

model because assuming a monopile base reaction would mean that the soil at this region would
exhibit a non-linear behavior, which is out of the scope of this thesis. Nevertheless, the behavior of
the OWT with the caisson model will be studied as well, in order to get an insight on the influence
of the soil modeling in the results.

5.2.4 Sensitivity to Soil Coefficients


For this study, the two mentioned models were chosen to get the spring and dashpot coefficients.
However, several authors have proposed other different linear and nonlinear models to account
for the soil reactions. Such coefficients are usually proportional to the soil characteristics such as
stiffness or density, and hence the variation between them is usually of a small order of magnitude.
Since the earthquake loading is applied through these springs and dashpots, it is a matter of interest
to find out how can these small variations influence the overall turbine response.
The influence of the soil stiffness is already investigated indirectly by using the different selected
soil types defined in Section 5.1. Therefore, only the influence of the soil damping will be studied.
This will be done by running additional simulation with multiplier factors of 0, 0.5 and 2 applied
to the distributed dashpot coefficients. It is worth mentioning that the values of these multipliers
are much higher than the usual factors to which the soil dashpot coefficients vary.
42
Chapter 6

Preamble of Simulations

Before going further with the results of the simulations, some aspects of them that have not been
part of previous chapters will be discussed.

6.1 Wind Turbine Model


In order to mimic reality as closely as possible, one representative turbine model was used in the
analysis, based on Siemens current offshore offer for Asian markets.

SWT-4.0-130 (G4 Platform) This 4.0 MW Turbine, which is an upgraded version of the previous
3.6 MW OWT, is equipped with 63-meter-long rotor blades, which gives a swept area of 13.300
m2 . Aditionally, it uses aeroelastic blade technology which gives an increase of about 15 % in the
power production.

The hub height and monopile length of these kind of turbines are dependent on the site. For
this analysis, a model from a previous project will be used. Its dimensions, as used also in the FE
model, are shown in Figure 6.1.
The foundation and tower are consisted of circular steel hollow sections with variable properties
along the height. The section properties of some interest points of the support structure are shown
in Table 6.1.
Location Height* [m] Diameter [m] Thickness [m]
Pile Tip -34.55 5.90 0.06
Mudline 0.00 5.90 0.06
Interface 35.34 5.00 0.047
Tower Bottom** 35.34 5.00 0.035
Tower Top 101.53 3.08 0.023

* Height with respect to mudline.


* Interface location corresponds to section properties of the founda-
tion, Tower Bottom location corresponds to section properties of the
tower.

Table 6.1: Section properties on points of interest of support structure.

6.2 Operational States


One of the particularities of the dynamic behaviour of OWT with respect to regular buildings,
is that they they can be subjected to different operational states, depending, for example, if the

43
44 6. PREAMBLE OF SIMULATIONS

Figure 6.1: Turbine model to be used in the analysis.

turbine is producing energy or if it is parked for some reason. The dynamic response of a wind
turbine structure can greatly depend on these operational states, as the turbine is in constant
motion not only due to the rotation of the rotor, but also to the pitching and yawing that the
turbine is subjected to, which is dependent on the wind conditions.
In order to analyze and design a wind turbine, several operational states are taken into account.
However, for the academic purposes of this study, only the following operational states, based on
the Siemens Wind Department standard Design Load Cases (DLC) will be included:

• Normal Operation (DLC-1.2): This is considered to be the “regular” state of the OWT. The
wind speed is such that the turbine is producing energy under normal conditions. The normal
operation state is possible for wind speeds in a range from 4 to 32 m/s. The control system
of the turbine makes the blades pitch in order to reach an optimal power production while
also minimizing the loads imposed on the wind turbine.
• Parked/Idling State (DLC-7.2/DLC-6.4): In this state, the OWT is not producing any power,
as the blades are fully pitched and not rotating. The parked state can happen because either
the wind speed is to low or too high to operate normally, or because of a fault or planned
maintenance. The parked operational state can be simulated for any wind speed, but for
these simulations a range of 2 to 40 m/s was selected.

For means of comparison, all the simulations were done for both of these operational states,
and for both cases with and without the presence of seismic loading.

6.2.1 Control Systems of Normal Operation


The power production of a wind turbine greatly depends on the wind speed. For very low wind
speeds, the torque generated on the blades is insufficient to make them rotate, and thus no power
is generated. On the other hand, for very high wind speeds, the loads on the turbine can become
so high that they can cause structural damage in the turbine. This is why turbines include different
kinds of control systems so their behavior can be adjusted automatically depending on this.
A typical power of a wind turbine is shown in Figure 6.2. The turbine will start producing power
at the cut-in speed (4 m/s). The power output will increase steadily until reaching its maximum (in
6.2. OPERATIONAL STATES 45

Figure 6.2: Typical power curve of a wind tubine.

terms of the turbine capacity) at the rated speed (11-12 m/s). For speeds above this, the control
system limits the power output so it does not go below the turbine capacity (this is usually done
through pitching the blades). Finally, for wind speeds above the cut-out wind speed (25 m/s), the
forces of the turbine become too large, such that the increased power output does not compensate
the fatigue load accumulation of the turbine. This is in most cases, the control system shuts down
the turbine (leading it into an idling position) for this wind speed range.
However, Siemens Wind Turbines have an unique feature called Hight Wind Ride Through
(HWRT). This system gradually reduces the power output for wind speeds above 25 m/s instead
of shutting the turbine down completely. Therefore, the turbine can still produce power for wind
speeds up to 32 m/s, which is the actual turbine cut-out wind speed. This causes not only an
increase in the power production, but a decrease in the wear and tear of the different turbine
components due to the fact that less stops and starts occur over the lifetime of the turbine.
Figure 6.3 shows the actual power curve of the modeled turbine, including the effect of the
HWRT feature.

Figure 6.3: Power curve of modeled tubine, including HWRT feature.

The HWRT system is included in the BHawC model, and is activated due to the acceleration
of the rotor, which is usually an indication of high wind speeds or a sudden wind gust. However,
the high accelerations of the earthquake that are transmitted to the tower top can activate the
HWRT even for wind speeds below the cut-out speed. This is why, during the earthquake event,
the turbine will slow down, and after the seismic event has passed, it will slowly try to recover its
initial state of power production.
Additionally, the turbine is also equipped with an “overspeed monitor”, which will fully shut it
down if the velocity, as measured after the gearbox on the high-speed shaft, reaches a certain limit.
This velocity is both influenced by both the earthquake forces and the turbine’s RPM during power
production.
46 6. PREAMBLE OF SIMULATIONS

Both of these features greatly influence the behavior of the operating turbine during the earth-
quake event, and this must be taken into account when interpreting the results in the following
chapters. It might have been more practical to do the simulations without enabling these con-
trol systems, but it must be taken into account that having these control systems will make the
simulation more accurate in terms of the real behavior of the turbine.

