VIBRATION-BASED STRUCTURAL HEALTH MONITORING OF WIND TURBINES Gustavo
VIBRATION-BASED STRUCTURAL HEALTH MONITORING OF WIND TURBINES Gustavo
VIBRATION-BASED STRUCTURAL HEALTH MONITORING OF WIND TURBINES Gustavo
VIBRATION-BASED STRUCTURAL
HEALTH MONITORING OF WIND
TURBINES
April 2016
Vibration-Based Structural Health Monitoring of Wind Turbines
To my parents
Vibration-Based Structural Health Monitoring of Wind Turbines
4
Vibration-Based Structural Health Monitoring of Wind Turbines
9
ABSTRACT
Wind power is one of the most attractive and reliable options for renewable and clean energy. In the
last years, the installed capacity for exploitation of wind energy has consistently increased worldwide.
Alongside, the technological evolutions of this sector have been based on the development of multi-
MW systems with increasingly larger structures elements, such as support structures and rotor blades.
These wind turbines present considerable advantages considering the power production; however,
they also represent several structural challenges: these wind turbines have become increasingly flexible
structures, prone to resonance phenomena and to rapid wear.
In that sense, the structural integrity and proper operation of large wind turbines is a concern for
owners and operators, specially in turbines installed offshore. Monitoring systems capable of
evaluating the condition of structural elements such as foundation, tower and blades, are seen as a
suitable response to address this problem. Among the several techniques available, vibration-based
systems reveal a great potential for structural monitoring.
Therefore, this thesis is focused on the development and implementation of a vibration-based dynamic
monitoring system for wind turbines. This system is based on Operational Modal Analysis tools,
capable of accurately estimate the modal properties of wind turbine structures from their dynamic
response. The efficiency of these tools in the modal identification of wind turbines was validated with
the analysis of vibrational data collected during a short period from onshore (Izar Bonus 1.3MW/62)
and offshore (Vestas V90-3.0MW) models.
Afterwards, a vibration-based monitoring system was implemented on a 2.0 MW Senvion MM82 wind
turbine allowing the collection of data during more than one year. Throughout this program, the
system was defined with two main purposes: detection of structural anomalies (based on the
continuous estimation and statistical analysis of the variations of the modal properties) and fatigue
assessment (based on a reduced number of installed sensors). The program was complemented with
the analysis of alternative solutions for optimization of the number of installed sensors. The usefulness
of the system was tested with different damage scenarios. It was concluded that the monitoring system
is capable of detecting small damages in the support structure of both onshore and offshore wind
turbines. The tools developed for fatigue evaluation using the estimated modal responses showed
promising results but a complete validation of their accuracy need further research.
Vibration-Based Structural Health Monitoring of Wind Turbines
Vibration-Based Structural Health Monitoring of Wind Turbines
9
RESUMO
Energia eólica é uma das mais atrativas e fiáveis opções para energia renovável e limpa. Nos últimos
anos, a capacidade instalada para aproveitamento do recurso eólico tem crescido de forma consistente
em todo o mundo. Ao mesmo tempo, a evolução tecnológica do sector tem sido baseada no
desenvolvimento de sistemas com múltiplos MW, recorrendo a elementos estruturais de dimensões
cada vez maiores, tais como estruturas de suporte e pás. Estas torres eólicas apresentam vantagens
consideráveis do ponto de vista da produção; no entanto, representam também diversos desafios
estruturais: estes aerogeradores tornaram-se estruturas cada vez mais flexíveis, propensos a fenómenos
de ressonância e a um desgaste rápido.
Nesse sentido, a integridade estrutural e a correta operação de modernas torres eólicas é uma
preocupação para donos e operadores, especialmente em turbinas instaladas em offshore. Sistemas de
monitorização capazes de avaliar a condição de elementos estruturais, tais como fundação, torre e pás,
são vistos como uma resposta adequada para lidar com este problema. De entre as várias técnicas
disponíveis, sistemas baseados na análise de vibrações revelam um grande potencial para
monitorização estrutural.
Assim, esta tese foca-se no desenvolvimento e implementação de um sistema de monitorização
dinâmico baseado na análise de vibrações para torres eólicas. Este sistema é baseado em ferramentas de
Análise Modal Operacional, capazes de estimar com precisão as propriedades modais das estruturas de
aerogeradores a partir da sua resposta dinâmica. A eficácia destas ferramentas na identificação modal
de torres eólicas foi validada com a análise de dados de vibração colhidos durante um curto período de
tempo para uma torre instalada em terra (Izar Bonus 1.3MW/62) e uma outra em offshore (Vestas
V90-3.0MW).
Posteriormente, um sistema de monitorização foi implementado numa torre eólica Senvion MM82,
permitindo a recolha de dados duranta mais de um ano. Ao longo deste programa, o sistema foi
definido com dois objetivos principais: deteção de anomalias estruturais (baseado na estimativa
contínua e análise estatística das variações das propriedades modais) e avaliação da condição de fadiga
(com base num reduzido número de sensores instalados). O programa foi complementado com a
análise de soluções alternativas para otimização do número de sensores instalados. A utilidade do
sistema foi testada com diferentes cenários de dano. Foi concluído que o sistema de monitorização é
capaz de detetar danos reduzidos na estrutura de suporte de torres instaladas em terra e em offshore.
As ferramentas desenvolvidas para avaliação de fadiga, através de estimativas de respostas modais,
mostraram resultados promissores, no entanto uma validação completa da sua precisão requere
desenvolvimentos adicionais.
Vibration-Based Structural Health Monitoring of Wind Turbines
Vibration-Based Structural Health Monitoring of Wind Turbines
19
ACKNOWLEDGEMENTS
This thesis would not be possible without the help and support of several people. Here I present my
sincere gratitude:
o to Prof. Álvaro Cunha, my advisor, for challenging me with such an interesting topic, for all
the support given during this period and for his care to ensure all the conditions and means
for the proper development of this thesis;
o to Prof. Filipe Magalhães, my co-advisor, for all the support, for the endless hours he received
me in his office and for the patience and sympathy to clarify unscheduled problems;
o to Prof. Elsa Caetano, my co-advisor, for her support whenever needed and for the
opportunity to cooperate on interesting projects;
o to Eng. José Carlos Matos, Eng. Miguel Marques and Eng. Silvina Guimarães from INEGI, for
the opportunity to work within the Rotor Blade Extension project;
o to Prof. Christof Devriendt and Doctor Wout Weijtjens from OWI-Lab, for the interesting
conversations about modal identification and for the opportunity to use the data of the Vestas
V90-3.0MW wind turbine;
o to Eng. José Saraiva from Senvion and to Eng. Carlos Cardoso and Eng. Frederick Saborano
from Cavalum, for all the support given within the study of the Senvion MM82 wind turbine;
o to Prof. Álvaro Rodrigues and Prof. Fernando Marques da Silva, for the interesting suggestions
and conversations that helped improve this thesis;
o to my work colleagues and friends Nuno, Joana(s), Celeste, Berawi, Sandro, Fernando, Aires,
Ana, André, Ricardo and Sérgio, for the fun that was working in our H304;
o to my parents and brother, to whom I owe who I am;
o to Paulina, for putting up with my bad mood and, even thus, always cheer me up;
o to Fundação para Ciência e Tecnologia, for the financial support through the PhD fellowship
SFRH/BD/79328/2011.
Vibration-Based Structural Health Monitoring of Wind Turbines
Vibration-Based Structural Health Monitoring of Wind Turbines
9
NOMENCLATURE
ܰ Number of blades
ܲ Power
Peak period
wind speed
Airfoil chord
Innovation sequence
Natural frequency
Pressure
Modal coordinates
Input vector
State vector (which contains the displacement and velocity vectors of the system)
State vector
Angle of attack
̂ Residual error
Pitch angle
Damping ratio
Mass density
Standard deviation
′ Local solidity
Matrix with the mode shape configurations at the measured degrees of freedom, angle of inflow
∗ Moore-Penrose pseudo-inverse
∗ Transpose operator
∗ Absolute value
∗ Covariance
∗ Expectation value
∗ Real part
∗ Trace of a matrix
CM Condition Monitoring
FA Fore-Aft
SS Side-Side
CONTENTS
1. INTRODUCTION ............................................................................................................................21
1.1 Research Context .............................................................................................................................. 21
1.2 Objectives and Main Contributions ............................................................................................... 24
1.3 Organization of the Thesis ............................................................................................................... 26
21
Vibration-Based Structural Health Monitoring of Wind Turbines
22
Vibration-Based Structural Health Monitoring of Wind Turbines
23
Vibration-Based Structural Health Monitoring of Wind Turbines
24
Vibration-Based Structural Health Monitoring of Wind Turbines
25
Vibration-Based Structural Health Monitoring of Wind Turbines
B.3. Analysis of Fibre Bragg Grating Sensors Measurements (After RBE) .....................................368
REFERENCES..................................................................................................................................... 371
26
Chapter 1
1
INTRODUCTION
21
Introduction
Commission. This interconnection, which could achieve 15 % of the total installed capacity by 2030,
would lead to a considerable boost of implementation of new wind power plants. Alongside, an
important portion of the installed wind power capacity in Portugal is achieving its design lifetime, with
repowering operations being expected in the next years.
Given this policy background, it is consensually accepted that wind power is a key element in the
present and future electric energy strategy. New wind turbines are expected to be installed in the next
few years with power outputs considerably higher than the nowadays common values, mainly due to
cost aspects. In addition, offshore locations are becoming very attractive to install turbines due to their
higher production yields. Three significant difficulties will be faced by this new wind turbine models:
o The increase of the wind turbines dimension, particularly the increase of the height of tower
and blades length, will turn the wind turbine into more flexible structural systems.
Consequently, the study of the dynamic characteristics of the structure will be of major
importance in order to avoid resonance problems and rapid wear of the components;
o The installation of wind turbines in offshore locations is still shrouded in some uncertainty
concerning the real wear suffered by the structures. On the other hand, the costs associated
with offshore wind turbines are still high, which demand for optimization in their design;
o Apart from reducing the cost of installation of the wind power plants (CAPEX), the reduction
of the costs over the period of its operation (OPEX) also represents an important vector for
competitiveness. A crucial aspect to lower this cost is the optimization of the maintenance
strategy. Considering the high power outputs of new models, the downtime of a single wind
turbine may represent a significant income loss. In that sense, several monitoring system are
being developed to remotely assess the condition of several components of the wind turbine,
allowing to optimize the maintenance interventions according to expected loss of production
during downtime, availability of replacing components and meteorological conditions (in the
case of offshore wind farms or onshore wind farms in mountains).
These facts address the new challenges that, from a structural point of view, will concern the owners of
wind farms. In that sense, a real-time monitoring system capable of assessing the structural integrity of
the support structure of wind turbines (including tower and foundation) can be a very useful tool to
aid decision-making. With a monitoring system, the wind farm owner can analyse the structural health
of the wind turbines in remote locations and adapt the maintenance strategy according to the received
data. In short, in possession of this information, the owner could maximize the investment.
In order to be effective, a monitoring system installed at a wind turbine support structure should be
able to answer two important questions:
o Is the structure damaged?
o How is the fatigue condition of the structure? Is the structure able to fulfil its period of life? Or
should the structure be disassembled before (or after) this period?
To meet these two aspects, different modules need to be integrated into one vibration-based structural
health monitoring (SHM) system: a structural anomalies detection monitoring system and a fatigue
monitoring system.
The structural anomalies detection monitoring system is a tool with the aim of detecting damage. It
can be based on the continuous tracking of the modal properties of the structure. For this purpose,
22
Chapter 1
tools already tested on civil engineering structures can be used in wind turbines in order to extract the
monitored features. These tools, from a branch of the experimental modal analysis field named as
Operational Modal Analysis (OMA), permit to identify the modal characteristics of structures during
the various operating conditions of the turbines. Thus, considering that damage is usually associated
with stiffness reductions, the natural frequencies of the various vibration modes can be used as damage
indicators. The study of the evolution and variation of these indicators indirectly evidence the
existence of damage in the structure.
The second module comprises the fatigue analysis of the structure. Fatigue assessment is a key point in
wind turbines, since these structures are mainly subjected to dynamic loads. Thereby, this module
permits to assess the evolution of the wear of the wind turbine support structure over time. This
information can then be compared with the estimated lifetime of the structure in order to validate the
suitability of this period.
23
Introduction
24
Chapter 1
identification of the modal parameters under all operating conditions, to remove the influence
of operational and environmental effects and to compute indices permitting to detect the
presence of small structural anomalies; and, in the second module, estimate the fatigue
conditions along the support structure of a wind turbine (section 7.2);
o Conception, assembly and operation of a monitoring system on a Senvion MM82 wind
turbine for a period of one year, where the developed routines were tested, demonstrating the
ability of the system to detect the occurrence of small damages (section 7.3);
o Optimization analysis concerning the ability (and loss of accuracy) of the dynamic monitoring
system to operate with a reduced number of installed sensors (section 7.3.8).
25
Introduction
26
Chapter 2
2
FROM WINDMILLS TO THE MODERN
WIND TURBINES
27
From Windmills to the Modern Wind
W Turbines
28
Chapter 2
Figu
ure 2.4 - Savonius Rotor Figure 2.5
2 - Darrieus Rootor (Darrieus, 1931)
29
9
From Windmills to the Modern Wind
W Turbines
2.1.1.2 Th
he First Largge Machines
One of thee first large projects waas the 100 kkW, 30 m diameterd Ballaclava windd turbine in 1931
(Sectorov, 11973) (Figuree 2.6). It had
d a 30 m heigght steel lattice tower an
nd was locateed at Balaclavva, on
Crimea, near the Blackk Sea. It had d a three-blaaded rotor with
w omatic yaw system and pitch
an auto
control (adjjustment oveer the angle of
o the blade ffacing the wiind – see Chaapter 3). Althhough it had
d some
primitive soolutions (succh as gears made
m of woodd and simple blade surfacce treatment)), it achievedd good
results. Thiis machine was
w connected to a steam m power statio on and is con
nsidered the first wind tu
urbine
to be intercoonnected to an AC utilityy system (Sheepherd, 19900).
In late 19399, the Ameriican Engineeer Palmer C.. Putnam approach the S. S Morgan Sm mith Compaany of
York with a new wind tu urbine with the latest tecchnology (Vooaden, 1943). The constrruction of thiis new
project startted in 1941 on
o the top off a mountain n called “Gran
ndpa’s Knobb” (near Vermmont, USA) and it
is considereed a landmaark in the wiind energy hhistory. Thiss model refleects the demmand for red ducing
production costs with laarge wind turrbines with hhigh power outputs.
o
30
Chapter 2
Technically, the Smith-Putnam wind turbine was a remarkable and giant project for that time. It had a
53.3 m diameter, 2-bladed upwind rotor that generated a power output of 1.25 MW with a
synchronous (AC) generator (Shepherd, 1990) and a steel lattice tower with a height of 33.5 m
(Voaden, 1943) (Figure 2.7). It included some technological innovations such as (Shepherd, 1990):
o Full span active control of the blade pitch (the blades were made of stainless-steel);
o The use of individual flapping hinges on the blades to reduce gyroscopic loads on the shaft (a
similar problem previously faced in the development of the Jacobs turbine);
o Active yaw control by mean of a servomotor turning a pinion meshing with a large bull gear
between the machinery house (or nacelle) and the tower.