6.3 Load Conditions


Usually, the analysis of an OWT takes into account two types of external loads: Wind and sea
waves. For this analysis, also seismic loads were included. A brief explanation of how each one of
these loads was taken into account in the model will be given below.

6.3.1 Wind Loads


As wind loads are usually design-driving for OWT, many different variations of its properties are
taken into account for a regular design. However, in an effort to minimize the number of simula-
tions and hence, the number of results to post-process and analyze, a very simple wind load state
was taken into account.
The chosen wind load parameters for this study were as follows:

• Wind Speed: The wind loads were calculated based on a mean wind speed for each simula-
tion. These wind speeds vary in steps of 2 m/s within the mentioned range for each operating
state (4-32 m/s for normal operation, 2-40 m/s for idling state).

• Wind Direction: An unique wind direction of 0 degrees (which means that the wind is fully
perpendicular to the plane of the rotor) was chosen.

• Wind Turbulence: Usually, the same wind speed is simulated by using different turbulence
seeds, in order to account for the random nature of the wind load. However, for this study
only one turbulence seed was used per simulation.

With these parameters, BHawC calculates automatically the wind load to be applied in the
turbine model.

6.3.2 Wave Loads


As it was explained in Chapter 3, seismic loads were included in the model as time series, using
the already existent feature of the model to include wave loads. Therefore, wave loads could not
be directly applied into the model.
It may have been possible to develop a way to include wave loads along with seismic loads.
However, due to the already unknown behavior of the turbine against earthquake events, it was
chosen to not include wave loads in the model and therefore minimize the amount of parameters
that could influence the results.
Nevertheless, the mass of the surrounding water on the foundation (above the mudline) was
taken into account by adding the corresponding terms to the mass matrix of this part of the model.

6.3.3 Seismic Loads


As it was mentioned before, the earthquake loads were applied as a time series of nodal forces and
moments that were calculated according to the procedure described in Chapter 3.
A total of three (3) pairs of seismic events (chosen according to Chapter 4) were used to
calculate such loads. Each pair was applied with rotations of 0 and 90 degrees.
6.4. REFERENCE MODEL AND PARAMETRIC STUDY 47

6.4 Reference Model and Parametric Study


As a result of the different load and operational states, a total of 364 simulations were run initially,
assuming that the turbine rests on a soil type 1 (stiff soil) and that the soil is modeled according
the pile (flexible)model, both described in Chapter 5. The set up for these simulations can be seen
in detail in the Load Case Table (Appendix B), used to define the loading state for each one of
the simulations on BHawC. The difference between each one of these simulations is due to the
differences in operational and environmental conditions, but not in the model itself. This first set
of simulations is what constitutes the results of the reference model, used in the results shown in
Chapter 8, which will answer the research sub-questions 1 and 2 as stated in section 1.4.
On the other hand, the parametric study which was done for the second part of the thesis
(Chapters 9 and 10) requires a variation on the stiffness and damping matrices of the soil. Since
these matrices are added to the respective matrices of the turbine structure to construct the whole
system to be analyzed in the time integration, a variation in the soil means that a whole new
set of simulations has to be run, as changes in the stiffness or damping of the system cannot be
considered as a variation of the loads.
Subsequently, the selection of soil parameters to be varied was chosen carefully in order to
minimize the number of new simulation sets to be run. According to this, the following additional
simulation sets were run as well:

1. Variation of soil types: 2 additional simulation sets (for soil types 2/medium and 3/soft).
2. Sensitivity to soil dashpot coefficients: 3 additional simulation sets (multipliers of 0, 0.5 and
2 for the soil dashpot coefficient, using soil type 1/stiff and the pile (flexible) soil model).

3. Sensitivity to selected soil-pile model: 1 additional simulation set (for caisson (rigid) soil
model, using soil type 1/stiff).
4. Sensitivity to discrete modeling of monopile: 1 additional simulation set (for the discrete
impedances model, as described in Chapter 3, using soil type 1/stiff and the pile (flexible)
soil model).

5. Variation of wave propagation model: 1 additional simulation set (assuming a Rayleigh wave
propagation ground model, using soil type 1/stiff and the pile (flexible) soil model).
6. Variation of seismic input: 1 additional simulation set (with a new set of seismic seeds, using
soil type 1/stiff and the pile (flexible) soil model).

The analysis of the results of the parametric study will be used to answer the research sub-
questions 3 and 4.
48
Chapters 7 to 9 are omitted for confidentiality reasons.

49
50
Chapter 10

Final Conclusions and Recommendations

In this chapter, some more general conclusions, regarding the most relevant findings of the whole
thesis will be discussed. Moreover, the research question and sub-questions, defined at Section
1.4, will be answered. Finally, some recommendations for future research on the subject of seismic
analysis of OWT will be shown.

10.1 General Conclusions


• The operational states of an OWT can significantly affect its response to a seismic event, par-
ticularly for the top structures of the turbine (tower top and blades), where it was found that
the extreme loads for each direction resulted from a different operational state (idling state
for fore-aft direction and normal operation for side-side direction of the tower). Therefore,
it cannot be said that one state is more critical than the other in case of seismic loading.
Moreover, an accurate seismic analysis of an OWT should include both operational states.
• The response of an OWT to a seismic event can be greatly influenced by its higher order
modes, which include both tower and blade modes. The effect of these higher order modes
can even be observed at the lowest parts of the support structure. Therefore, blades cannot
be neglected in the analysis, even if the analysis is aimed for a tower or foundation design.
• The ultimate loads of an OWT are very dependent on the applied earthquake events. This
is why they should be chosen with great care, depending on the site conditions. Moreover,
selecting these events based solely on a Target Response Spectrum could not be sufficient to
ensure realistic results, as the response also depends on the characteristics of the earthquake
event in the time domain. Moreover, the response spectrum is based on the response of a
single degree of freedom system, which is a behavior that can be extended to many regular
civil structures, but might not be valid for the rather complex dynamic behavior of a wind
turbine.
• Due to the stochastic nature of earthquakes, there is always a big uncertainty in applying time
series of previous earthquake events on structures, specially regarding the extreme values of
its responses. This is particularly true for the case of a wind turbine due to the complexity
of the dynamic response. Unlike wind and wave loads, where the randomness of the load
is accounted for by applying statistical methods to the measurement data in order to obtain
the loads to be applied to the model, designing with the maximum extreme load obtained
from real earthquake time histories could lead to conservative results due to modeling a very
particular scenario of an event with an already large return period.
• Earthquakes can simultaneously excite a wide range of frequencies, which follows their nat-
urally random behavior. This is why OWT cannot be designed to avoid earthquake frequen-