The generator was connected to the electrical system at wind speeds higher than 8 m/s, reached its
rated power at 13.4 m/s (Voaden, 1943) and it ran until the maximum wind speed of 32 m/s (Johnson,
2001). The Smith-Putnam wind turbine is considered the first large wind turbine with two blades
(Johnson, 2001) and it was the largest wind turbine for almost 40 years. Notwithstanding, it faced
some important drawbacks. From the technical point view, it was concluded that the blades were
undersigned, presenting several cracks after some operating time. From the economic point of view, it
demonstrated that some improvements would have to be taken in order to be become cost effective
with respect to the conventional power generation processes (Gasch and Twele, 2012).
At this time, the increase of the capacity of the wind turbines was already a confirmed trend. Although
the Smith-Putnam wind turbine had been a one-off project, some previous and future projects
confirmed this trend, some of them with utopic designs (like the design of Hermann Honnef with a
power output of 20 MW (Hau, 2006)).
In 1939, the World War II began. From the end of World War I (1918) to 1939, the interest in wind
power in Denmark decreased (Vestergaard, Brandstrup et al., 2004). The price of coal and oil had
become sufficiently low to decelerate the interest in wind energy. However, with the beginning of the
World War II, the increase in the price of fossil fuels led, once again, to the rise of wind energy interest.
Some old commercial LaCour machines were reactivated and some new wind turbines were built.
Also in Denmark, a new company (F.L. Smidth Company) developed a new model during this period,
relying in a small, conservative design. In addition, new concepts of modern aerodynamics (modern
airfoils) were applied to the blades (Ackermann and Söder, 2002). Under this philosophy, reliable
models with power outputs ranging from 50 kW to 70 kW were developed, with rotor diameters of
17.5 m and 24 m, respectively. It is interesting to note that the model with the higher power output was
designed with a concrete tower, not a very common solution at the time.
The models developed by the Smidth Company are an example of a development trend, choosing
simple designs over heavy and complicated models, in opposition to the Smith-Putnam turbine design.
In fact, the Smidth wind turbines are considered the forerunners of modern wind turbine generators
by some authors and one of the pioneers of the Danish wind energy market (Shepherd, 1990;
Ackermann and Söder, 2002). By the first half of the 20th century, the wind energy market was already
a reality due to the number of small wind turbines installed, being the primary energy source in some
areas (Hau, 2006).
31
From Windmills to the Modern Wind
W Turbines
a
a) b)
During the 1930s and 1940s, much theoretical aand design acctivity in the wind energyy field was held h in
Germany. T This developm ment was un ndertaken byy the German n Reich with the aim of a self-sufficien ncy in
fuel power (Hau, 2006). For this purpose,
p thee “Reichsarbeitsgemeinscchaft Windkkraft” (RAW W) was
h renowned scientists,
created with s tecchnicians an nd industrialss. Among thee projects dev
eveloped und der the
RAW progrram, the pro oject led by the engineerr Franz Kleiinhenz, in 19 937, stands oout, reflectin
ng the
ambition off building large machinees to reduce the cost of energye produuction. This wind turbin ne had
remarkable characteristtics like a 130 m rotor ddiameter with h a three or four-bladedd downwind rotor.
This rotor w
would have a power outp put of 10 MW W at a tip speed of 5 and the
t generatorr would havee been
from the diirect drive tyype. The designed towerr was a guyed d tubular steeel tower, wiith the uppeer part
yawing withh the rotor. The
T concept was improvved from 19338 to 1942 an nd was readyy for constru uction.
However, thhe war preveented its consstruction (H au, 2006). It is interesting to note thaat, even nowadays,
some of theese characteriistics are stilll not widely iimplemented d in commerrcial wind turrbines
Apart fromm Denmark and a USA, oth her regions aalso showed some intereest in the expploitation off wind
energy, alth
hough in a sm mall scale. In
n the Unitedd Kingdom, there
t was a wind
w energy program ru unning
from 1948 to the earlyy 1960s (Speera, 1994). A Among some different designs
d deveeloped durin ng the
program, th he Enfield-A
Andreau wind d turbine prooject stands out from th
he others duee to its innoovative
design, althoough associaated with low
w efficiency (H
Hau, 2006))..
nce, there waas some interest in the ddevelopment of experimental wind tuurbines durin
Also in Fran ng this
time. The 30.1 m rotor diameter, 8000 kW Best-R Romani turb
bine and the 35 m rotor ddiameter, 1.0
0 MW
Vadot turbiine are examples of projects from thiss time (Hau, 2006).
During the post-war perriod in Germmany, Ulrich Hutter was the main name in the fieeld of wind power.
p
Hutter, whoo had alreadyy been distin
nguished on previous papers on the theory
t of win
ind turbines (Hau,
2006), cond
ducted some experimentaal tests durinng the 1950s. His studies culminated in the 34 m rotor
diameter W
W-34 wind turrbine with an n AC synchrronous generrator (Gasch and Twele, 22012) with a rated
power output of 100 kWW at a wind speed of 8 m//s (Figure 2.99) (Johnson, 2001). This tturbine, ereccted in
32
Chapter 2
1957,, had a two o-bladed, dow wnwind rotoor with a siimple hollow w pipe toweer with 22 m of heightt
suppoorted by gu uy wires (Jo ohnson, 20001). One off Hutter’s main m prioritiies was the lightweightt
consttruction. Forr this reason
n, the bladess were madee of an advaanced glass-ffibre compossite materiall
(Hauu, 2006) and they were aerodynamic
a cally refined, being connected to a teeetering hub b (see Figuree
3.6). T
The solution
n found for th his connectioon was anothher innovatioon: instead oof a more com
mplex design n
like tthe one usedd in the Smmith-Putnam wind turbin ne, the teeteering movem ments of thee rotor weree
aeroddynamically damped by mechanically
m y coupling th
he teetering angle to the blade-pitch angle (Hau,,
2006)). The pitch mechanism
m of
o the bladess changed theeir angles at higher
h wind speed to keeep a constantt
speedd of rotation of the rotorr (Vargo, 19774). This win
nd turbine is considered as an examp ple of a third
d
desiggn philosophhy, relying on n the reducttion of the structure
s weeight and onn the optimization of itss
aeroddynamics.
33
3
From Windmills to the Modern Wind Turbines
refurbished in the mid-1970s at the request of NASA for measurements, in the framework of an
American program (Hau, 2006).
Although Hutter and Juuls’ projects were considered a success, the wind turbine projects in general
were not yet reliable, with numerous problems and faults to be dealt with (Hau, 2006). Besides these
reasons, the extremely low prices of primary fuels after the World War II, along with the increased
investments in nuclear energy, demonstrated that the wind generated electricity was too expensive to
compete with other sources (Gasch and Twele, 2012). Due to this, the wind energy industry suffered
another breakdown during the 1960s.
34
Chapter 2
35
5
From Windmills to the Modern Wind
W Turbines
Figure 2.12
2 – MOD-1 w
wind turbine (W
Wikipedia, 2009
9)
The next iteeration of thee NASA proggram started in 1977 with h the MOD-2 2 project (Goordon, Andrews et
al., 1983) (F
Figure 2.13). This project was design ned as a sequuence of the findings andd developmeents of
the previous NASA/DO OE projects (M MOD-0, MO OD-0A and MOD-1)
M and
d, for that reaason, it is referred
to as a secoond generatiion machinee (Ramler an nd Donovan, 1979). Thee project of the MOD-2 wind
turbine wass a design evvolution of the t proposedd design forr the MOD-1A model. T The MOD-2 wind
turbine wass a two-bladeed, upwind rotor
r with a ddiameter of 91.4
9 m. Thiss model had a power output of
2.5 MW at a wind speed Gordon, Anddrews et al., 1983).
d of 12.3 m/s and used a ssynchronouss generator (G
Three of thee major mod difications fro
om the MOD D-1 design were
w (Ramler and Donovaan, 1979):
o Thee power conntrol. Instead
d of the heavvy and comp
plex full-span pitch conttrol of the blades,
b
pow o the wind turbine was handled thrrough the pitch of just thhe outer 30 % the
wer control of
blad
de span;
o Thee tower desiggn philosoph hy. The desiggned tower was w considerrably more fflexible than those
used d in previous projects. It consisted inn a 61 m high h steel tower with a cylinddrical shell design.
d
In tthis design, the natural frrequency of tthe tower is lower
l than th
he frequencyy of rotation of the
bladdes. For this reason this design
d is calleed “soft-soft”” (see 3.3.4);
o Thee load reducttion. The rottor was desiggned to allow p to 5o in annd out of the plane
w teeter of up
of th
he rotor.
This design
n approach permitted
p to
o save weighht and, conseequently, red
duce costs. C
Comparing to
t the
MOD-1 dessign (which was smallerr), the MOD D-2 was arou und 10 % liighter (Ramller and Don
novan,
1979).
Three MOD D-2 machinees were built and tested in n a cluster in
n 1980 in Goodnoe Hills,, Washington n. The
disposition of the thrree wind tu urbines was defined in n order to study possiible aerodyn namic
interferencees between laarge turbines in downwiind (wake efffects) (Speraa, 1994). Tw wo other macchines
were also bbuilt for utiliity companiees. The MO D-2 machin nes suffered some techniical problem ms and
economic sttudies showeed that the costs were stilill not compeetitive. Howeever, this prooject proved that a
group of larrge wind turbbines could operate
o in a ttotally autom
matic and unaattended moode (Spera, 19 994).
36
Chapter 2
The tthird generaation of horizzontal-axis w wind turbinees from the NASA/DOE E program inncluded fourr
projeects: MOD-33, MOD-4, MOD-5A M annd MOD-5B B (although the MOD- 3 and MOD D-4 projectss
stoppped in an inittial phase) (SSpera, 1994).. The main goal
g of the MOD-5 prograam, started in i 1980, was,,
once again, to devvelop cost coompetitive, laarge wind turrbines (Baldw
win and Kennnard, 1985).
The MMOD-5A design, develop ped by Geneeral Electric Company
C forr NASA/DO OE, had a 122
2 m diameterr
rotorr with two bllades and thee hub was pllaced 76 m above
a the ground (Baldw win and Kennnard, 1985)..
The p power outpuut of this maachine was 7..2 MW with a rotor rotaation of 13 oor 17 rpm, deepending on n
wind conditions (Johnson, 20 001). Howevver, this macchine was no ot built due to low persp
pectives of a
comm mercial succeess from General Electriic which withdrew from the project.. Neverthelesss, this wind d
turbinne would rep
present, for nowadays
n staate-of-art, on
ne of the mosst powerful pprojects.
The M MOD-5B, developed
d byy Boeing forr NASA/DO OE, was builtt in 1987 (SSpera, 1995). This wind d
turbin
ne followed thet design co oncept of thee MOD-2 maachine: a two o-bladed, upw wind, teetereed rotor withh
a parttial span pitcch control off the blades (SSpera, 1995). The turbinee was supporrted by a cyliindrical steell
shell and had a rated
r power of 3.2 MW W and a roto or diameter of 97.5 m (FFigure 2.14). The majorr
innovvation of thee model was the t ability off its turbine to
t operate su
uccessfully att variable speeed (from 133
to 17.3 rpm).
37
7
From Windmills to the Modern Wind
W Turbines
F
Figure 2.15 – WTS-4
W wind turb
bine (Wikipediaa, 2009; Schwerrin, 2010)
During the 1970s, theree was also so ome relevantt research in wind powerr outside thee USA. In Caanada,
the research h efforts weere towards the verticall axis wind turbines (V VAWT), speecifically witth the
Darrieus deesign. Amongg others (sm mall) turbiness, the Nation
nal Research Council (NR RC) developeed the
Éole VAWT T. The installlation of thiss turbine wass complete in uebec, Canadda (Spera, 1994). It
n 1987 in Qu
was the firstt megawatt-cclass Darrieuus turbine, wi
with a 64 m diiameter and with an origginal power output
o
of 4.0 MW, operating att variable speeed (Figure 22.16). Howevver, to ensuree the longeviity of the machine,
the rated poower was lowwered to 2.5 MW.M The exxperience gained and the results achieeved were no ot very
encouragingg and the Caanadian proggram was teerminated (H Hau, 2006). Some
S researcch in the Daarrieus
concept field was also peerformed in the USA in tthe Sandia National
N Labo
oratories in A
Albuquerque.
a) b)
Figure 2.16
6 – Éole VAWT: a) general view
w (Wikipedia, 2005);
2 b) illustra
ation (Spera, 19994)
38
Chapter 2
In Euurope, some Governmen nts from Norrdic and Cen ntral Europeaan countries also develop ped research
h
progrrams during this period. In Denmarkk, an ambitio ous programm started in 11974. The main goal wass
“thatt it should bee possible to generate
g 10%
% of the Danish power requirement frrom wind energy withoutt
creatiing particulaar problems in the publiic power grid” (Hau, 20 006). For thiis purpose, the
t 200 kW W
Gedser wind turb bine was refufurbished forr a joint stud
dy with NAS SA. Furthermmore, two 63 30 kW wind d
turbinnes with 40 m diameter were builtt in Nibe, which w had th
heir first rotaation in 197
79 and 19800
(Interrnational En nergy Agencyy (IEA), 19884). These windw turbines had a threee-bladed, upwind rotorr
with steel blade spars
s and bo oth had a cooncrete toweer (Spera, 1994) (Figure 2.17). Each of the wind d
turbinnes had a diffferent poweer control meechanism: on ne had a tip--controlled rrotor while th
he other had
d
a full-span pitch control.
c Thesse wind turbiines operatedd for 15 yearrs.
Previiously in 1975, a 2.0 MW wind turbin
ne was built in
n Tvind, Den
nmark (Hau,, 2006) (Figu
ure 2.18).
Figure 2.17
2 – Nibe A wind
w turbine Fig
gure 2.18 – Tvinnd wind turbine
e
39
9
From Windmills to the Modern Wind
W Turbines
Fig
gure 2.19 – Nassudden wind tu
urbine Figure 2.20 – Maglarp windd turbine
The Germaan state su ubsidized program on the develop pment of wind w energy started in 1974
(Internationnal Energy Agency
A (IEAA), 1984). TThe most em mblematic prroject was thhe Growian wind
turbine, whhich was asseembled in 19 982 (Thiele, 1984). This wind
w turbinee was, at thaat time, the largest
l
Figure 2.21). It had a 100 m diameter , two-bladed
ever built (F d, downwind rotor whichh was supportted by
a 100 m talll steel tower stabilized
s byy several guy cables (Spera, 1994). Thee carbon-filaament blades had a
full-span pittch control. The
T Growian n machine wwas also the fiirst wind turbine to attemmpt variable--speed
operation aalthough it was
w not well succeed. Addditionally, the Growian n wind turbiine suffered other
problems, iincluding faatigue in major compo nents which h conducted d to its disaassemble in 1987
(Internationnal Energy Agency
A (IEA), 1989). De spite of the low success of the Grow wian project, some
smaller macchines operaated satisfacttorily (such as the 370 kW k Monopteros with a one-bladed, 48 m
diameter rootor (Figure 2.22)
2 (Spera, 1994).
40
Chapter 2
The large wind turbines developed during the 1970s and 1980s failed, in general, one of their main
purposes: to produce energy at a competitive price. They were not as reliable as expected, with short
life periods due to several failures. In (Johnson, 2001), it is referred the fact that smaller turbines (in
the 100 kW range) could be built at lower costs and with better performance than the larges turbines as
one of the reasons to the end of the American program. In fact, the seek for larger and powerful wind
turbines underestimated the difficulty of building reliable machines of this size. The Growian wind
turbine is an example in which the size of the turbine is too large for the time in which it was built.
On the other hand, for the first time in the history of the wind energy, there was a consistent
experimental and theoretical knowledge of wind turbines. This knowledge was fundamental for the
beginning of the commercial market of wind turbines as it is known today.
In hindsight, it is possible to evidence three distinct groups of approaches on the structural and control
level over the evolutionary history of wind turbines (Thresher and Dodge, 1998):
o Turbines designed to withstand high wind loads. These models were considerably robust, with
low rated power and were optimized for reliability. An example of this philosophy was the
Gedser wind turbine;
o Turbines designed to be compliant and shed loads. These models were optimized for
performance. Their design had an important concern with the reduction of weight and with
the optimization of the aerodynamic profile of the blades. An example of this design concept
was the W-34 wind turbine;
o Turbines designed to manage loads mechanically and/ or electrically. These turbines were
optimized for control, with several mechanical and electrical innovations (such as hinged
blade connections and variable speed generators). An example of this concept was the Smith-
Putnam turbine.
The first approach was indubitably the most durable and reliable. However, the increase of knowledge
in the industry has led to less conservative designs, with the aim of weight reduction (to reduce cost)
and control (to optimize the production). Thus, a joint of the second and third approaches is currently
the main driver of the design concepts.
An illustration of the principal projects held during this period of development is shown in Figure
2.23. The evolution of the dimension of the wind turbines is notorious.
41
Figure 2.23 – Main projects developed during the 1970s and 1980s (adapted from (Spera, 1995; Gasch and Twele, 2012))
Height [m]
150
(Wind Direction)
125
100
100
79.2
75
75 78
97.5
54 91.4
50
61
38
25
Name MOD-0 Tvind MOD-1 MOD-2 WTS-4 WTS-3 Growian WTS-75 MOD-5B
From Windmills to the Modern Wind Turbines
Year 1975 1975 1979 1980 1982 1982 1982 1983 1987
Rated Power
0.20 2.00 2.00 2.50 4.00 3.00 3.00 2.00 3.20
[MW]
42
Chapter 2
1
In this study, the manufacturers considered were: Alstom, AMSC, Darwin, Enercon, Gamesa, GE Energy, GoldWind, MY
Wind, Nordex, Samsung, Senvion, Siemens, Sinovel, Suzlon and Vestas.
43
From Windmills to the Modern Wind Turbines
120
80
Senvion MM82
60
Figure 2.24 – Rated power and rotor diameter of onshore wind turbines
44
Chapter 2
45
5
From Windmills to the Modern Wind Turbines
175
Rotor Diameter [m]
150
125
100
75
Vestas V90-3.0MW
50
2 3 4 5 6 7 8 9 10 11
Rated power [MW]
Figure 2.26 – Rated power and rotor diameter of offshore wind turbines
Another important aspect in the offshore development is related to the distance to the shore and water
depth at which the wind farms are installed. Figure 2.27 presents the average distance to shore and
water depth of European wind farms, alongside with their dimension (defined by the size of the
46
Chapter 2
circles). Over the years, the distance to shore has increased, as well as the water depth of the sites of the
projects. The average distance to shore was 32.9 km and the average water depth of wind farms was
22.4 m, at the end of 2014 (European Wind Energy Association (EWEA), 2015a). Looking at Figure
2.27, it is possible to assume that the trend of installation of wind farms away from the shore will
continue.
120
Online
100
Consented
80
Distance to shore [km]
Under
construction
60
40
20
0 10 20 30 40 50 60
Water depth [m]
Figure 2.27 – Average water depth and distance to shore of online, under construction and consented European wind
farms, at the end of 2014 (adapted from (European Wind Energy Association (EWEA), 2015a))
The main design drivers for the future of offshore wind turbines will be the cost reduction and the
implementation of turbines at higher water depths. Furthermore, the increase of the turbine power
output and of their reliability will continue, alongside with a better understanding of the influence of
the environment on the structure of the wind turbine.
Currently, the levelized cost of offshore energy is slightly higher than twice the onshore cost (World
Energy Council (WEC), 2013; Bloomberg New Energy Finance, 2015; U.S Energy Information
Administration, 2015). In order to become viable, the capital and operational expenditure of offshore
investments need to be reduced. A part of the reduction of the initial investment is achieved with the
decrease of the structure weight (Butterfield, Musial et al., 2007). With light-weight structures, the
loads tend to decrease through all the structure, including in the foundation. Some new and more
expensive materials might be used for this purpose, such as, in special cases, carbon in the blades
(Butterfield, Musial et al., 2007). However, less material (and then less stiffness and mass) leads to a
more flexible structure which, considering the dynamic loads, requires a thorough understanding of
the dynamic behaviour of the structural system.
Another possibility for weight reduction is the adoption of two-bladed rotors (Butterfield, Musial et al.,
2005). This solution was almost abandoned for the onshore turbines due to lower yield and for
aesthetic reasons, as well as some technical problems that were not solved in the past - like the one
described for the Jacobs wind turbine (Butterfield, Musial et al., 2007). However, the reduction of one
blade (with the consequent weight decrease) could be balanced by the increase in blade tip speed of the
offshore turbines. In addition, the higher rotor speeds lead to a lower input torque and lower gear
ratios, which permit to reduce some mechanical components located in the nacelle. The increase in
rotor tip speed could also lead to the utilization of direct-drive generators (drivetrains without
gearboxes) (Musial and Ram, 2010).
47
From Windmills to the Modern Wind Turbines
In the future, another major change may take place for the offshore turbines. The continuous
increasing of rotor diameters may lead to a definition of the blade design by a deflection limit in
upwind rotors, instead of a load limit criterion (Butterfield, Musial et al., 2007). This is due to the
increased levels of strength and flexibility achieved in blades. For this reason, downwind rotors may be
an option for future offshore wind turbines because, contrary to the upwind solution, the blades
deflect away from the tower. The downwind solution was abandoned in the past due to vibrations and
noise problems (Butterfield, Musial et al., 2007) with the blades passing through the tower shadow.
Nevertheless, for large distances from the shore, noise disturbances are not a problem.
Some other perspectives for future developments of wind turbines are presented in (European
Commission, 2011). In the report of the UpWind program, some considerations are made for the
design of a giant 20 MW wind turbine. Although the program was not exclusively for the offshore
market, the dimensions of the turbines under analysis make them suitable for the implementation at
offshore locations. Among other considerations, some important aspects are referred for the feasibility
of turbines with this dimension:
o Reduction of the magnitude of fatigue loads. For this, advanced rotor control strategies are
required (the concept of “smart turbine” is introduced);
o Implementation of control systems for detection and evaluation of upcoming wind loads
(gusts or vortex) with the aim of load alleviation (such as LIDAR systems mounted at the
nacelle);
o Individual pitch control of each blade;
o Application of materials with lower mass to strength ratio;
o Introduction of monitoring technologies to assess the structural integrity of the system;
o Due to transportation issues, the blades might have to be splitted into two pieces since the
rotor of the studied 20 MW wind turbine has a diameter of 252 m. For this reason, this joint
might be an important focus of attention;
o The need for optimization of the offshore substructure (foundation and transition piece).
The offshore substructure is another important topic for the future of offshore wind turbines.
Nowadays there is a large variety of commercially implemented, tested and concept designs (see
Chapter 3). The future trend, as stated earlier, is to increase the distance from the shore. In this
situation, due to a larger water depth, floating structures seem to be the most suitable option. Until
today, a few number of experimental projects have been deployed. For the next years, many floating
wind turbine projects are schedule to start in several countries (Main(e) International Consulting LLC,
2012).
48
Chapter 2
49
From Windmills to the Modern Wind Turbines
250000
Latin America & Caribbean
[2.3 %]
200000
Europe [36.3 %]
150000
50000
Africa & Middle East [0.7 %]
0
2005 2006 2007 2008 2009 2010 2011 2012 2013 2014
Figure 2.28 – Cumulative installed capacity (the share at the end of 2014 results is presented in brackets) (EWEA 2007,
2008, 2009, 2010, 2011b, 2012b, 2013c, 2014b, 2015b) (GWEC 2007, 2008, 2009, 2010, 2011, 2012, 2013, 2014, 2015)
Top 10 Cumulative Capacity (2014)
Brazil
Rest of the
France Italy 2% world China
2% 2% 16% 31%
Canada
3%
UK
3%
India USA
6% Germany 18%
Spain 11%
6%
Figure 2.29 – Top 10 Countries in cumulative installed capacity in the end of 2014 (European Wind Energy Association
(EWEA), 2015b; Global Wind Energy Council (GWEC), 2015)
On the other hand, European investments stagnated during the last years which can be justified by the
financial crisis installed in this continent (Figure 2.30). In the opposite side, China was responsible,
during 2013 and 2014, by more than 45 % of the new installations (Figure 2.31). These recent
investments led China to a position of market domination.
50
Chapter 2
50000
North America [14.3 %]
40000
Latin America & Caribbean
[7.3 %]
30000
Europe [25.0 %]
20000
Asia [50.5 %]
10000
Africa & Middle East [1.8 %]
0
2006 2007 2008 2009 2010 2011 2012 2013 2014
Figure 2.30 – Annual installed capacity (the share at the end of 2014 results is presented in brackets) (EWEA 2007, 2008,
2009, 2010, 2011b, 2012b, 2013c, 2014b, 2015b) (GWEC 2007, 2008, 2009, 2010, 2011, 2012, 2013, 2014, 2015)
Top 10 New Installed Capacity (2014)
France
2% Turkey
2% Rest of the
Sweden world
2% 13%
UK China
Canada
3% 45%
4%
India
5%
Brazil
USA
5%
9% Germany
10%
Figure 2.31 – Top 10 Countries in new installed capacity in the 2014 (European Wind Energy Association (EWEA), 2015b;
Global Wind Energy Council (GWEC), 2015)
In Portugal, large investments were done in wind energy since the first wind farm was commissioned
at Madeira in 1986. Portugal is the 12th country in the world with largest installed power and the 8th in
Europe with a cumulative installed capacity of 4 914 MW (European Wind Energy Association
(EWEA), 2015b; Global Wind Energy Council (GWEC), 2015) (Figure 2.32).
However, the investments have been declining in recent years. This situation may be explained by the
financial crisis, by the minor incentives given by the Government and by some limits imposed for
penetration in the national electrical system.
51
From Windmills to the Modern Wind Turbines
5.000
Cummulative
installed Capacity
Power output [MW]
4.000
3.000
Annual installed
capacity
2.000
1.000
?
0
2005 2006 2007 2008 2009 2010 2011 2012 2013 2014
Figure 2.32 – Annual cumulative and installed capacity in Portugal (since 2005) (EWEA 2007, 2008, 2009, 2010, 2011b,
2012b, 2013c, 2014b, 2015b)
Figure 2.33 shows the geographical distribution of wind farms in the continental territory of Portugal.
As can be observed, the major part of wind farms is located in the north and centre of the country,
mainly in interior highlands.
Wind Farms
Total Power Output [MW]
[
[0.5 – 1.9]
[2.0 – 9.9]
[10.0 – 24.9]
[25.0 – 49.9]
>50
Figure 2.33 – Installed wind farms in Portugal at the end of 2014 (Institute of Mechanical Engineering and Industrial
Management (INEGI) and Portuguese Renewable Energy Association (APREN), 2015)
52
Chapter 2
Norway [0.02 %]
8.000 Spain [0.06 %]
7.000 Portugal [0.06 %]
Norway [0.02 %]
6.000 Ireland [0.29 %]
5.000 Japan [0.57 %]
Finland [0.57 %]
4.000
Sweden [2.42 %]
3.000 Belgium [8.13 %]
2.000 Germany [11.98 %]
China [7.51 %]
1.000
Netherlands [2.82 %]
0 Denmark [14.51 %]
2009 2010 2011 2012 2013 2014 UK [51.32 %]
Figure 2.34 – Annual cumulative capacity of offshore wind turbines .The share values in brackets are referred to the end
of 2014 (EWEA 2011a, 2012a, 2013b, 2014a, 2015a) (GWEC 2012, 2013, 2014, 2015)
In today’s market, the UK is clearly dominant with more than 50 % of the entire global capacity. The
UK has been doing a large investment in the offshore market in the last years, as can be seen in Figure
2.35. Portugal has a small portion of the entire market with just one 2.0 MW experimental floating
offshore wind turbine.
Anual Installed Power US [-]
2000 Korea [-]
Portugal [-]
1750 Norway [-]
Annual Installed Power [MW]
Spain [-]
1500 Portugal [-]
Norway [-]
1250
Ireland [-]
1000 Japan [-]
Finland [0.01 %]
750 Sweden [-]
Belgium [8.76 %]
500 Germany [32.84 %]
China [7.89 %]
250
Netherlands [-]
0 Denmark [-]
2010 2011 2012 2013 2014 UK [50.51 %]
Figure 2.35 – Annual installed capacity of offshore wind turbines (data from 2013 is only referred to European countries).
The share values in brackets are referred to the end of 2012 (EWEA 2011a, 2012a, 2013b, 2014a, 2015a) (GWEC 2012,
2013, 2014, 2015)
The dominance of the offshore market by the Northern countries of Europe is a clear evidence. For
this fact, the water depth conditions present in these countries should be considered as an important
driver. Analysing Table 2.1, it is evident the large portion of shallow waters present in the north of
Europe, in opposition to the situation at the south regions, where the water depth increases rapidly
53
From Windmills to the Modern Wind Turbines
with the distance from the coast. In addition, it is also possible to note the apparent good conditions
for the USA to install offshore wind farms. However, in the last years, investments in this area were not
considered a priority by this country.
Table 2.1 – Percentage of water depths in different regions up to 100 km offshore (Henderson, 2003)
Water Depth
Region
< 25 [m] 25 – 50 [m] 50 – 100 [m] 100 – 300 [m]
North Europe 21 26 32 20
South Europe 16 11 23 49
Japan 22 9 18 51
USA 50 26 13 11
54
Chapter 2
55
From Windmills to the Modern Wind Turbines
56
Chapter 3
3
BASICS OF WIND TURBINES
57
Basics of Wind Turbines
Blade Blade
Wind Direction
Hub
Nacelle Nacelle
Rotor
Tower
Tower
Boat
Landing
Transition Support
(Water Level)
Piece Structure
J-tube
Scour Foundation
Protection Foundation
a) b)
Figure 3.1 – Main wind turbine components of horizontal-axis upwind wind turbines: a) offshore; b) onshore
Flapwise Edgewise
Side-Side Fore-Aft
Figure 3.2 – Terminology of the principal axes of translation and rotation of a wind turbine structure
58
Chapter 3
3.1.1.1 Rotor
The rotor is the key component of a wind turbine, converting the kinetic energy of the wind into
mechanical energy. It is composed by the blades which are connected to the hub.
The design of a blade is one of the most complex tasks in the development of a wind turbine. Blades
have to be designed in order to optimize the energy extraction from the wind flow. However, at the
same time, the progressive increase in the power output of the turbines demands increasingly larger
blades (with models of 83.5 m already being produced (Quilter, 2013)).
For these reasons, blades present complex cross sections (named airfoils) which vary along their length
(Figure 3.3). In addition, blades have to be sufficiently light, stiff and strong, at the same time, to not
contribute with enormous gravitational loads (for the blade itself and for the tower), to not deflect too
much (which would drive to loss of efficiency and, even more important, could lead to a collision with
the tower) and to resist to a very high number of cycles during their lifetime (of around 107 to 108
(Hau, 2006)). In order to accomplish all these prerequisites, composite materials are used (usually
epoxy resins with glass fibres, although carbon fibres may be also used in the blade composition).