51
52 10. FINAL CONCLUSIONS AND RECOMMENDATIONS

cies, like it is done with wave loads or with the distinct rotational frequency of the rotor
(“p-excitation”). However, it was found that seismic loading seems to consistently excite cer-
tain particular turbine modes, in a way that can be less or more severe depending on the
frequency content of the earthquake itself. This could serve as a basis for the OWT seismic
design.
• Wind turbines are very complex structures whose behavior and dynamic properties vary over
time and are linked to its own operation and external factors such as the wind speed. When
combining this with loads as uncertain as seismic loads, it could be concluded that time do-
main analysis is the only option if one would like to account for aspects such as aerodynamic
damping, control systems and soil non-linearities. However, given the fact that the modeled
turbine presented a somehow defined and constant modal and frequency response, it could
be feasible to perform simplified frequency domain seismic analysis and obtain relatively ac-
curate loads with respect to the time domain. In order to do this, the operational states and
the aerodynamic damping derived from them should be taken into account accordingly, as
they have been shown to significantly influence the response.
• Soils with low stiffness can significantly increase the design loads of OWT, in case of earth-
quake loading, due to amplification of the response in the frequency range for which the
turbine gets more excited during the seismic event. In this case, the chosen soft soil lead to
an increase of the loads of up to 50 % when compared to the stiff soil. However, a simplified
constant soil profile was assumed, which could have influenced this result. This is why a
proper identification of the soil profile is needed in order to obtain accurate seismic loads.
• The response of an OWT model against seismic loading is very dependent on the way the
soil is modeled. This is why great attention should be paid into selecting an accurate soil-
structure interaction model that fully captures the soil behavior for the frequency range of
interest.
• Soil damping is very important for seismic analysis of OWT and should be taken into ac-
count, as not including it could lead to largely unrealistic loads. Linearizing the frequency-
dependent soil damping towards a certain range of frequencies (in order to be used in a time
domain analysis) could not have a significantly large impact in the design loads.
• Discretizing the monopile foundation into a single node system of coupled springs and dash-
pots could lead to fairly accurate design loads (at least, for a linear soil case). This means
that it would be possible to model the soil-foundation system separately from the OWT with-
out a significant loss of accuracy, as long as the correct loads are transferred from one model
to the other.

10.2 Answers to Research Questions


10.2.1 Sub-Questions
1. What is the influence of the operational state on the seismic response of an OWT?
Answer: As the operational states vary significantly the dynamic properties of the OWT, they
also influence the response particularly in the case of higher order modes. This leads to
significant differences in the extreme loads for both states (specially at the upper part of the
turbine).
2. Which is the most critical part of an OWT when seismic events occur?
Answer: For the modeled turbine, it was found that the moments at tower top and the blades
were closer to the turbine design loads, which make them the most critical locations. How-
ever, different factors regarding the modeling and the loading could make other locations
become critical as well.
10.3. RECOMMENDATIONS 53

3. How do the soil properties influence the dynamic response of an OWT?


Answer: In this case, softer soils lead to large increases of the turbine response. However,
this is dependent on the soil profile, therefore this might not be the case for other soil profiles
where the stiffness is distributed in a non uniform way over the height of the foundation.

4. How sensitive is the dynamic response of an OWT on the way the soil is modeled?
Answer: It was found that the response of modeled OWT was very sensitive to the choice
of soil model, which is why an accurate representation of the soil should be done to obtain
accurate seismic loads.

10.2.2 Main Question


Can earthquake loading become design driving for Offshore Wind Turbines?

Answer: Yes. Earthquake loading can excite an OWT in a very different way as it is excited with
waves or wind loading, therefore, it could affect the design for locations where large seismic events
are expected. However, according to the developed model, such seismic loads are very sensitive to
the way the soil is modeled, which is why an accurate representation of the soil should be done in
order to not under or over estimate the loads.

10.3 Recommendations
Since the subject of seismic analysis of wind turbine is relatively new and there is not a lot of
knowledge available regarding this, the present thesis was intended to simplify many aspects of
the modeling and the analysis order to get a general idea on the OWT behavior due to seismic
loading and how could it influence the design. As a following step, more detailed analysis should
follow, taking care of the aspects which were found to be more critical during the past chapters.
The author would like to draw the following recommendations for further research in this
matter:

• Include earthquake loading in the OWT turbine design for sites with large seismic activity, at
least for idling and normal operation states (DLC-1.2, DLC-7.2 and DLC-6.4).

• Investigate on the effect of wave loading combined with earthquake loading on an OWT.

• Investigate on the effect of the vertical component of the earthquake event on the OWT
seismic response.

• Investigate on the differences between time-domain and frequency-domain simulations (us-


ing an accurate representation of the OWT that includes the variations in eigenfrequencies
and modal damping ratios due to wind speed and operational state).

• Investigate on the sensitivity of the seismic response with blade modeling, using, for example,
matrix reduction methods to include the blade modes in a more simplified analysis.

• Develop a more accurate soil model that takes into account non-local soil effects, soil nonlin-
earities and soil damping, and investigate on the differences of the seismic response between
such model and a simplified spring and dashpot model with coefficients ”tuned” to this de-
tailed model.

• Develop a more strict criterion of selection of seismic events in accordance to the location of
the OWT.
54 10. FINAL CONCLUSIONS AND RECOMMENDATIONS

• When performing seismic analysis in the time domain, apply a large number of seismic seeds
selected using the developed criterion, and find the ultimate loads using statistical methods
(such as finding the mean value of the extreme loads of each seismic seed, or the expected
value of the peaks for a certain return period).
Appendix A

Element properties for Finite Element Model

A.1 Degrees of freedom


 T
ue = ux,1 uy,1 uz,1 θx,1 θy,1 θz,1 ux,2 uy,2 uz,2 θx,2 θy,2 θz,2

Figure A.1: Degrees of freedom of finite element.