In Figure 3.4 two common solutions for the structure of a blade are presented.
Considering the rotor position, wind turbines may have a wide variety of configurations. As previously
referred, the rotor can be placed in front (upwind) or behind (downwind) the tower. This latter
solution is rarely used today because of noise, vibrations and efficiency problems. Due to the presence
of the tower upstream of the rotor, the flow is disturbed before it reaches the blades and, every time a
blades pass behind the tower, the momentary absence of aerodynamic force leads to additional
vibrations (shadow effects). In addition, large downwind rotors also need a yaw system (which is not
necessary in small downwind turbines). Several examples of downwind wind turbines were presented
in Chapter 2.
The number of blades is another aspect that contributes for the variety of rotor configurations.
Historically, wind turbines have one, two or three blades. The difference between two or three-bladed
rotors implies, besides differences in tip-speed and efficiency, different vibration problems. Figure 3.5
presents an example of the yaw moment created by an asymmetrical wind flow on wind turbines with
different rotor solutions (Hau, 2006). As can be seen, one and two-bladed rotor wind turbines suffer
from a noteworthy time-varying yaw moment while this excitation is almost undetected with rotors
with three or more blades. This effect is due to the variation of the rotor moment of inertia with the
blades revolution. Whereas in one or two-bladed rotors, this asymmetry in the moment of inertia is
59
Basics of Wind Turbines
relevant and leads to a pulsating load profile during a revolution (Hau, 2006), three or more blades
confer an almost symmetrical moment of inertia to the rotor.
Shear web
Trailing Edge
Leading Edge
a)
Box girder
Trailing Edge
Leading Edge
b)
Figure 3.4 – Two common design concepts for blade cross sections with different load-carrying solutions (adapted from
(Tong, 2010)): a) shear web; b) box girder
3000
2000
1 blade
1000
Yaw Moment [kNm]
2 blade
0 3 blade
4 blade
-1000
-2000
-3000
0 90 270 180 360
Angle of Rotation [º]
Figure 3.5 – Yaw moment caused by asymmetrical wind flow on wind turbines with different number of blades (adapted
from (Hau, 2006))
With the aim of alleviating these loads created by two-bladed rotors, some solutions were developed.
These solutions focus on the flexibility of the connections between blades and the hub (Figure 3.6).
The simplest solution is the rigid hub connection. This solution connects the blades to the hub at fixed
positions and, consequently, the loads are directly transmitted to the drive train by the hub. This is the
common solution for three-bladed rotors and for some rotors with two blades (such as the 1983 WTS-
75 Nasudden, see Figure 2.19)).
A more flexible solution is the teetering hub. This hub connection allows the rotor to teeter about the
rotor shaft due to a hinge located at the hub. This solution permits to balance out some asymmetrical
loads. This solution is intended for two-bladed rotors (and was used in the 1982 Growian wind
turbine, see Figure 2.21).
60
Chapter 3
Another flexible solution is the flapping hinge rotor. In this solution, a hinge is installed in each blade
allowing the individual rotation of the blades (the 1941 Smith-Putnam wind turbine used this solution,
see Figure 2.7)).
3.1.1.2 Nacelle
The nacelle is the component of a wind turbine which encloses the mechanical and some electrical
elements (Figure 3.7).
In the nacelle, the rotor shaft (or main shaft) receives the torque (mechanical energy) from the hub. In
nowadays most common turbines, this shaft is connected to a gear box. The gear box increases the
rotational speed from the rotor shaft to a more convenient speed to the generator. The connection
between the gearbox and the generator is made by the generator drive shaft. The generator produces
electric energy from the mechanical energy. There is also a mechanical brake with the aim of stopping
the turbine when necessary and a yaw system that mechanically orients the nacelle to the wind
direction. The components of the nacelle are supported by the bedplate.
Although the previous description refers to the most common nacelle layout, there are also some
different concepts, such as the absence of a gear box (direct-driven generators).
Bearings
Bedplate
Yaw System Rotor Hub
a) b)
Figure 3.7 – a) Main components inside the nacelle. b) Detail of the drive train (adapted from (Hau, 2006))
61
Basics of Wind Turbines
3.1.1.3 Tower
Wind turbine tower elevates the rotor to the desire height. Since the wind flow presents higher speed
and lower turbulence at higher heights, the towers tend to be very tall. In addition, wind turbine
towers are also very slender which, alongside with the rotor motion, leads to complex dynamic
problems.
The more common towers are of four types: steel tubular, steel lattice, concrete tubular and hybrid
solutions. There are also some special designs, such as guyed supported towers. However, since these
are almost not used nowadays, they are not covered in this work.
Prestressing
tendons
Steel
Concrete
a) b) c) d)
Figure 3.8 – Various wind turbine tower types: a) steel tubular; b) lattice; c) concrete tubular; d) hybrid solution (adapted
from (Tricklebank, Halberstadt et al., 2007))
62
Chapter 3
introduces some difficulties concerning the stiffness of the tower. For that reason, this solution is not
used in some models due to the required structural stiffness that a steel tubular tower cannot deliver.
Hybrid Tower
A compromise solution is obtained with tubular hybrid towers. This solution is characterized by a
concrete tower for the lower part of the tower (usually mixed on-site), while the upper portion is made
of steel. Thus, it is possible to increase the stiffness of the structural system with a short period of
concrete curing and without transportation of large steel components. This is a solution mainly for
high wind turbine towers (100 m or higher).
63
Basics of Wind Turbines
transition piece
(includ. access platform,
cable ducts, ladder,
grounted joint add-ons)
monopile
submarine cable
scour protection
Figure 3.9 – Grouted joint for the connection between the foundation (monopile) and the transition piece. In this
situation, the verticality is corrected by the connection (Gasch and Twele, 2012)
3.1.1.5 Foundation
The foundation is the component that best distinguishes onshore and offshore wind turbine
structures. While onshore turbine foundations follow traditional in-land civil engineering solutions,
offshore models have been progressively adopting solutions from oil and gas offshore industry.
However, both onshore and offshore foundations have the same function: to ensure stability and drive
the loads from the structure to the ground, which means, to prevent the structure from sinking due to
gravitational loads, from sliding due to horizontal loads and from fall over due to base overturning
moments. Besides this, it is important to bear in mind that the definition and implementation of the
designed foundation will influence the dynamic behaviour of the entire structural.
Onshore Solutions
The foundations for onshore wind turbines employ well-known and controlled solutions. There are
mainly two possibilities: slab foundations or pile foundations (Figure 3.10).
64
Chapter 3
a) b)
Figure 3.10 – Examples of foundations for onshore wind turbines: a) slab foundation; b) pile foundation
Slab foundations are used when the ground near the top is of good quality. They consist on reinforced
concrete slabs which usually have a polygonal geometry (such as an octagonal shape) with a tapered or
constant thickness.
This kind of foundation resists to the overturning moment with the eccentricity of the reaction. The
vertical and horizontal loads are resisted by the dispersion of the load through the foundation area and
by the friction between the ground and the foundation, respectively.
Pile foundations are used for weaker soils. They are constituted by a pile cap from which a certain
number of piles are connected and extended until a good soil layer is reached. In this situation, the
loads are absorbed by the soil through axial and lateral pile resistance against the soil. Once again,
reinforced concrete is used in this solution.
Offshore Solutions
In contrast to onshore turbines, there is the need for diversified offshore solutions, depending on the
water depth. In addition, these solutions are considered far more complex than for onshore.
Offshore solutions for wind turbines can be divided in three categories: for shallow waters (up to 25 –
30 m), for transitional waters (from around 30 to around 60 m) and for deep waters (more than 60 m).
The most appropriate solutions for each water depth are presented in Table 3.1.
65
Basics of Wind TTurbines
T
Table 3.1 – Offsshore foundatio
on solutions co
oncerning wate
er depth
Gravity-b
based ✓
Monop
pile ✓
Tripod ✓
Jacke
et ✓
Tripile ✓
ng
Floatin ✓
Gravity-bassed foundatio on is the second most ussed solution for current offshore
o winnd turbine prrojects
(Figure 3.111). It consistss on a caisson
n (usually maade of concrete) which iss filled with m
material to acchieve
the requiredd weight. Th herefore, a large footpriint is used to
o absorb mo oments due to environm mental
loads. The ccaisson is faabricated on land and thhen transportted to the sitte, where it is submergeed and
filled with ballast. For the operatio on on the siite, heavy veessels are required to seet the found dation.
Besides the need for heaavy machinerry, another ddisadvantage is the necesssary preparattion of the seeabed.
This operatiion ensures compaction
c and homogeenization of the
t soil to prevent unevenn settling.
This solutioon is considered, with reggard to the viibrational ch
haracteristics, as “stiff”. FFor this reaso
on, the
aerodynamiic damping of o the rotor do not contrribute in largge measure to t alleviate thhe response of the
structure duue to the dyn
namic loads (see
( section 44.5) (Hau, 20006).
Gravity fouundations aree indicated fo
or very shalllow waters (aaround up too 10 m – e.gg. the Nysted
d wind
farm in Den nmark). How wever, there are examplees of implemmentation at deeper
d waterrs (around 27 m –
e.g. the Thoornton Bank wind farm, ini Belgium (T Thomsen, Fo orsberg et al., 2007)).
Tower
(Se
ea Level)
(SSeabed)
66
Chapter 3
Monoopile foundaation is anotther alternattive solution for shallow w waters andd is, nowadayys, the mostt
used solution in offshore
o wind turbines (FFigure 3.12).. It consists on
o a free-stannding steel pipe
p which iss
driven or drilled (depending
( on
o soil qualitty) into the seabed
s by meeans of speciial and heavyy equipment..
In sittuations in which the monopile
m is implementeed by driving, hydraulicc hammers are a used forr
penettration. Sincce the pressu ure introducced by the hammers
h is very high, appurtenancces (such ass
outsidde ladders an nd the j-tubee) cannot bee installed directly on thee monopile. FFor this reasson, they aree
installled on the trransition piece. The prevvious preparaation of the seabed is not necessary foor monopiless
although significaant boulders should be avvoided on the foundation n area.
Contrary to the gravity-based
g d solution, m
monopile fou
undations arre consideredd as “soft”, in
i respect off
their dynamic beh
haviour (Hauu, 2006).
Monoopile foundaations are useed in deeper waters than gravity-baseed solutions. However, a water depth
h
of aroound 25 m to
t 30 m is a consensual llimit for the implementaation of this ffoundation. This type off
founddation was used
u in the Horns Rev win nd farm (in Denmark).
D
Tower
Transition Piiece
(Sea Levvel)
(Seabeed)
Figure 3.12
2 – Monopile fo
oundation (pho
oto (LORC, 2012
2b) and illustraation)
67
7
Basics of Wind TTurbines
This type of foundatiion is still in an expeerimental ph hase. There are alreadyy some exaamples
implementeed (in Fredeerikshavn, Denmark) (Ibbsen, Liingaaard et al., 20
005) and moore recently at the
Dogger Bannk (UK) (C Carbon Trust, 2013)). H
However a project
p was cancelled ddue to installlation
problems in
n 2005 (Enerccon GmbH, 2005).
Tower
Transition Piece
(Sea Level)
(Seabed)
Water deptths of around 25 m – 30 0 m are the limit of currrent commo on commerccial offshore wind
turbines. Beeyond this liimit, the app
plicability off the presentted solutionss becomes veery expensivve and
demanding on material weight and on o machinerry for implem mentation. Fo or this reasonn, new found
dation
technologiees are under development
d t nowadays ((for transitional and deep p waters).
Tripod founndation is on
ne of the alteernatives for transitional waters. It is an expansioon of the mon
nopile
concept witth a larger footprint
f (Figure 3.14). T Tripods connsist on a strructure withh three legs which
w
diverge from
m a single no
ode. These thhree points arre anchored in the seabed d.
The tripod ggeometry alllows this solu
ution to be liight-weight and
a “stiff”. Since the loadds are mainlyy axial
in the legs, scour protection is geneerally not neeeded (Tongg, 2010). In addition,
a onlly reduced seabed
s
prearrangem ment is requ
uired. On thee other handd, these stru uctures are complex to bbuild (leading to a
high producction expendditure) and difficult
d to trransport (Haau, 2006). Also, the mainn joint of thee three
legs is comp
plex and a po
otential focuss of fatigue pproblems.
Tripod foun ndations are suitable for water depthhs up to arou
und 50 m – 60 m. This kiind of foundations
has been alrready implem
mented on th he Alpha Ven ntus wind farrm, in Germaany (Bartschh, 2012).
68
Chapter 3
Tower
Transition Piece
(SSea Level)
(Seabed)
Figure 3.14
4 – Tripod foun
ndation (photo (Wikipedia, 2012) and illustraation)
69
9
Basics of Wind TTurbines
Tower
Transition Pie
ece
(Sea Level)
L
(Se
eabed)
Fig
gure 3.15 – Jacket foundation
n (photo (LORC
C, 2012a) and illustration)
70
Chapter 3
Tower
(
(Sea Level)
(Seabed)
F
Figure 3.16 – Trripile foundatio
on (photo (BARD Engineering GmbH) and illuustration)
As reeferred in Ch hapter 2, onee of the todaay’s main fiellds of develo opment in thhe offshore wind
w turbinee
indusstry is the floating strructure for deep waterr depths. Nowadays, N thhere are already somee
experrimental pro ojects under developmen nt, however there
t is still a long way of developmment for thee
feasibbility of commmercial offsh
hore wind tuurbines basedd on floating foundationss.
Besid
des the possib
bility of intro
oduction of wwind turbinees at deep water depths, floating con ncepts enablee
a maass-productioon of structtures since they can achieve a high independdence from the seabed d
condiitions. Anoth
her advantagge of floatingg platforms iss the possibility of being ttowed to their location.
The design of floating
f stru
uctures folloows mainly three diffeerent conceppts to achieeve stabilityy
(Buttterfield, Musiial et al., 2005):
o Ballast – This kind ofo solution uuses the weiight of ballasst below a ceentral buoyaancy tank to o
create a righting
r momment and higgh inertial reesistance to pitch
p and rolll. The elongaated shape off
the struccture helps to
o minimize tthe heave mo otion. This so olution has bbeen used in the offshoree
oil indusstry for manyy years (Musiial, Butterfield et al., 2004
4);
o Mooringg Lines - Plattforms usingg this technollogy achieve stability throough the usee of mooringg
lines und
der tension;
o Buoyanccy – This lastt concept rellies on a disttributed buoy
yancy, achievved through the use of a
weighted
d plane area for
f righting m moment (Bu utterfield, Mu
usial et al., 20005).
Althoough each deescribed conncept presentts a differentt physical principle to acchieve stabiliity, in realityy
floatiing platforms usually preesent a hybriid system. Commonly, so olutions are designed based on all off
71
Basics of Wind Turbines
the three concepts, although generally relying on one primary source for stability (Butterfield, Musial
et al., 2005).
A representative design of each of the listed concepts is presented in Figure 3.17.
Tower
(Sea Level) (Sea Level)
(Sea Level)
Tower Tower
Mooring
Mooring Lines
Mooring
Lines Lines
a) b) c)
Figure 3.17 – Floating solutions for offshore wind turbines(adapted from (Butterfield, Musial et al., 2005)): a) ballast; b)
mooring lines; c) buoyancy
72
Chapter 3
The NREL study present in (Tegen, Hand et al., 2012) describes some costs related to onshore models
of U.S. wind projects during 2010. Figure 3.18 presents the principal results of this study, namely the
capital cost share of the principal components of a wind turbine located in wind farm. These costs are
grouped in larger groups, namely:
o Turbine – comprises the wind turbine itself, which includes the tower, nacelle and rotor
components (Blanco, 2009);
o Balance of plant – including all the infrastructure except the turbine (foundations, buildings,
roads, among others) (Department of Energy & Climate Change, 2010);
o Installation and commissioning – including the installation and commissioning of the turbine
and balance of plant (Department of Energy & Climate Change, 2010);
o Development and consent – includes the multifaceted process of taking a wind farm from
inception through to the point of financial close or commitment to build, depending on the
contracting model, including several studies and contracts (Department of Energy & Climate
Change, 2010).
73
Basics of Wind Turbines
Engineering
Other Costs Foundations
& permits
9% 14%
6%
Turbine
transportation
9%
Balance of Electrical
Plant interface and
23% connections
37%
Turbine
Roads & civil
68%
work
20%
Turbine
Installed Capital Costs
Blades
14%
Tower
22%
Hub
Control, safety 4% Pitch mechanism &
system, and condition bearings
monitoring 4%
2%
Low-speed shaft
Nacelle cover 3%
1% Bearings
Electrical 2%
Hydraulic, cooling
connections
system
6% Gearbox
1%
11%
Main frame
10%
Variable-
speed Generator
electronics 7% Mechanical brake,
9% High sped shaft
Yaw drive & bearing coupling
2% 0%
Figure 3.18 – Installed capital costs for onshore wind turbine farms (Tegen, Hand et al., 2012)
The main installed capital costs of offshore wind turbines are shown in Figure 3.19 (Department of
Energy & Climate Change, 2010). These values are adjusted in order to give representative results for a
500 MW+ wind farm of approximately 5 MW turbines on jacket foundations in typical UK east coast
conditions around 80 km from shore in water depth of 30 m and considering market dynamics
prevailing at the start of 2010.
74
Chapter 3
Balance of
Plant
31% Turbine
39%
Development
Installation and
and Consent
Commissioning
4%
26%
Tower Blades
15% 18% Offshore
Other Costs Hub Substations
12% Assembly 23%
9%
Foundations
Cables
Electrical 52%
16%
System Gearbox
20% 26%
Onsohre
Electrical
9%
Figure 3.19 – Installed capital costs for offshore wind turbine farms (Department of Energy & Climate Change, 2010)
Since offshore models are largely dependent on the water depth of implementation, a qualitative
representation of costs for different substructures is presented in Figure 3.20 (Musial and Ram, 2010).
Tripods,
Jackets,
Trusses Floating
Structures
Substructure Cost
Monopiles,
Gravity Foundations
Shallow-Water Transitional Deep-Water
Technology Technology Technology
0
0 20 40 60 80 100 120 140 160
Figure 3.20 – Cost of offshore wind turbine substructures with water depth (adapted from (Musial and Ram, 2010))
75
Basics of Wind Turbines
As can be seen from the previous figures, the main costs come from the turbine group. However, for
offshore turbines, the costs are more evenly distributed mainly due to the increased cost of
foundations and installation of the turbines (Department of Energy & Climate Change, 2010).
Another interesting aspect is the high relative cost importance of the tower, blades and gearbox. This is
highlighted in Figure 3.21 (capital costs of a 5.0 MW turbine from Senvion).
Turbine Components
Brake System
1%
Cables
1% Screws
1% Other Costs
Nacelle Housing 12% Tower
1% 26%
Transformer
4%
Power Converter
5%
Pitch System
3% Rotor Rotor Hub
Bearings 1%
Yaw System 1%
1% Generator
3% Main Frame Main Shaft
3% 2%
Gearbox
13%
Rotor Blades
22%
Figure 3.21 – Share of capital costs of the Senvion 5M wind turbine (Wind Directions, January/ February 2007)
76
Chapter 3
scheduled activities defined according to statistical models based on the historic of failures. On
the other hand, predictive maintenance is a strategy based on the assessment of the
components health through sensors installed in the components/ structure. The monitoring
system developed and described in this work falls in this category. This is the strategy expected
for future large onshore and offshore wind turbines.
Modern wind turbines achieve a high availability1, of around 97 %, when the maturity and teething
problems of the turbines are solved (Harman, Walker et al., 2008). However, several failures still occur,
implying unscheduled down times (Faulstich, 2010).
The UpWind project (Faulstich, 2010) analysed wind turbine failures in the German market between
1989 and 2006. It presents two major groups of failure causes:
o External causes (such as lightning, grid outage, icing, storm). These failures mostly affect
electric subassemblies and occur more frequently. However, the downtime associated with
each occurrence is usually short. For these cases, design optimization is suggested as solution
for prevention;
o Operation causes (such as wear out and relaxation). This type of failure mainly affects
mechanical components and, although seldom, usually implies large downtimes. An
assessment of the condition of the components through the use of sensors is recommended for
prevention of these failures.
Among different components, electrical systems represent the most common failures for onshore wind
turbines (Figure 3.22). However, rotor blades (including pitch system), the yaw system, the gearbox
and also electrical systems are the ones which imply longer downtime, according to Figure 3.23
(Faulstich, 2010; Wilkinson and Hendriks, 2011).
40 %
Power Module
Contribution to overall failure rate (failures/ turbine/ year) [%]
35 %
30 %
Rotor Module
25%
20 % Control &
Comms
15 %
Nacelle
10 %
Drive Train Auxiliary System
5% Structure
0%
HYDRAULIC SYSTEM
TOP
UNKNOWN
ANCILLARY EQUIPMENT
UNKNOWN
GENERATOR SILENT BLOCKS
eLECTRICAL PROTECTION & SAFETY
SERVICE CRANE
GROUNDING
FIREFIGHTING SYSTEM
ELECTRICAL AUXILIARY CABLING
FOUNDATIONS
WIND FARM SYSTEM
SENSORS
COMMUNICATION SYSTEM
SAFETY CHAIN
CONTROLLER H/W
CONTROLLER S/W
HUMAN & OPERATIONAL SAFETY ...
UNKNOWN
GEARBOX ASSEMBLY
MECHANICAL BRAKE
MAIN SHAFT
COOLING SYSTEM
WTG METEOROLOGICAL STATION
LIGHTNING AND POWER POINTS
GROUND
LIFT
ELECTRICAL CABINETS
UPS CABINET
BEACON
UNKNOWN
HUB CABINET
TOWER
COMMON FACILITIES
UNKNOWN
DATA LOGGER
CONDITION SENSORS & CABLES
PROTOCOL ADAPTER CARD FOR...
FREQUENCY CONVERTER
GENERATOR ASSEMBLY
LV SWITCH GEAR
MV SWITCH GEAR
TRANSFORMER
POWERFEEDER CABLES
UNKNOWN
POWER CABINET
PORTECTION CABINET
PITCH SYSTEM
BLADES
HUB
SLIPRINGS
HUB COVER
UNKNOWN
BLADE BEARINGS
SENSORS
YAW SYSTEM
NCAELLE COVER
NACELLE BEDPLATE
Figure 3.22 – Normalized failure rate of sub-systems and assemblies for turbines of multiple manufacturers (for onshore
wind turbines) (adapted from (Wilkinson and Hendriks, 2011))
1
Availability time is defined as the ratio between the time during which the wind turbine is available to produce (with
none component failure) and the entire time under analysis.
77
Basics of Wind Turbines
40 %
Power Module
35 %
Contribution to overall downtime (hours lost) [%]
Rotor Module
30 %
25%
20 %
15 %
Control &
Comms
10 %
Nacelle Drive Train
5% Auxiliary System
Structure
0%
HYDRAULIC SYSTEM
TOP
UNKNOWN
ANCILLARY EQUIPMENT
UNKNOWN
GENERATOR SILENT BLOCKS
eLECTRICAL PROTECTION & SAFETY
SERVICE CRANE
GROUNDING
FIREFIGHTING SYSTEM
ELECTRICAL AUXILIARY CABLING
FOUNDATIONS
WIND FARM SYSTEM
SENSORS
COMMUNICATION SYSTEM
SAFETY CHAIN
CONTROLLER H/W
CONTROLLER S/W
HUMAN & OPERATIONAL SAFETY ...
UNKNOWN
GEARBOX ASSEMBLY
MECHANICAL BRAKE
MAIN SHAFT
COOLING SYSTEM
WTG METEOROLOGICAL STATION
LIGHTNING AND POWER POINTS
GROUND
LIFT
ELECTRICAL CABINETS
UPS CABINET
BEACON
UNKNOWN
HUB CABINET
TOWER
COMMON FACILITIES
UNKNOWN
DATA LOGGER
CONDITION SENSORS & CABLES
PROTOCOL ADAPTER CARD FOR...
FREQUENCY CONVERTER
GENERATOR ASSEMBLY
LV SWITCH GEAR
MV SWITCH GEAR
TRANSFORMER
POWERFEEDER CABLES
UNKNOWN
POWER CABINET
PORTECTION CABINET
PITCH SYSTEM
BLADES
HUB
SLIPRINGS
HUB COVER
UNKNOWN
BLADE BEARINGS
SENSORS
YAW SYSTEM
NCAELLE COVER
NACELLE BEDPLATE
Figure 3.23 – Normalized hours lost per turbine per year due to faults in sub-systems and assemblies for turbines of
multiple manufacturers (for onshore wind turbines) (adapted from (Wilkinson and Hendriks, 2011))
Although the tower is not considered as a major source of problems, some aspects must be analyzed,
since the support structure of a wind turbine is fundamental for the integrity of the whole system. The
main problems assigned to the tower refer to corrosion and cracking (Liu, Tang et al., 2010), alongside
with more routine maintenance tasks, such as bolt tightening.
Some dramatic failures are registered in literature. In Devon, England, a small 50 kW wind turbine
collapsed in 2013. The cause was said to be the lack of resistance of the tower to the registered high
wind speeds (Gray, 2013). In Taiwan, a large wind turbine collapsed in 2008 due to insufficient
strength of bolts (Chou and Tu, 2011). Other cases of tower buckling due to a typhoon are reported in
(Li, Chen et al., 2013). A list of tower collapses is presented in (Chou and Tu, 2011), however, in some
cases, the causes are not available. In all these cases, the wind turbine was completely destroyed. The
share of wind turbine accident causes up to 2009 is presented in Figure 3.24. As can be seen, structural
failure represents a significant portion.
78
Chapter 3
Other
13%
Environmental
damage Blade failure
9% 23%
Human injury
5% Structural failure
Fatal 12%
accidents
9%
Fire
Ice rainfall Transport
19%
4% 6%
Figure 3.24 – Failure types of wind turbine accidents (Chou and Tu, 2011)
Figure 3.25 presents the estimated maintenance, repair and replacement costs of wind turbines (Gasch
and Twele, 2012). It classifies the support components (tower and foundations) as the elements with
the higher number of years between significant investments. This is expectable, since tower and
foundations are designed to maintain their structural integrity during the entire life period of the wind
turbine. However, this figure also demonstrates the high investment expected for the tower before the
expected period of 20 - 25 years, which means that, in average terms, this expenditure would have to be
considered.
180
Rotor blade
160
Maintenance and/ or replacement costs in €/ kW
140
Critical components
120
Uncritical components
100
Generator Tower
Pitch system
80
Gearbox
Control and
60 electronics Transformer Foundation
Yawing system station
40
Brakes
Shafts
Main bearing
20
Hydraulic system
Today’s period of experience Source: DEWI 2002
0
0 2 4 6 8 10 12 14 16 18 20
Number of years between significant investments in maintenance or replacement
Figure 3.25 – Estimated maintenance, repair and replacement costs of wind turbines (adapted from (Gasch and Twele,
2012))
79
Basics of Wind Turbines
Flow Direction
Figure 3.26 – The energy extracting stream tube of a wind turbine (adapted from (Burton, Sharpe et al., 2001))
This theory considers the following assumptions (Manwell, McGowan et al., 2010):
o Homogenous, incompressible, steady state fluid flow;
o No frictional drag;
o An in-plane continuous disc (similar to an infinite number of blades);
o Uniform thrust over the disc;
o A non-rotating wake;
o The static pressure far upstream and far downstream of the rotor is equal to the undisturbed
ambient static pressure.
A schematic illustration of the flow mechanism according to this theory is presented in Figure 3.27.
80
Chapter 3
Stream Tube
U0
Actuator Disk U0
UW
FT UD
U0
Axis
Stream Tube
Actuator Disc
Velocity
Velocity
Pressure
Axis
Pressure
Figure 3.27 – Actuator disc with stream tube. Velocity and pressure development along the stream tube (adapted from
(Tempel, 2006))
Considering the variation of linear momentum of the volume of air in the stream tube and the mass
conservation principle, it is possible to calculate the axial thrust force of the rotor ( ), considering the
variation of the cross section between 0 and ), represented in Figure 3.27:
. . . . (3.1)
with:
81
Basics of Wind Turbines
. . . . (3.2)
with:
(3.3)
Applying the Bernoulli equation between the sections far ahead and in the vicinity ahead of the rotor
(∗ , ), and between the vicinity of the rotor wake (∗ , ) and the far wake of the rotor, it becomes:
1 1
. . , . . ,
2 2
(3.4)
1 1
, . . , . .
2 2
with:
As exposed in Figure 3.27, the flow velocity is the same immediately before (section , 0) and after
(section , ) the actuator disc and the pressure is identical far away of the rotor disc (ahead and
after).
So, defining the thrust force according to the variation of pressure on each vicinity of the disc:
, , (3.5)
1
. . (3.6)
2
82
Chapter 3
From equation (3.6) and considering . . in equation (3.3), it is possible to conclude that
the velocity in the disc section is the average value of the upstream and downstream flow speed:
(3.7)
2
(3.8)
1
(3.9)
1 2.
As shown in the previous equations, defines the degree of slowdown of the wind at the actuator disc.
From equation (3.9), it is concluded that, in the extreme case, the induction factor can be = 1/2.
However, in this situation, the flow would cease and, thus, the flow mechanism would not be possible.
Since the power output ( ) is obtained as the rate of work done by the force at the actuator disc, it can
be written as:
1
. . . . . 4. 1 (3.10)
2
where the term defined in the first square brackets represents the power that is present in the flux,
while the second square brackets represents the efficiency of the power capturing.
Considering equations (3.6) and (3.9), the thrust force is defined as:
1
. . . . 4. 1 (3.11)
2
83
Basics of Wind Turbines
Usually, both and parameters are characterized by non-dimensional coefficients: the power
coefficient ( ) and the thrust coefficient ( ), respectively:
1
. . . . 4. 1
2 4. 1
1 (3.12)
. . .
2
1
. . . . 4. 1
2 4. 1
1 (3.13)
. . . .
2
In Figure 3.28 both coefficients are represented. As expected, the maximum values of the coefficients
are attained for different values of (as already referred, values of greater than 1/2 are unrealistic).
CT
(Dimensionless Amplitudes)
0.8
C P, C T
0.6 CP
0.4
0.2
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
a (Induction factor)
The calculation of the induction factor for the maximum values of the coefficients is obtained with:
d 16 1
0→ , ≅ 0.593, (3.14)
d 27 3
d 1
0→ , 1, (3.15)
d 2
While the results obtained for the thrust coefficient are not of too much importance (due to the fact
that the maximum value is obtained for an unrealistic situation), the result of equation (3.14)
represents an important achievement. The value of 16/27, named as Betz Limit, is considered the
theoretical maximum limit for power extraction from the flow. This means that, no matter the
efficiency of the machine, no wind turbine can extract more than 59.3 % of the kinetic energy of the
wind.