A.2 Stiffness Matrix


 
Ke1,1 Ke1,2
Ke = (A.1)
Ke2,1 Ke2,2
Where:
 EA
− EA

Le 0 0 0 0 0 Le
12EIz 6EIz

 0 L3e (1+Φy ) 0 0 0 L3e (1+Φy ) 0 
12EIy 6EIy
0 0 0 − L3 (1+Φ 0 0 
 
 L3e (1+Φz ) z)
Ke1,1 = GJ
e 

 0 0 0 Le 0 0 0 
6EIy (4+Φz )6EIy

 0 0 − L3 (1+Φ z)
0 Le (1+Φz ) 0 0 
e
6EIz (4+Φy )6EIz
0 L3e (1+Φy ) 0 0 0 Le (1+Φy ) 0

55
56 A. ELEMENT PROPERTIES FOR FINITE ELEMENT MODEL

 EA
− Le

0 0 0 0 0
 0 − L312EIz
(1+Φy ) 0 0 0 6EIz
L3e (1+Φy )

 e 
12EI 6EIy
 0 0 − L3 (1+Φyz ) 0 − L3 (1+Φ 0
 
z)
Ke1,2 = Ke2,1 T

= e e
− GJ

 0 0 0 0 0 
 Le 
 0 6EIy (2−Φz )EIy
 0 L3e (1+Φz ) 0 Le (1+Φz ) 0 

(2−Φy )EIz
0 − L36EIz
(1+Φy ) 0 0 0 Le (1+Φy )
e

 EA 
Le 0 0 0 0 0
12EIz
 0 L3e (1+Φy ) 0 0 0 − L36EI z 
 e (1+Φy ) 
12EIy 6EIy
0 0 0 0
 
 L3e (1+Φz ) L3e (1+Φz ) 
Ke2,2 = GJ


 0 0 0 Le 0 0 

6EIy (4+Φz )EIy

 0 0 L3e (1+Φz ) 0 Le (1+Φz ) 0 

(4+Φy )EIz
0 − L36EI z
0 0 0
e (1+Φy ) Le (1+Φy )

Where:
12EIz
Φy = GAsz L2e
12EIy
Φz = GAsy L2e
A = cross-sectional area
E = Young’s modulus
Le = element length
G = shear modulus
J = torsional moment of inertia
Ii = moment of inertia normal to direction i
Asi = shear area normal to direction i

A.3 Shape Functions


 
N1 0 0 0 0 0 N2 0 0 0 0 0
0
 N1y 0 0 0 N2y 0 N3y 0 0 0 N4y 

0 0 N1z 0 N2z 0 0 0 N3z 0 N4z 0 
N=
0

 0 0 N1 0 0 0 0 0 N2 0 0 
0 0 N5z 0 N6z 0 0 0 N7z 0 N8z 0 
0 N5y 0 0 0 N6y 0 N7y 0 0 0 N8y
Where:

N1 = 1 − s
N2 = s
1  3
N1i = 2s − 3s2 − Φi s + (1 + Φi )

1 + Φi
1  3
N2i = − 2s − 3s2 − Φi s

1 + Φi
     
L Φ 2 Φ
N3i = s3 − 2 + s + 1+ s
1 + Φi 2 2
   
L Φ 2 Φ
N4i = s3 − 1 − s − s
1 + Φi 2 2
A.3. SHAPE FUNCTIONS 57

6
N5i =
 2 
s −s
(1 + Φi )L
6
N6i = −
 2 
s −s
(1 + Φi )L
1
N7i =
 2 
3s − (4 + Φi )s + (1 + Φi )
(1 + Φi )
1
N8i =
 2 
3s − (2 − Φi )s)
(1 + Φi )
58
Appendix B

BHawC Load Case Table

59
60 B. BHAWC LOAD CASE TABLE

Load Case Table


LCT revision # 6 Load Case Table
Case id Rev1 Adapted for Seismic sensitivity study
Design Standard IEC_ed3 All wave parameters, refer to the seismic events
Site Type Offshore
Turbine Type 4.0-120

Page 1/4

Design Design Load ACTIVE / Yaw errors Wind directions Wind wave
Description Options Wind range [m/s]
situation Case INACTIVE [deg] [deg] misalignments [deg]

BSH 35h Storm BSH35i Special_EWM-ESS_Idling INACTIVE 35, 48 -8, 0, 8 (0 : 30 : 330) 0

BSH35i_-EWH Special_EWM-ESS_Idling INACTIVE 48 -8, 0, 8 (0 : 30 : 330) 0

BSH35p Special_EWM-ESS INACTIVE 25 -8, 0, 8 (0 : 30 : 330) 0

Power production DLC11 INACTIVE (4 : 32) -5.6, 0, 5.6 0 0

DLC12 NTM_Seismic m0500_NoWave ACTIVE (4 : 2 : 32) 0 0 0

DLC12 NTM_Seismic m0500 ACTIVE (4 : 2 : 32) 0 0 0, 90

DLC12 NTM_NSS m1495 INACTIVE (4 : 32) -5.6, 0, 5.6 (0 : 30 : 330) 0

DLC13 ETM_NSS INACTIVE (4 : 32) -5.6, 0, 5.6 0 0

DLC14 ECD_NSS INACTIVE (9 : 15) 0 0 0

DLC15 EWS_NSS INACTIVE 4, (9 : 15), 25 -5.6, 0, 5.6 0 0

DLC16 NTM_SSS-SWH_50y INACTIVE (4 : 32) -5.6, 0, 5.6 (0 : 30 : 330) 0


Parked (standing
DLC61 EWM_ESS-EWH_50y_Idling INACTIVE 50 -8, 0, 8 (0 : 30 : 330) -30, 0, 30, 90
still or idling)
DLC62 EWM_ESS-EWH_50y_Idling_NoGrid INACTIVE 50 (5 : 10 : 355) (0 : 30 : 330) -30, 0, 30, 90