Obviously, this theory does not accurately characterize the real physical situation: the rotor is not
composed by a uniform disc with only axial thrust acting on it and the wind does not continue as an
undisturbed axial flow after passing through the rotor, among other things. To overcome these
84
Chapter 3
simplifications, the variation of angular momentum in the flow is considered in the theory presented
in the next section.
Ω
Axis
R
dr U0 (1-a)
r
ω.r
Figure 3.29 – Annular ring in the stream tube and decomposition of wake wind speed (adapted from (Manwell, McGowan
et al., 2010))
As shown in Figure 3.29, the wake flow has an axial speed of 1 (like in the axial momentum
theory) and a rotational speed of .
Defining an angular induction factor as:
′ (3.16)
2. Ω
85
Basics of Wind Turbines
2. Ω. ′ (3.17)
The torque of the annular stream is defined as the rate of change of angular momentum (due to the
rotational speed):
d . 2. . . d . 1 . 2. Ω. .
(3.18)
4. . 1 . . Ω. . .d
with:
With the defined rotor torque, it is possible to calculate the power output due to it:
d T. Ω (3.19)
Defining the relationship between the tangential speed of the rotor at a radial distance and the wind
speed as the local speed ratio (or tip speed ratio, for the case of ):
Ω.
(3.20)
Ω.
The power output (3.19) for each ring of radius and thickness d can be written as:
1
d . . . 2. . . d . 4. ′ 1 . (3.21)
2
Once again, the term defined in the first square brackets represents the power of the flux, while in the
second square brackets, the efficiency of the capturing element is quantified (also known as blade
element efficiency (Burton, Sharpe et al., 2001)).
86
Chapter 3
As shown in (3.21), the power output extracted from the wind when the rotational wake is considered,
depends not only on and , but also on the tip speed ratio. The power coefficient is defined by:
d 8
d ⇒ ′ 1 d
1 (3.22)
. . .
2
Although the calculation of the integral in the equation (3.22) is not presented in this work (it can be
consulted in (Manwell, McGowan et al., 2010)), it is important to note two things:
o The power coefficient ( ) depends on the tip speed ratio (λ);
o The maximum power that can be extracted from the wind tends to the Betz Limit (16/27).
Figure 3.30 illustrates the previous statements.
Betz limit
0.6
0.5
0.4
CP
0.3
0.2
0.1
0
0 2 4 6 8 10 12 14 16 18 20
λ (Tip speed ratio)
Considering the previous statements, it is possible to conclude that the evolution from rotors with
fixed-speed to rotors with variable-speed was a major evolution step for the efficiency in energy
harvesting. Whereas in a fixed-speed, the optimum tip-speed ratio is only achieved for a specific wind
speed, with a variable-speed turbine, it is possible to have different optimum points of operation for
each wind speed (Figure 3.31).
87
Basics of Wind Turbines
1200
variable speed
1000
800
fixed
speed 13 m/s
P [kW]
600
400 11 m/s
200 9 m/s
7 m/s
0
0 10 20 30 40 50 60 70 80 90 100
Rotational speed [RPM]
Figure 3.31 – Power output as function of rotational speed for a fixed- and variable-speed wind turbine for different wind
speed classes (adapted from (Tempel, 2006))
Although the previous momentum theories already explain the power extraction from a wind flow, the
problem is still not solvable yet due to unknown axial and angular induction factors ( and ). For
this reason, an additional approach to this problem is needed: the blade element theory.
88
Chapter 3
Ω.r.a’
δr
r
U0(1-a)
r Ω.r
Figure 3.32 – Rotor annular ring and detail of a blade element (adapted from (Burton, Sharpe et al., 2001))
As previously introduced, a blade element with a radius and a radial width d is subjected to an axial
wind speed of 1 and a rotational wind speed2 of Ω. created by a rotor angular speed of Ω.
Since tangential speeds are obtained from rotational speed multiplied by , the resultant relative
velocity between the blade element and the wind is:
1 Ω. 1 ′ (3.23)
An illustration of the relative velocity, its components and the associated lift ( ) and drag ( ) forces
is presented in Figure 3.33 (where the chord of the airfoil is represented by , represents the pitch
angle and represents the angle of inflow).
Ω.r(1+a’)
θ
α Ф
URel
FL
Wind Direction
U0(1-a)
c FD
Figure 3.33 – Relative wind velocity applied to the airfoil and the subsequent lift and drag forces
2
The rotational speed of the blade element is obtained assuming a linear increase from 0 (at the far ahead location) to
2. Ω. ′ (in the far wake location).
89
Basics of Wind Turbines
Having in mind the relative velocity defined in equation (3.23), it is possible to define the drag and lift
forces:
1
d . . . . .d (3.24)
2
1
d . . . . .d (3.25)
2
with:
Drag force
Lift force
Angle of attack
d 4. 1 . . . .d ,
(3.26)
Considering: d 2. . . d
d 4. . 1 . . Ω. . .d (3.27)
Considering now the results presented for the blade element theory (equations (3.24) and (3.25)) and
calculating their effects on the rotor, the following expressions for d and are obtained:
1
d . . .d (3.28)
2
1
d . . . .d (3.29)
2
with:
Number of blades
90
Chapter 3
.
(3.30)
2. .
1
(3.31)
with:
′ Local solidity
it is possible to obtain expressions for the axial and angular induction factors, equating (3.11) with
(3.28) and (3.18) with (3.29), respectively:
′ . .
(3.32)
1 4.
′ ′ . .
(3.33)
1 4. .
(According to some authors (Wilson and Lissaman, 1974), it is acceptable to set as zero)
However, it is important to consider that this theory refers to an independence of each annular section
of the blade to the adjacent ones. Although this does not correspond to reality, for normal operations
tip speed ratios, the error is considered as acceptable. Beside this aspect, some corrections should be
introduced to the BEM theory:
o According to equation (3.9), the axial induction factor has to be lower than 0.5, otherwise a
reversal of the wake flow would be induced. Since this cannot occur, a different explanation is
needed for this situation. In reality, for high values of (which lead to high values of ), the
permeability of rotor is almost comparable to a disc. Some authors, such as Glauert (Glauert,
1926), proposed a correction for the axial induction factor curve (Figure 3.34);
2. 0
Glauert empirical relation
1. 5
CT
1. 0 CT = 4a(1-a)
0. 5
0. 0
Figure 3.34 – Trend of thrust coefficients for high values of axial induction factor (adapted from (Manwell, McGowan et
al., 2010))
91
Basics of Wind Turbines
o Another correction that should be taken into account is the loss of power production near the
tip of the blades - called tip-loss factor. This is due to the discrete number and finite length of
real blades, which is not accounted in the classical approach of the BEM theory. The most
common approach to this problem is through the Prandtl’s approximation.
1.5
Tip/root-loss factor
0.5
0
0 0.2 0.4 0.6 0.8 1
r/ R
Figure 3.35 – Span-wise variation of combined tip/root-loss factor for a three-blade turbine optimized for a tip speed ratio
of 6 and with a blade root at 20% span (Burton, Sharpe et al., 2001)
92
Chapter 3
Ω.r(1+a’) Ω.r(1+a’)
θi = θf θi = θf
URel,i αi URel,f α f > αi
Wind Direction
1.0
CL
0.0
-10 0 10 20 30
α[º]
Wind turbines with this design present a lower aerodynamic efficiency below rated regime than with a
solution based on pitch control, since the rotor must operate closer to stall.
Although stall control seems to be a simple solution, it has some drawbacks, such as a less precise
control of the power production and the inability to drop the thrust force of the rotor at high wind
speeds (when compared to the pitch control solution) (Hau, 2006). For this reasons, the stall control is
usually only used with turbines with a power output up to 1.5 MW.
With large wind turbines, the attempts to introduce this control technology have been made with a
more complex construction, with the use of an active blade pitch angle adjustment. That is the case of
the Izar Bonus 1.3MW/62 studied in Chapter 5.
θ αf θ
URel αi URel
Wind Direction
U0(1-a) U0(1-a)
Pitch rotation
93
Basics of Wind Turbines
Pitch control of the blades allows a better control of the rotor speed, resulting in a more constant
power output. In Figure 3.38, the power curves of two wind turbines with stall and pitch control are
shown. As can be seen, the last part of the curve (after the rated speed), is considerably more flat with
pitch control. This is the solution implemented in both the Vestas V90-3.0 MW and Senvion MM82
studied in Chapters 5 and 7.
600
Pitch regulated
500 turbines
300
200
100
0 Rotor diameter = 40 m
0 5 10 15 20 25 30
Wind Speed [m/s]
Figure 3.38 – Power curve of a 500 kW wind turbine with rotor regulation by active pitch and passive stall (adapted from
(Burton, Sharpe et al., 2001))
1
. . π. . .Ω (3.34)
2
The maximum aerodynamic power is defined by introducing the optimum tip speed ratio (λ ) and
the maximum power coefficient ( ) in equation (3.34), while the rotor speed is optimum
(Ω ). Equation (3.34) can then be rewritten as:
.Ω (3.35)
where
1
. . π. . (3.36)
2
94
Chapter 3
Equation (3.35) shows that the maximum aerodynamic efficiency is proportional to the cubic of the
rotor speed. In the same sense, the torque is defined as proportional to the square of the rotor speed:
.Ω (3.37)
Considering these equations, it is concluded that, under steady state conditions (and assuming that the
variables from equation (3.36) are at their optimum values), the maximum efficiency can be achieved
by varying the rotor torque according to the squared value of the rotor speed. This is the control
strategy for maximum efficiency. However, some restraints prevent the operation of the turbines
following this strategy over the whole operating regimes.
Figure 3.39 illustrates the relationship between the torque and speed of the generator from a variable-
speed wind turbine with pitch control. For this turbine, the referred strategy is followed between the
operating points B and C. In this operating regime, the generator torque is proportional to the square
of the generator speed, in order to be in line with the optimum power coefficient (red line in Figure
3.39). This means that, (theoretically) optimal power extraction occurs within this operating zone.
Due to aerodynamic noise, loads and other design constraints, it is not possible to follow the optimum
curve for all wind speed regimes, being the maximum allowable rotor speed achieved at a relatively
low wind speed (Burton, Sharpe et al., 2001). For that reason, when reaching the operating point C, a
different control strategy is followed. At this point, the increase of power production (with the increase
of wind speed) is obtained by increasing the torque, while the rotor speed is kept at its rated value (line
CE). When the rated generator rotor torque value is achieved, the power production is maintained at
the same level by adjusting the pitch angle of the blades, which allows controlling the input torque
generated by higher wind speeds (point E).
It should be noted that the described chain of operating regimes are referred to a situation in which no
limitations are present for control. In reality, some constraints are present in the definition of the
control algorithms, leading to the need of transitional regimes. Thus, the dashed line A’B’C’D’E’
defines a chain of operating conditions more in line with the operation of wind turbines in real
conditions.
180
Rated speed
18 m/s
160
16 m/s
Generator torque
140
14 m/s
Rated torque
120 12 m/s
D’ E
100
10 m/s
80
8 m/s
C
60
6 m/s
C’
40
4 m/s
20
B
Optimum Cp
0
0. 00 200.00 40 00 600. 800.00 1000.00 1200.00 1400.00 1600.00 2000.00
A’ A Generator speed
Figure 3.39 – Illustration of the torque-speed curve for a variable-speed wind turbine with pitch control (adapted from
(Burton, Sharpe et al., 2001))
95
Basics of Wind Turbines
96
Chapter 4
4
STRUCTURAL BEHAVIOUR OF WIND
TURBINES
Waves
Tides and
Currents
Figure 4.1 – Sources of dynamic excitations of an offshore wind turbine (for onshore turbines, tides, currents and waves
are, obviously, not applicable)
97
Structural Behaviour of Wind Turbines
Thus, an accurate description of the dynamic behaviour is extremely important in this context,
otherwise resonance phenomena are very likely to occur due to several sources.
A complete wind turbine structure can be idealized as an assembly of coupled multi degree-of-freedom
mass-spring-damper systems. With this in mind, the structure is defined according to its mass,
damping and stiffness properties (Clough and Penzien, 1995):
. . . (4.1)
with:
Mass matrix
Damping matrix
Stiffness matrix
Displacement vector
Excitation force
Although the complete characterization of a wind turbine might be complex, simplifications for some
components can be made for initial considerations. Within the scope of this thesis, the most important
aspects related with the dynamic characterization of the tower, blades and the tower-blades coupling
are presented.
Figure 4.2 – Simplified structure for calculation of the natural frequency of the tower bending mode
98
Chapter 4
With this configuration, approximate values of the tower first natural frequency can be obtained
through:
, .
2. (4.2)
. , . .
with:
, Constant value: 1.732 (Young and Buynas, 2002) or 1.7436 (Vugts, 2000), depending
on literature sources
, Constant value: 0.236 (Young and Buynas, 2002) or 0.227 (Vugts, 2000), depending
on literature sources
Tower top mass (including nacelle, all the equipment in it and the rotor)
Tower height
Figure 4.3 presents an illustration with the mode shape configuration of the first vibration mode for
the fore-aft (1 FA) and side-side (1 SS) directions.
1 FA 1 SS
Figure 4.3 – First tower vibration mode for fore-aft and side-side directions: undeformed (grey) and deformed (black)
structure (adapted from (Skjoldan, 2011))
99
Structural Behaviour of Wind Turbines
The rotor modes of wind turbines can be divided into two groups: flapwise and edgewise (see Figure
3.2). For each order of these modes, three “sub-modes” need to be considered, namely, one symmetric
and two asymmetric modes in each direction. These modes are presented in Figure 4.4. The symmetric
mode is characterized by the deflection of the three blades in the same direction:
o Flapwise modes: the blades deflect in the rotor out-of-plane direction (1 FS);
o Edgewise modes: the blades deflect in the rotor plane direction (1 ES).
On the other hand, the two asymmetric modes are defined accordingly to the direction of the motion
of the blades:
o Flapwise modes: tilt (1 FT) and yaw mode (1 FY);
o Edgewise modes: horizontal (1 EH) and vertical mode (1 EV).
1FS 1FT 1FY
1 ES 1EH 1EV
Figure 4.4 – First order rotor modes (for low values of pitch angle): undeformed (grey) and deformed (black) structure
(adapted from (Skjoldan, 2011))
These asymmetric modes are associated with translation and rotation of the rotor. This is the reason
why the stiffness of the supporting elements of the rotor has an important role in these modes.
When the wind turbine is under operation and the rotor is spinning, two particular effects need to be
accounted for: gyroscope effects and centrifugal stiffening. These effects should be considered in the
equation of motion (equation (4.1)) of the rotor as:
. (4.3)
with:
Gyroscopic matrix
100
Chapter 4
Gyroscope Effect
The rotation of the rotor introduces a well-known characteristic in the context of rotor dynamics: the
gyroscopic effect. This effect is due to the flexibility of supporting bearings of the rotor which leads to
deviation of the rotating axis from the bearing centre line (Yoon, Lin et al., 2013). Essentially, when the
rotor is rotating at the shaft axis, the motion in the two orthogonal axes is coupled. This is evident by
the skew-symmetric nature of the gyroscopic matrix (Ewins, 2000; Yoon, Lin et al., 2013). Thus, under
rotation, the rotor suffers two rotational movements: the rotation of the shaft itself; and a whirl
rotation due to gyroscopic effects. There are two types of whirl rotation, namely, forward (when the
rotation is in the same direction of the shaft) and backward (when the rotation is in the opposite
direction). Figure 4.5 illustrates these two phenomena for two different situations: one where the
rotor/ stator system is symmetric and another where the system is non-symmetric.
Ω Ω Ω Ω
The dynamic behaviour of wind turbine rotors is similar to what was described above. With the
rotation of the rotor, the two asymmetric modes couple (the tilt with the yaw flapwise mode; and the
horizontal with the vertical edgewise mode), resulting in the development of the whirl modes. The
frequency of these whirl modes (backward – BW - and forward - FW) drift, from a fixed-frame
reference, due to the rotor rotation. This situation is illustrated in Figure 4.6. As exposed in Figure 4.5,
the forward and backward definition of the whirl modes refers to the rotation of the rotor centre of
mass in the same or against the rotor direction, respectively (Skjoldan, 2011).
101
Structural Behaviour of Wind Turbines
2 FT
2 FY 1. edgewise forward whirling
2. flapwise backward whirling
1 EH -Ω
3
1 EV +Ω
Natural frequency [Hz] 1. edgewise backward whirling
0
0 5 10 15 20 25
Figure 4.6 – Campbell diagram of fixed-frame natural frequencies (O) for the first 10 structural modes of a 600 kW three-
bladed wind turbine. Lines denote the centre frequencies of the rotor whirling modes given Ω added to the fixed-frame
natural frequencies (adapted from (Hansen, 2007))
Figure 4.6 illustrates some important phenomena that should be considered when studying the
dynamic behaviour of wind turbine rotors.
When the rotor is stopped, modes 4 and 5 (first flapwise yaw and flapwise tilt modes) are very close.
The difference between them is due to different elastic supports in the corresponding directions.
Usually, the yaw mode is lower than the tilt mode because towers are stiffer in tilt than in yaw
direction (Hansen, 2003). The 6th mode, the flapwise symmetric mode, is slightly higher than the
asymmetric ones. This is justified by the fact that this mode, in which all the blades vibrate
simultaneous in-phase, beholds also some fore-aft motion of the tower in out-of-phase (Hansen,
2003). When the rotor starts to operate, a change in the frequency value of the modes is evident. As
highlighted in Figure 4.6, the pair of asymmetric modes starts to couple and became pairs of whirling
modes. Thus, FW and BW modes start to drift, increasing and decreasing their frequency values,
respectively, in an amount equal to the rotor speed (±Ω). This splitting phenomenon, identified
considering a fixed reference of observation, is due to the gyroscopic effect.
Modes 7 and 8 have a similar behaviour as the described for the 4th and 5th mode. These frequencies are
also slightly different when the rotor is stopped due to different flexibilities of the rotor supports.
When the rotor starts to spin, a similar splitting effect is noticeable.
In the last two modes (2 flapwise FW and BW modes), the difference at parked situation of these two
frequencies indicates an even greater difference in the requested flexibility of the support in both
directions. Although the behaviour of these two modes is very similar to modes 4 and 5, it is important
to refer that it is less pronounced due to the lower contribution of the gyroscopic reaction forces
(Hansen, 2007).
102
Chapter 4
The evolution of the frequencies with the rotor speed usually leads to situations where two different
modes stay very close to each other. In the presented situation, the 1st flapwise FW with the 1st
symmetric mode and the 1st edgewise FW with the 2nd flapwise BW feature very close frequencies. In
these situations, the vibrations modes interact with each other, presenting a mode shape with a mixed
configuration.
One last aspect should also be mentioned. If Figure 4.6 is observed carefully, it is possible to verify that
the frequency of the 1 FS mode, which does not contain any gyroscopic effect, also increases with the
rotor speed. This situation is also noticeable for the centre lines of the whirl modes (solid line) which
imply that another effect is present that increases the value of the frequencies too. This phenomenon is
called centrifugal stiffening.
. .Ω d (4.4)
with:
Centrifugal force
Blade length
Alongside with the tensile force presented in the blade due to the centrifugal forces, the influence of
the gravity (self-weight) can also be introduced. This effect, which is reflected by an axial force acting
on the blade, depends on the position of the blade (Murtagh, Basu et al., 2005). Thus, the effect of the
self-weight of the blade should be added or subtracted to its stiffness, either it is a tensile or a
compressive force, respectively. However, it is referenced that this gravity effect is negligible when
compared to the effect of the centrifugal stiffening (Arrigan, Pakrashi et al., 2011).
103
Structural Behaviour of Wind Turbines
Several authors, among which (Naguleswaran, 1994) and (Yoo and Shin, 1998), studied the problem of
determination of natural frequencies of simple beams rotating on a hub taking into account the
centrifugal stiffening. However these studies focused on simple beams, with constant parameters along
the length of the beams, which diverge substantially from a common wind turbine blade. More
complex studies were developed for the specific case of the blades, e.g (Hansen, 2003), however they
are outside the scope of this work. Nevertheless, a rough estimation of the increase of the natural
frequency of a blade due to centrifugal stiffening is presented (Putter and Manor, 1978):
, .Ω (4.5)
with:
Although this effect is important and should be considered when studying the dynamic behaviour of
the rotor vibration modes, it is not very noticeable in edgewise modes. This is justified by the softening
effect that is also introduced by the centrifugal forces on the linear stiffness of deflection in the rotor
plane that almost cancels the stiffening effect (Hansen, 2007).
104
Chapter 4
0.4
0.3
m = 250 t
0.2
ground shear module 106 N/m2 t = 18 mm
H = 70 m
D = 5.0 m
- sand, loose, round 20 - 50
- sand, loose, sharp 40 - 80
0.1
- sand, med. dense, round 50 - 100
- sand, med. dense, sharp 80 - 150
- pebbles, without sand 100 - 200
12 m
- brash, sharp 150 - 300
0
0 100 200 300
Ground shear module ES 106 N/m2
Figure 4.7 – Influence of the soil stiffness on the first natural bending frequency of a wind turbine tower (Hau, 2006)
For the modelling of the soil stiffness, the recommendations defined in the DNV guidelines (Det
Norske Veritas (DNV), 2002) are then presented. In order to introduce the effect of the soil, four
foundation springs must be included:
o Vertical spring stiffness - ;
o Horizontal spring stiffness - ;
o Rotational (rocking) spring stiffness - ;
o Torsional spring stiffness - .
It is important to note that soil behaves in a non-linear manner. But instead of using non-linear
springs, it is common to model the springs according to the strain level of the soil and then using
springs with linear behaviour. DNV (Det Norske Veritas (DNV), 2002) suggests three different levels
of dynamic loading (and corresponding shear strains ):
o From earthquakes: large shear strains up to 10-2 to 10-1;
o From wind and ocean waves: moderate shear strains up to 10-2, typically 10-3;
o From rotating machines: small shear strains, usually less than 10-5.
105
Structural Behaviour of Wind Turbines
After deciding which load case is of interest, it is necessary to compute the shear modulus of the soil.
This is usually accomplished through the use of the plot presented in Figure 4.8 together with the
initial shear modulus . The value of can be obtained according to the empirical expression:
3
′ (4.6)
1
with:
Void ratio
Over consolidated ratio for clay soil (equal to 1.0 for sand)
Once the shear modulus of the soil is obtained, the Young modulus of soil is calculated according to:
2. 1 (4.7)
with:
Young modulus
Table 4.1 – Coefficient k Table 4.2 – Coefficient K Table 4.3 – Poisson ratio
Plasticity Index
k Type of soil K Type of soil ν
(I )
106
Chapter 4
Two simpler alternative expressions for the determination of can be used for sand:
1000. . ′ (4.8)
with:
2600. (4.9)
with:
After computing the value of the initial shear modulus, the value of shear modulus ( ) and damping
( ) of the soil can be obtained with the help of Figure 4.8. Once these values are calculated, the stiffness
of the foundation springs for circular slab footing are given by the equations presented in Table 4.4.
1.00
0.80 0.20
Typical
Range
G( γ)/G 0
0.60 0.15
ζ(γ)
0.40 0.10
0.20 0.05
0 0
10-5 10-4 10-3 10-2 10-1 10-5 10-4 10-3 10-2 10-1
γ γ
Figure 4.8 – Shear modulus and damping ratio according to strain level (adapted from (Det Norske Veritas (DNV), 2002))
It is important to refer that the values obtained from these tables are related to static stiffness (for
excitation frequencies approaching zero). This assumption is correct for majority of the dynamic
loading acting on wind turbines, although may introduce some errors for high frequency excitations.
After the foundation stiffness is calculated, springs should be introduced in the wind turbine model
according to Figure 4.9 in order to more accurately reproduce the boundary conditions between
structure and soil.
107
Structural Behaviour of Wind Turbines
KR
KT
KH
KV
108
Chapter 4
Table 4.4 – Stiffness of foundation springs for circular footing on different substratum conditions (adapted from (Det Norske Veritas (DNV), 2002))
R R R
D
G
Foundation
H ν
G G1
H H ν1
spring ν
Bedrock
G2
Bedrock ν2
D/R < 2
0 ≤ G1/G2 ≤ 1 D/H < 1/2
4. . 4. . 1 1.28. 4. . ⁄
Vertical 1 1.28 . ;1 ⁄ 5 1 1.28. 1 1 0.85 0.28.
1 1 1 1.28. . 1 2. 1 ⁄
4. . 8. . 1 8. . 2. 5.
. 2.
Horizontal 1 1.28 1 ;1 ⁄ 4 1 1 1
1 1 . 1 2. 3. 4.
2.
8. . 8. . 1 8. .
. 6.
Rotation 1 3 1 ;0.75 ⁄ 2 1 1 2. 1 0.7.
3 1 6. 1 . 3 1 6.
6.
16. . 16. . 8.
Torsion - 1
3 3 3.
109
Structural Behaviour of Wind Turbines
1Ω 2Ω 3Ω 4Ω
Log Frequency
Figure 4.10 – Illustration of the spectral energy distribution for rotationally sampled spectrum. The peaks refer to integers
of the rotor frequency of rotation (adapted from (Murtagh, Basu et al., 2005))
110
Chapter 4
1Ω 3Ω
Rotor Rotational Frequency
Figure 4.11 – Frequency intervals for tower stiffness for a fixed rotational speed wind turbine
Figure 4.11 refers to a situation of a fixed-speed wind turbine. For these cases, since there is only one
operational rotational speed, the range of possible values for the tower natural frequency is wider.
However, in the case of variable-speed wind turbines, further considerations had to be taken into
account. In these situations, the range for the tower natural frequency becomes much narrower. With
the purpose of a clearer explanation, a Campbell diagram of a 1.5 MW three-bladed variable-speed
wind turbine is presented in Figure 4.12. In it, the natural frequencies of several components of the
wind turbine are plotted against the rotational speed. The range of operation of the rotor is also shown.
As can be seen, the first natural frequency of the tower remains above the rotational speed (1Ω) and
below the integer multiple 3Ω. Thus, this is considered a soft-stiff tower.
111
Structural Behaviour of Wind Turbines
2 nd Flapw
Flapwise bending 8Ω
3 7Ω
2 nd To wer bending 6Ω
2 1 st Edgewise bending 5Ω
4Ω
3Ω
1 1 st Flapw.
2Ω
bending
1st Tower Longitudinal 1Ω
bending Lateral
Rotational speed
in rpm
0 5 10 15 20 25
Figure 4.12 – Campbell diagram of a 1.5 MW three-bladed rotor, variable-speed wind turbine (adapted from (Gasch and
Twele, 2012))
112
Chapter 4
4.5 DAMPING
In the dynamic characterization of the wind turbine system, other important parameters that need to
be determined, besides the value of the natural frequencies, are the damping coefficients. With the aim
of illustrate the importance of these parameters on the supporting structure, a simple single degree of
freedom (SDOF) structure, defined by its mass , stiffness and damping , is presented in Figure
4.13 a). Its dynamic amplification factor (DAF – which represents the magnification the structure
response suffers with the application of a dynamic load relative to the response to a static force) is
defined by (Clough and Penzien, 1995):
(4.10)
1 2. .
with:
In Figure 4.13 b), the DAF for a simple SDOF structure was calculated for two different values of
damping ratios. As can be observed, the DAF value is highly dependent on the damping value. The
value of the structural response in resonance is considerably higher for a structure with a damping
ratio of 0.2 % (a common value for a wind turbine in an idling or parked situation) than for a structure
with = 4 % (a conservative value for a wind turbine under production). For this reason, it is easy to
understand that, for structures subjected to dynamic excitations (such as wind turbines), their
damping value is a key-factor to handle with the problem of fatigue.
The damping of a wind turbine can be considered as linear sum of different damping sources:
structural, aerodynamic, from devices, soil and hydrodynamic. From their nature, the last two are only
referred to offshore wind turbines.
3
10
2
10 ξ = 0.2 %
1
DAF
10
ξ = 4.0 %
0
10
1
-1
10
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
f /f
load 1
a) b)
Figure 4.13 – a) SDOF structure. b) DAF values for different damping ratios
113
Structural Behaviour of Wind Turbines
“A motion of the tower top in the wind direction (…) results in a smaller angle of attack at the rotor
blades [see Figure 3.33] because the apparent out-of-plane velocity component [U 1 a ] is reduced by
the tower top velocity. A smaller angle of attack corresponds, for attached flow conditions, to lower
aerodynamic lift and drag forces [equations (3.24) and (3.25)] and to a reduction of the thrust force (…).
Likewise a movement of the nacelle against the wind direction increases both the angle of attack and the
thrust force. In both situations the alternation of the thrust is oriented opposite to the disturbing tower
top motion and is experienced as aerodynamic damping. (…)
For higher angles of attack, stall occurs and the dynamic damping is lower or may even become negative
because the slope of the lift curve is reduced. Under such conditions aero-elastic instability can occur if
insufficient damping exists.”
For the purpose of illustration, Figure 4.14 presents the experimental results for lift and drag
coefficients for the NACA 63-415 airfoil obtained in a wind tunnel by Bak et al (Bak, Fuglsang et al.,
2000). As can be seen, the lift coefficient has an almost linear increase for the initial range of angle of
attack. Then, the slope of the curve starts to decrease until a peak is reached. After this point, the
variation of the lift coefficient with the angle of attack becomes negative. The values of drag
coefficients for angle of attack of around 15º are considerable low.
114
Chapter 4
2.0 0.5
1.5 0.4
1.0 0.3
CD
CL
0.5 0.2
0.0 0.1
-0.5 0.0
-10 0 10 20 30 -10 0 10 20 30
α[º] α[º]
Figure 4.14 – Variation of lift ( ) and drag ( ) coefficients with the angle of attack (adapted from (Bak, Fuglsang et al.,
2000))
Garrad (Kuhn, 2001) presented a simple equation (4.11) for the calculation of the aerodynamic
damping for a wind turbine support structure. It assumes that the wind turbine is operating under
stationary rotor dynamics at a high tip speed ratio, where the inflow angle ( ) is small and the drag
forces are negligible compared to lift forces (Freris, 1990; Salzman and Tempel, 2005).
. .Ω d
. . d (4.11)
8. . . d
with:
Airfoil chord
From the previous equation, some important aspects should be highlighted. The aerodynamic
damping is inversely proportional to the natural frequency (and modal mass of the structure). This
mean that structures with different philosophies of tower stiffness design (as presented in section 4.4)
will lead to very different values of aerodynamic damping. On the other hand, the increase of the
rotational speed of rotor increases the value of this additional portion of damping.