DLC63 EWM_ESS-EWH_1y_Idling INACTIVE 40 -20, 20 0 -30, 0, 30, 90

DLC64_h NTM_Idling_Seismic m0500 ACTIVE (34 : 2 : 40) 0 0 0, 90

DLC64_h NTM_Idling m1495 INACTIVE 33, 34, 35 -5.6, 0, 5.6 (0 : 30 : 330) 0

DLC64_l NTM_Idling_Seismic m0500 ACTIVE 2 0 0 0, 90

DLC64_l NTM_Idling m1495 INACTIVE 1, 2, 3 -5.6, 0, 5.6 (0 : 30 : 330) 0


Parked and fault
DLC71 EWM_ESS-EWH_1y_IdlingFault INACTIVE 40 -8, 0, 8 0 -30, 0, 30, 90
conditions
DLC72 NTM_IdlingNoGrid_Seismic m0500_NoWave ACTIVE (2 : 2 : 40) 0 0 0

DLC72 NTM_IdlingNoGrid_Seismic m0500 ACTIVE (2 : 2 : 40) 0 0 0, 90

DLC72 NTM_IdlingNoGrid m1495 INACTIVE (1 : 2 : 35) 0 (0 : 30 : 330) (-60 : 30 : 90)


Transport,
DLC81_a NTM_RotorAndYawLock INACTIVE 23 (60 : 20 : 300) 0 0
assembly,
DLC81_b NTM_RotorAndYawLock INACTIVE 18 (-40 : 20 : 40) 0 0

DLC83 NTM_NoGridDuringInstall m0500 INACTIVE (1 : 2 : 35) 0 (0 : 30 : 330) (-60 : 30 : 90)

DLC83 NTM_NoGridDuringInstall m1495 INACTIVE (1 : 2 : 35) 0 (0 : 30 : 330) (-60 : 30 : 90)

Internal DLC96 NTM_FrontalPassage m0500 INACTIVE 4, 10 0 0 0

DLC96 NTM_FrontalPassage m1495 INACTIVE 4, 10 0 0 0


Sea ice - Power
DLCE3 NTM_SeaIce INACTIVE (9 : 15), 32 -5.6, 0, 5.6 0 0
production
DLCE4 NTM_SeaIce m0500 INACTIVE (4 : 32) -5.6, 0, 5.6 0 0

DLCE4 NTM_SeaIce m1495 INACTIVE (4 : 32) -5.6, 0, 5.6 0 0

Sea ice - Parked DLCE6 EWM_SeaIce_Idling INACTIVE 40 -5.6, 0, 5.6 0 0

DLCE7 NTM_SeaIce_Idling m0500 INACTIVE (1 : 2 : 35) -5.6, 0, 5.6 0 0

DLCE7 NTM_SeaIce_Idling m1495 INACTIVE (1 : 2 : 35) -5.6, 0, 5.6 0 0

Prerun simulations PrepStartup EOG_EDC_Timings NoWave INACTIVE 4, (9 : 15), 20 0 0 0


Internal - Ver_BladeCrossS
INACTIVE
Verification ectionalPlot
Ver_BladeEigenV
ACTIVE
alues
Ver_Controller INACTIVE
Ver_FoundEigen
ACTIVE
Values
Ver_ModalAnaly
ACTIVE
sis
Ver_PowerCurve
INACTIVE (4 : 32)
NoTurb
Ver_PowerCurve
INACTIVE (4 : 0.5 : 32)
Turb
Ver_Status ACTIVE
Ver_TowerFound
ACTIVE
EigenValues
Ver_TurbineEige
ACTIVE
nValues
61

Load Case Table


LCT revision # 6
Case id
Design Standard IEC_ed3
Site Type Offshore
Turbine Type 4.0-120

Page 2/4

Apply number of Additional


Turbulence seeds Simulation
Design Design Load seeds multiplier Description of
per Description of conditions duration # Simulations
situation Case per yaw error? per additional multiplier
unique simulation [s]
[Y/N] unique simulation

BSH 35h Storm BSH35i 2 Y 600 144

BSH35i_-EWH 2 Y 600 72

BSH35p 2 Y 600 72
Load extrapolation based
Power production DLC11 29
on DLC12 with aligned
DLC12 3 120 45

DLC12 3 120 90

DLC12 3 600 1044

DLC13 9 600 261


Positive and negative
DLC14 2 60 14
direction change
Combinations of
DLC15 Y 4 60 108
horizontal or vertical and
DLC16 6 600 2088
Parked (standing
DLC61 6 Y 600 864
still or idling)
DLC62 2 Y Loss of electrical network 600 3456

DLC63 6 Y 600 48

DLC64_h 3 120 24

DLC64_h 3 600 108

DLC64_l 3 120 6

DLC64_l 3 600 108


Parked and fault
DLC71 12 Failure in one air brake 600 48
conditions
DLC72 3 Y Loss of electrical network 120 60

DLC72 3 Y Loss of electrical network 120 120

DLC72 1 Y Loss of electrical network 600 1296


Transport,
DLC81_a 6 Y Rotor and yaw lock 600 78
assembly,
DLC81_b 6 Y Rotor and yaw lock 600 30
No grid during
DLC83 1 Y 600 1296
installation period
No grid during
DLC83 1 Y 600 1296
installation period
Internal DLC96 120 2

DLC96 120 2
Sea ice - Power Horizontal load from
DLCE3 6 600 48
production moving ice floe at
Horizontal load from
DLCE4 6 600 174
moving ice floe at
Horizontal load from
DLCE4 6 600 174
moving ice floe at
Pressure from
Sea ice - Parked DLCE6 6 600 6
hummocked ice and ice
Horizontal load from
DLCE7 6 600 108
moving ice floe at
Horizontal load from
DLCE7 6 600 108
moving ice floe at
Prerun simulations PrepStartup 600 9
Internal - Ver_BladeCrossS
1
Verification ectionalPlot
Ver_BladeEigenV
2 Damping + No damping 2
alues
Various wind speeds and gust
Ver_Controller 9 9
types
Ver_FoundEigen Isolated foundation (Damping +
2 2
Values No Damping)
Ver_ModalAnaly
2 Wind + RPM modal analysis 2
sis
Ver_PowerCurve
29
NoTurb
Ver_PowerCurve
10 10
Turb
Ver_Status 1
Ver_TowerFound (Without RNA + Brake off + Idling
6 6
EigenValues Pos.) with (Damping + No
Ver_TurbineEige (All + Idling Pos + Clamped) with
6 6
nValues (Damping + No Damping)
62 B. BHAWC LOAD CASE TABLE

Load Case Table


LCT revision # 6
Case id
Design Standard IEC_ed3
Site Type Offshore
Turbine Type 4.0-120
ULS post processing setup
ULS_All_SupportStruct
Page 3/4 ULS_All_RNA
ure
INACTIVE ACTIVE

Design Design Load Sea current Relevant for RNA / Partial safety
Wind condition Sea state Water level Analysis type Averaging scheme Averaging scheme
situation Case model Support structure factor [-]

BSH 35h Storm BSH35i EWM ESS ECM EWLR extreme Soil

BSH35i_-EWH EWM ESS EWH50 ECM EWLR extreme Soil

BSH35p EWM ESS ECM EWLR extreme Soil

Power production DLC11 NTM I_eff NSS - >= MSL extreme R 1.25 max

DLC12 NTM I_eff - - - fatigue R/S 1 max

DLC12 NTM I_eff NSS - >= MSL fatigue R/S 1 max

DLC12 NTM I_eff NSS - >= MSL fatigue R 1.35mean_misalignment|windDir.windSpd

DLC13 ETM NSS NCM MSL extreme R/S 1.35 mean_windDir|windSpd


mean_windDir|windSpd

DLC14 ECD NSS NCM MSL extreme R/S 1.35 max max

DLC15 EWS NSS NCM MSL extreme R/S 1.35 max max

DLC16 NTM I'_char SSS50 SWH50 NCM NWLR extreme R/S 1.35 mean_windDir.windSpdmean_windDir.windSpd
Parked (standing
DLC61 EWM ESS50 EWH50 ECM EWLR extreme R/S 1.35mean_misalignment|windDir.yawErr
mean_misalignment|windDir.yawErr
still or idling)
DLC62 EWM ESS50 EWH50 ECM EWLR extreme R/S 1.1meanrun_misalignment|windDir.yawErr
meanrun_misalignment|windDir.yawErr

DLC63 EWM ESS1 EWH1 ECM NWLR extreme R/S 1.35mean_misalignment|windDir.yawErr


mean_misalignment|windDir.yawErr

DLC64_h NTM I_eff NSS - >= MSL fatigue R/S 1 max

DLC64_h NTM I_eff NSS - >= MSL fatigue R 1.35 max

DLC64_l NTM I_eff NSS - >= MSL fatigue R/S 1 max

DLC64_l NTM I_eff NSS - >= MSL fatigue R 1.35 max


Parked and fault
DLC71 EWM ESS1 EWH1 ECM NWLR extreme R/S 1.1
meanhalf_misalignment|windDir.windSpd
meanhalf_misalignment|windDir.windSpd
conditions
DLC72 NTM I_eff - - - fatigue R/S 1 max

DLC72 NTM I_eff NSS - >= MSL fatigue R/S 1 max

DLC72 NTM I_eff NSS - >= MSL fatigue R 1.35 max


Transport,
DLC81_a NTM I'_char NSS NCM MSL extreme R/S 1.5 mean_azimuth|yawErrmean_azimuth|yawErr
assembly,
DLC81_b NTM I'_char NSS NCM MSL extreme R/S 1.5 mean_azimuth|yawErrmean_azimuth|yawErr

DLC83 NTM I_eff NSS - >= MSL fatigue R/S 1.35 max max

DLC83 NTM I_eff NSS - >= MSL fatigue R 1.35 max

Internal DLC96 NTM I_eff NSS NCM MSL fatigue R 1.35 max max

DLC96 NTM I_eff NSS NCM MSL fatigue R 1.35 max


Sea ice - Power
DLCE3 NTM I'_char - - NWLR extreme R/S 1.35 max max
production
DLCE4 NTM I_eff - - NWLR fatigue R/S 1.35 max max

DLCE4 NTM I_eff - - NWLR fatigue R 1.35 max

Sea ice - Parked DLCE6 EWM - - NWLR extreme R/S 1.35 max max

DLCE7 NTM I_eff - - NWLR fatigue R/S 1.35 max max

DLCE7 NTM I_eff - - NWLR fatigue R 1.35 max

Prerun simulations PrepStartup NWP - - -


Internal - Ver_BladeCrossS
Verification ectionalPlot
Ver_BladeEigenV
alues
Ver_Controller
Ver_FoundEigen
Values
Ver_ModalAnaly
sis
Ver_PowerCurve
NoTurb
Ver_PowerCurve
Turb
Ver_Status
Ver_TowerFound
EigenValues
Ver_TurbineEige
nValues
63

Load Case Table


LCT revision # 6
Case id
Design Standard IEC_ed3
Site Type Offshore
Turbine Type 4.0-120
ULS post processing setup
ULS_All_SupportStruct ULS_Foundation_Foun ULS_Foundation_Foun
Page 4/4 ULS_Tower_Tower
ure_AllMax dation dation_AllMax
INACTIVE ACTIVE INACTIVE INACTIVE

Design Design Load Weibull A Weibull k ITUCI


Averaging scheme Averaging scheme Averaging scheme Averaging scheme
situation Case [-] [-] [Y/N]

BSH 35h Storm BSH35i

BSH35i_-EWH

BSH35p

Power production DLC11

DLC12 max 14.04 3.24

DLC12 max max max 14.04 3.24

DLC12 14.04 3.24

DLC13 max mean_windDir|windSpdmean_windDir|windSpd max

DLC14 max max max max

DLC15 max max max max

DLC16 max mean_windDir.windSpd mean_windDir.windSpd max


Parked (standing
DLC61 max mean_misalignment|windDir.yawErr
mean_misalignment|windDir.yawErr max
still or idling)
DLC62 max meanrun_misalignment|windDir.yawErr
meanrun_misalignment|windDir.yawErr max

DLC63 max mean_misalignment|windDir.yawErr


mean_misalignment|windDir.yawErr max

DLC64_h max max max max 14.04 3.24

DLC64_h 14.04 3.24

DLC64_l max max max max 14.04 3.24 Y

DLC64_l 14.04 3.24 Y


Parked and fault
DLC71 max meanhalf_misalignment|windDir.windSpd
meanhalf_misalignment|windDir.windSpd max
conditions
DLC72 max max max 14.04 3.24 Y

DLC72 max max max max 14.04 3.24 Y

DLC72 14.04 3.24 Y


Transport,
DLC81_a max mean_azimuth|yawErr mean_azimuth|yawErr max
assembly,
DLC81_b max mean_azimuth|yawErr mean_azimuth|yawErr max

DLC83 max max max max 14.04 3.24 Y

DLC83 14.04 3.24 Y

Internal DLC96 max max max max

DLC96
Sea ice - Power
DLCE3 max max max max
production
DLCE4 max max max max

DLCE4

Sea ice - Parked DLCE6 max max max max

DLCE7 max max max max

DLCE7

Prerun simulations PrepStartup


Internal - Ver_BladeCrossS
Verification ectionalPlot
Ver_BladeEigenV
alues
Ver_Controller
Ver_FoundEigen
Values
Ver_ModalAnaly
sis
Ver_PowerCurve
NoTurb
Ver_PowerCurve
Turb
Ver_Status
Ver_TowerFound
EigenValues
Ver_TurbineEige
nValues
64
Bibliography

[1] BTM Consult, International wind energy development - World update 2010, Technical Re-
port, 2010.
[2] Ernst & Young, Offshore wind in Europe: Walking the tightrope to success, Technical Report,
2015.
[3] Global Wind energy Council, Global wind report: Annual market update, Technical Report,
2015.
[4] Siemens, Fact sheet: Offshore wind power - April 2015, Technical Report, 2015.
[5] International Electrotechnical Comission, IEC 61400 Ed. 3: Wind turbines - Part 1: Design
requirements, 2005.
[6] Germanischer Lloyd, Guideline for the certification of offshore wind turbines, 2005.
[7] Risø DTU National Laboratory for Sustainable Energy and Det Norske Veritas, Guidelines for
design of wind turbines, 2002.
[8] American Society of Civil Engineers and American Wind Energy Association, Recommended
practice for compliance of large land-based wind turbine support structures, 2011.
[9] N. Bazeos, G. D. Hatzigeorgiou, I. D. Hondros, H. Karamaneas, D. L. Karabalis, and D. E.
Beskos, Static, seismic and stability analyses of a prototype wind turbine steel tower, Engi-
neering Structures (2002).
[10] O. Kiyomiya, T. Rikiji, and P. van Gelder, Dynamic response analysis of onshore wind energy
power units during earthquakes and wind, Wind Energy 3, 520–526 (2002).
[11] I. Lavassas, G. Nikolaidis, P. Zervas, E. Efthimiou, D. I. N., and C. C. Baniotopoulos, Analysis
and design of the prototype of a steel 1 MW wind turbine tower, Engineering Structures
(2003).
[12] E. Nuta, C. Christopoulos, and J. Packer, Methodology for seismic risk assessment for tubular
steel wind turbine towers :application to Canadian seismic environment, Canadian Journal
of Civil Engineering (2011).
[13] M. Hongwang, Seismic analysis for wind turbines including soil-structure interaction combin-
ing vertical and horizontal earthquake, 15th World Conference on Earthquake Engineering
(2012).
[14] O. Diaz Nevarez and L. Suarez, An analytical model for wind turbines subjected to seismic
motions, Revista Internacional de Desastres Naturales, Accidentes e Infraestructura Civil
(2012).
[15] X. Zhao and P. Maißer, Seismic response analysis of wind turbine towers including soil-
structure interaction, Proceedings of the Institution of Mechanical Engineers, Part K: Journal
of Multi-body Dynamics (2006).

65
66 BIBLIOGRAPHY

[16] T. Ishihara and M. Sawar, Numerical and theoretical study on seismic response of wind
turbines, European Wind Energy Conference and Exhibition (2008).
[17] X. Jin, H. Liu, and W. Ju, Wind turbine seismic load analysis based on numerical calculation,
Strojnis̃ki vestnik - Journal of Mechanical Engineering (2014).
[18] V. Indu and P. Binu, Seismic analysis of wind turbines using Opensees, American Journal of
Engineering Research (2013).
[19] I. Prowell, M. Veletzos, A. Elgamal, and J. Restrepo, Experimental and numerical seismic
response of a 65 kW wind turbine, Journal of Earthquake Engineering (2009).
[20] I. Prowell, An experimental and numerical study of wind turbine seismic behavior, PhD thesis,
University of California, San Diego, 2011.
[21] D. Witcher, Seismic analysis of wind turbines in the time domain, Wind Energy (2005).
[22] U. Ritschel, I. Warnke, and J. Kirchner, Wind turbines and earthquakes, 2nd World Wind
Energy Conference (2003).
[23] N. Alati, G. Failla, and F. Arena, Seismic analysis of offshore wind turbines on bottom-fixed
support structures, Philosophical Transactions of the Royal Society (2015).
[24] E. Ntambakwa and M. Rogers, Seismic forces for wind turbine foundations: Wind turbine
structures, dynamics, loads and control, AWEA Windpower Conference (2009).
[25] G. N. Stamatopoulos, Response of a wind turbine subjected to near-fault excitation and com-
parison with the Greek Aseismic Code provisions, Soil Dynamics and Earthquake Engineering
(2013).
[26] V. Valamesh and A. T. Myers, Aerodynamic damping and seismic response of horizontal axis
wind turbine towers, Journal of Structural Engineering (ASCE) (2014).
[27] E. Bossanyi, GH Bladed Theory manual, Garrad Hassan and Partners Limited, 2003.
[28] J. M. Jonkman and Buhl, FAST User’s guide, National Renewable Energy Laboratory, 2005.
[29] M. L. Buhl Jr and A. Manjock, A comparison of wind turbine aeroelastic codes used for
certification, American Institute of Aeronautics and Astronautics (2006).
[30] I. Prowell, El, and Jonk, FAST simulation of wind turbine seismic response, Technical Report,
National Renewable Energy Laboratory, 2010.
[31] M. Asareh and I. Prowell, Seismic Loading for FAST, Technical Report, National Renewable
Energy Laboratory, 2011.
[32] T. K. Datta, Seismic analysis of structures, John Wiley & Sons, 2010.
[33] B. Song, Y. Yi, and J. C. Wu, Study on seismic dynamic response of offshore wind turbine
tower with monopile foundation based on M method, Advanced Materials Research (2013).
[34] M. Ghaffar-Zadeh and F. Chapel, Frequency-independent impedances of soil-structure sys-
tems in horizontal and rocking modes, Earthquake Engineeering and Structural Dynamics
(1983).
[35] R. A. Kjørlaug, A. M. Kaynia, and A. Elgamal, Seismic response of wind turbines due to
earthquake and wind loading, Proceedings of the 9th International Conference on Structural
Dynamics, EURODYN 2014 (2014).
[36] K. Hacıefendioğlu, Stochastic seismic response analysis of offshore wind turbine including
fluid-structure-soil interaction, The Structural Design of Tall and Special Buildings (2010).
BIBLIOGRAPHY 67

[37] G. Gazetas, Foundation vibrations, Foundation engineering handbook , 553–593 (1991).

[38] D. H. Kim, S. G. Lee, and I. K. Lee, Seismic fragility analysis of 5 MW offshore wind turbine,
Renewable E (2014).

[39] G. Gazetas and R. Dobry, Horizontal response of piles in layered soils, Journal of Geotechnical
Engineering (1984).

[40] A. E. Kampitsis, E. J. Sapountzakis, S. K. Giannakos, and N. A. Gerolymos, Seismic soil-


pile-structure kinematic and inertial interaction - a new beam approach, Soil Dynamics and
Earthquake Engineering (2013).

[41] I. C. Sapountzakis, Evangelos J.and Dikaros, A. E. Kampitsis, and A. D. Koroneou, Nonlinear


response of wind turbines under wind and seismic excitations with soil-structure interaction,
Journal of Computational and Nonlinear Dynamics (2015).

[42] R. W. Clough and J. Penzien, Dynamics of structures, Computers & Structures, Inc., 1995.

[43] J. P. Wolf, Dynamic soil-structure interaction, Prentice-Hall, 1985.

[44] I. M. Idriss and J. I. Sun, SHAKE 91: A computer program for conducting equivalent linear
seismic response analyses of horizontal layered soil deposits, University of California, 1992.

[45] A. R. Kottke and E. M. Rathje, Technical manual for Strata, Pacific Earthquake Engineering
Research Center, 2009.

[46] S. L. Kramer, Geotechnical earthquake engineering, Prentice Hall, 1996.

[47] J. S. Przemieniecki, Theory of matrix structural analysis, McGraw-Hill, 1968.

[48] Z. Friedman and J. B. Kosmatka, An improved two-node Timoshenko beam finite element,
Computers and S (1993).

[49] B. Oskin, What is a subduction zone?, http://www.livescience.com/


43220-subduction-zone-definition.html, Accesed: 05-07-2016.

[50] GLGArcs, Japan in a subduction zone, http://www.glgarcs.rgr.jp/intro/subduction.


html, Accesed: 05-07-2015.

[51] S. Wei, D. Helmberger, Z. Zhongwen, and R. Graves, Rupture Complexity of the MW 8.3
sea of okhotsk earthquake: Rapid triggering of complementary earthquakes?, Geophysical
Research Letters (2013).

[52] Earth Observatory of Singapore, The great East Japan (Tohoku) 2011 earth-
quake: Important lessons from old dirt, http://www.earthobservatory.sg/news/
great-east-japan-tohoku-2011-earthquake-important-lessons-old-dirt, 2011, Ac-
cesed: 05-07-2016.

[53] N. C. J. Venture, Selecting and scaling earthquake ground motions for performing response-
history analyses, U.S. Department of Commerce, 2011.

[54] Japan Society of Civil Engineers, Guidelines for design of wind turbine support structures and
foundations, 2007.

[55] T. Ancheta, R. B. Darragh, J. P. Stewart, E. Seyhen, W. J. Silva, B. S. J. Chiou, K. E. Wooddell,


R. W. Graves, A. R. Kottke, D. M. Boore, T. Kishida, and J. Donahue, PEER NGA-West2
Database, Pacific Earthquake Engineering Research Center, 2012.

[56] D. M. Boore, A. Azari Sisi, and S. Akkar, Using pad-stripped acausally filtered strong-motion
data, Bulletin of the Seismological Society of America (2012).
68 BIBLIOGRAPHY

[57] International Institute of Seismology and Earthquake Engineering IISEE, Seismic bedrock
and engineering bedrock, http://iisee.kenken.go.jp/net/yokoi/bedrock/index.htm,
Accesed: 05-07-2016.
[58] M. Davoodi, M. Sadjadi, P. Goljahani, and M. Kamalian, Effects of near-field and far-field
earthquakes on seismic response of SDOF systems considering soil-structure interaction, Pro-
ceedings on the 15th World Conference on Earthquake Engineering (2012).
[59] American Society of Civil Engineers, ASCE/SEI 7-10: Minimum design loads for buildings and
other structures, 2010.
[60] E. L. Wilson and I. Suhawardy, A clarification of the orthogonal effects in a three-dimensional
seismic analysis, Earthquake Spectra (1995).
[61] O. A. Lopez and R. Torres, The critical angle of seismic rotation and structural response,
Earthquake E (1997).
[62] E. Kalkan and N. S. Kwong, Pros and cons of rotating ground motion records to fault-
normal/parallel directions for response history analysis of buildings, Journal of Structural
Engineering (2014).
[63] J. Penzien and M. Watabe, Characteristics of 3-dimensional earthquake ground motions,
Earthquake (1975).
[64] L. Al Atik and N. Abrahamson, An improved method for nonstationary spectral matching,
Earthquake Spectra (2010).
[65] Seismosoft Ltd., SeismoMatch, http://www.seismosoft.com/seismomatch, Accesed: 05-
07-2016.
[66] N. Abrahamson, Non-stationary spectral matching, Seismological Research Letters (1992).

[67] J. Hancock, J. Watson-Lamprey, N. A. Abrahamson, J. J. Bommer, A. Markatis, E. McCoyh,


and R. Mendis, An improved method of matching response spectra of recorded earthquake
ground motion using wavelets, Journal of Earthquake (2006).
[68] M. Teshigawara, Appendix A: outline of earthquake provisions in the Japanese building
codes, Preliminary Reconnaissance Report of the 2011 Tohoku-Chiho Taiheiyo-Oki (2012).

[69] N. Subramanian, Steel structures: Design and practice, Oxford University Press, 2011.
[70] M. Novak, T. Nogami, and A. Fakhry, Dynamic soil reactions for plane strain case, Faculty of
Engineering Science, University of Western Ontario (1977).
[71] N. Makris and G. Gazetas, Dynamic pile-soil-pile interaction. Part II: Lateral and seismic
response, Earthquake Engineering & Structural Dynamics (1992).
[72] G. Gazetas and R. Dobry, Simple radiation damping model for piles and footings, Journal of
Engineering Mechanics (1984).
[73] D. Varun, Assimaki and G. Gazetas, A simplified model for lateral response of large diameter
caisson foundations - Linear elastic formulation, Soil Dynam (2009).

S-ar putea să vă placă și