Some other important aspects should be also mentioned:
o The aerodynamic damping is specially noted for the first bending mode of the tower in the
rotor out-of-plane direction (the first fore-aft vibration mode). This is expected since the
aerodynamic forces act mainly in this direction;
o In the rotor plane direction (the side-side direction) the effect of the aerodynamic damping is
expected to be negligible since there is a very low level of aerodynamic forces in this direction;
o In idling or parked situations, the effect of aerodynamic damping is expected to vanish for
both orthogonal directions. This situation is due to the fact that, in these situations, the blades
are usually in a feathered position (with high angles of attack). Nevertheless, some effect could
115
Structural Behaviour of Wind Turbines
be noted for the side-side direction due to the opposition of the blade to the wind flow in this
direction;
o Variable and fix rotor speed turbines present different values for aerodynamic damping
(Salzman and Tempel, 2005). Fix speed turbines operate at high tip speed ratios in low wind
speeds and lower tip speed ratios at high wind speeds (due to the constant rotor speed). For
these reasons, in low wind speed the angle of attack is small and, consequently, the flow is
attached to the blade. When the wind speed is higher, the angle of attack increases which lead
to higher angles of attack and some separation of the flow at the blade. In the case of variable
speed rotors, since they operate with their blades close to stall under rated wind speed, there is
some flow separation which leads to an aerodynamic damping below the optimal value. For
the higher wind speed, the flow stays attached to the blade and the damping is higher.
k1 k2
m1 m2
c1 c2
Figure 4.15 – Idealized structural system with the damper system (index 2) attached to the main system (index 1)
116
Chapter 4
The damping introduced by the soil is dependent on the type of soil and also, to some extent, on the
type of foundation. Along with the inherent complexity of this phenomenon, it is comprehensible that
there is a quite wide range of values between published works. Damping ratio values between 0.23 %
and 1.00 % for monopile foundations are referred in (Tarp-Johansen, Andersen et al., 2009;
Versteijlen, Metrikine et al., 2011; Damgaard, Ibsen et al., 2013a). Damgaard et al (Damgaard, Ibsen et
al., 2013b) also studied the influence of soil damping in an offshore prototype wind turbine with a
bucket foundation and found a value of 0.16 % for the damping ratio during power production.
.
(4.12)
with:
Wave period
For small values of (below 2 (Naess and Moan, 2012)), the radiation damping is dominant; while
for high value of , the drag component prevails. Thus, the contribution of each damping term
depends on the type and size of the foundation solution (and also on the sea conditions).
Some authors refer the larger contribution of the radiation damping with respect to the viscous drag
component for monopile wind turbine foundations (Tarp-Johansen, Andersen et al., 2009; Shirzadeh,
Devriendt et al., 2013). Some typical values for the wave radiation component of the hydrodynamic
damping are present in published works. Leblanc and Tarp Johansen (LeBlanc and Tarp-Johansen,
2011) present a value of 0.12 % for the damping ratio of an offshore wind turbine with a pile diameter
of 4.7 m, with a water depth of 20 m and a natural frequency value of 0.3 Hz. Also in (Germanischer
Lloyd (GL), 2005), the value of 0.11 % for damping ratio of a monopole wind turbine is presented.
Likewise, Tarp-Johansen et al (Tarp-Johansen, Andersen et al., 2009) present a value of 0.22 %,
following a numerical procedure.
Besides the hydrodynamic effect of the pile, the presence of mooring lines in floating wind turbines
also represents an important contribution to the damping of the system (Hall, Buckham et al., 2013).
117
Structural Behaviour of Wind Turbines
-2
10
Amplitude
-4
10
-6
10
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Frquency [HZ]
Figure 4.16 – Frequency domain spectrum of the acceleration of the top of the tower of a wind turbine due to turbulent
wind inflow
In the situation of the Figure 4.16, the cyclic loads are only due to aerodynamic forces of three types:
o Rotational sampled turbulence spectrum: As already explained in section 4.3, for each
revolution, each blade passes through a gust once. This will lead to an excitation of 1Ω for the
blades and of .Ω for the structures that support the rotor (3Ω in the case of Figure 4.16);
o Tower shadow: For an up-wind rotor, the presence of the tower disturbs the wind flow,
retarding the wind speed. The decrease in the wind speed leads to a decrease of the force on
the blade every time it passes in front of the tower, causing a cyclic excitation on the blade
(and, consequently, on the tower). For down-wind rotors, the effect of the tower is more
severe. The retardation of the flow is felt at larger distances and lead to strong excitations;
o Vertical wind shear: The mean wind velocity increases with height (Figure 4.17). With today’s
large rotor diameters, blades are long enough to experience a reasonable variation of wind
speed from top to bottom position. Thus, blades will complete a cycle from the maximum to
the minimum mean wind load in each rotor revolution.
118
Chapter 4
Unbalanced rotors represent another source of cyclic loads. It leads to a rotating force that can excite
lateral vibrations of the nacelle and tower (Gasch and Twele, 2012). As examples of causes of
unbalanced rotors, manufacturing defects and accumulation of ice can be pointed out. In Figure 4.18, a
comparison of the response between an unbalanced mass rotor and a balanced rotor is shown. The
time history was obtained through a numerical analysis, where the unbalanced rotor was defined by an
abnormal mass at the tip of one of the blades with the purpose of evidence the described phenomenon.
As can be observed, the unbalance presented in the rotor introduced an excitation with a frequency of
1Ω in the support structure.
2
10
Balanced rotor
Unbalanced rotor
0
10
Amplitude
-2
10
-4
10
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Frequency [Hz]
Figure 4.18 – Time history of the tower top accelerations for a balance and unbalance mass rotor
For offshore wind turbines, another source of excitation forces needs to be accounted for. Wave
excitation acts on support structures in a wide range of frequencies, usually bellow the tower natural
frequency (Tempel, 2006). Thus, time histories of the motion of an offshore tower will likely record
these excitations, which need to be analysed and discharged when a dynamic continuous monitoring
system is installed in a wind turbine. The usual wave excitation range of frequency is illustrated in
Figure 4.19. As can be seen, the frequencies of excitation are not very well defined as are for
aerodynamic forces.
119
Structural Behaviour of Wind Turbines
100
f1
80
Ocurrence [%] 1Ω 3Ω
60
40
20
0
0 0.2 0.4 0.6 0.8 1.0 1.2
Frequency [HZ]
Figure 4.19 – Range of frequencies of wave excitation on an offshore wind turbine (Tempel, 2006)
120
Chapter 4
5000 12
126 m
+ 87.600
4000 S3
9
Power Output [kW]
3000 S2
+ 60.000
6
2000 90 m
S1
3 + 30.000
1000
0 0
0 5 10 15 20 25 + 0.000
Wind speed [m/s]
Figure 4.20 – Power curve (in blue) and relationship Figure 4.21 – Illustration of the numerical model of the
between rotor speed and wind speed (in red) of the NREL NREL 5MW and position of the accelerometers
5MW
The aeroelastic analysis was performed for a time length of 700 s, with a time step of 0.02 s. The first
100 s were disregarded to avoid transient responses from the start of operation of the turbine.
121
Structural Behaviour of Wind Turbines
During the analysis, the wind turbine was subjected to artificially generated wind speed time series
with a mean wind speed of 15 m/s (Figure 4.22). The turbine was operating at the rated power output
conditions. At this regime, the pitch angle is used to control the rotor speed and, for that reason, it
varies along the analysed 600 s. The rotor mean speed is 12.1 rpm. The evolution of the rotor speed
and of the blades pitch angle throughout the analysis is shown in Figures 4.23 and 4.24.
Wind speed at hub height 6 Wind speed at hub height
17 10
16 4
10
Wind spedd [m/s]
Amplitude
15
2
10
14
0
10
13
-2
12 10
100 200 300 400 500 600 700 0 1 2 3 4 5
Time [s] Frequency [Hz]
Figure 4.22 – Generated wind speed at the hub height (time and frequency domain)
12.2 11
Rotor speed [rpm]
12.1 10
Pitch angle [º]
12 9
11.9 8
11.8 7
11.7 6
100 200 300 400 500 600 700 100 200 300 400 500 600 700
Time [s] Time [s]
The acceleration time series obtained from the analysis of the considered sensors are presented in
Figure 4.25, in both time and frequency domain. In this figure, only the considered 600 s are shown. It
is interesting to note that the vibration levels recorded by sensor S2 (height + 60.000) are the highest
from the three sensors.
In the frequency domain plots, the frequencies of the 3Ω, 6Ω, …, 24Ω harmonics are indicated with
vertical dashed lines. The influence of the first 5 harmonics in the structural response of the wind
turbine is clearly visible.
122
Chapter 4
Sensor S1 3 Sensor S1
10
0.2
2
10
Acceleration [m/s2]
0.1
1
Amplitude
10
0
0
10
-0.1
-1
10
-0.2
-2
10
100 200 300 400 500 600 700 0 1 2 3 4 5
Time [s] Frequency [Hz]
Sensor S2 3 Sensor S2
0.4 10
0.3
2
10
0.2
Acceleration [m/s2]
0.1 1
Amplitude
10
0
0
-0.1 10
-0.2 -1
10
-0.3
-2
-0.4 10
100 200 300 400 500 600 700 0 1 2 3 4 5
Time [s] Frequency [Hz]
Sensor S3 3 Sensor S3
10
0.2
2
10
Acceleration [m/s2]
0.1
1
Amplitude
10
0
0
10
-0.1
-1
10
-0.2
-2
10
100 200 300 400 500 600 700 0 1 2 3 4 5
Time [s] Frequency [Hz]
Figure 4.25 – Wind turbine acceleration captured by the sensors (time and frequency domain)
In order to ease the comprehension of some procedures presented throughout this thesis, only the two
first tower vibration modes were considered. The dynamic properties of the wind turbine were
obtained through an eigensolver implemented in the HAWC2 code. The natural frequency and the
modal damping ratio of these two modes are presented in Table 4.5. It is important to note that the
damping values presented are only referred to the structural component of damping and, for that
reason, do not necessary reflect the effective damping of the modes during the analysis since the
aerodynamic damping is not considered in the eigensolver and, consequently, in the values of shown
in Table 4.5.
123
Structural Behaviour of Wind Turbines
The configuration of the considered vibration modes is illustrated in Figure 4.26. The 1st tower
bending mode is clearly a pure tower bending mode, with a first order configuration of the tower
motion (Figure 4.26 a)). On the other hand, the 2nd tower bending mode beholds, besides the tower, an
important contribution of the rotor to the global motion of the mode (Figure 4.26 b)).
a) d)
Figure 4.26 – Illustration of the considered vibration modes: a) 1 FA; b) 1 FS and c) 2 FA
124
Chapter 4
125
Structural Behaviour of Wind Turbines
126
Chapter 5
5
OPERATIONAL MODAL ANALYSIS OF
WIND TURBINES
127
Operational Modal Analysis of Wind Turbines
Nevertheless, OMA presents some disadvantages inherent to its concept (Magalhães and Cunha,
2011):
o Although the excitation is assumed to be of broad band nature, the frequency content may not
cover the whole spectrum of interest, mainly in the case of structures with high natural
frequencies. Also, if the excitation does not correspond to a flat distribution of energy along
the frequency spectrum, some misidentification might occur;
o The modal mass is not estimated.
OMA has been widely used in the last years in several engineering domains. In civil engineering, this
technique has been applied in diverse structures. Bridges (Cunha, Caetano et al., 2001), (Caetano,
Magalhães et al., 2007), (Brownjohn, Magalhães et al., 2010), dams (Rodrigues, 2004)) and buildings
((Ventura, Finn et al., 2003), (Shi, Shan et al., 2012), (Gentile and Saisi, 2007) are among the most
commons examples of structures tested with OMA methodology. In mechanical engineering, OMA is
also applied in different structures, such as helicopters (Peeters, Cornelius et al., 2007), (Ameri,
Grappasonni et al., 2013), aircrafts (Mevel, Benveniste et al., 2006), (Hermans and Auweraer, 1999) or
vehicles (Goursat, Döhler et al., 2010), (Hermans and Auweraer, 1999).
128
Chapter 5
129
Operational Modal Analysis of Wind Turbines
(Carne, Lauffer et al., 1988), evidencing the advantages of ambient vibration tests (considerably faster
and more practical) with regard to forced vibration tests.
The success achieved with the previous application led to the development of a new identification
method named Natural Excitation Technique (NExT) (James, Carne et al., 1993). This technique was
further validated and used in numerical (FloWind Corporation 19 m VAWT) and experimental tests,
both in parked - FloWind Corporation 19 m VAWT - and in operation – DOE/ Sandia 34 m VAWT -
situations (James, Carne et al., 1993). Reference should also be made to the application of NExT to a
rotating 100 kW horizontal axis wind turbine (James, 1994).
After these initial tests, other authors have also performed some modal identification analysis on wind
turbine structures. Molenaar (Molenaar, 2003) performed an experimental modal test on a 750 kW
horizontal axis wind turbine, under parked conditions, in order to compare the results with a
numerical model. For the test, the wind turbine was excited through the step-relaxation technique by
applying a static load on the tower top and the structural response was measured with 19
accelerometers mounted both on tower and blades.
In (Osgood, Bir et al., 2010), continuing earlier works (Osgood, 2001), a parked 600 kW three-bladed
horizontal axis wind turbine was experimentally tested following both forced and ambient vibration
test procedures. For the forced vibration test, two independent hydraulic actuators were used, which
excited the structure at the top in both orthogonal directions. The turbine was heavily instrumented
with accelerometers, totalling 75 channels distributed in the tower, rotor and drive train. A very good
correlation was obtained with the two tests (although it was possible to detect additional modes in the
forced vibration test).
(Hansen, Thomsen et al., 2006) explored different excitation techniques for wind turbine testing.
Experimental results obtained with two different types of harmonic excitation were confronted with
the results from an ambient vibration test. The harmonic excitation was induced by blade pitch angle
variation and by electrical torque variation. However, it was concluded that the excitation of wind
turbines with these techniques presents some problems, namely:
o The excitation of high-frequency and highly damped modes is not technically possible due to
limitation of the blade pitch actuators;
o The excited structure vibrations are not pure modal vibrations and the estimated damping is
therefore not the actual modal damping. Since it is not possible to isolate the desired modes,
the estimation of damping for modes with close frequencies will result in considerable errors.
In the opposite way, the authors (Hansen, Thomsen et al., 2006) concluded that the results obtained
with the adopted output-only tool confirmed its suitability for identifying closely-spaced modes.
Griffith et al (Griffith, Mayes et al., 2010) conducted a modal identification test on a small 60 kW
VAWT under parked conditions. As in the above references, an ambient vibration and forced
vibration (excitation from an impact force, from step-relaxation and from human random excitation)
were performed and both accelerometers and strain gauges were used. The results revealed good
agreement between all the tests.
Chauhan et al (Chauhan, Tcherniak et al., 2011) and Carcangiu et al (Carcangiu, Tcherniak et al.,
2012) experimentally studied a 3 MW, 100 m diameter wind turbine. In this work, OMA techniques
were applied in order to extract the main vibration modes of the tower and rotor. For that purpose,
130
Chapter 5
several accelerometers were placed on the tower, nacelle, gearbox and generator. Several vibration
modes were identified and presented for different rotor speed regimes. A numerical aeroelastic model
was also developed and the results obtained were correlated with the experimental ones.
Ozbek et al (Ozbek, Rixen et al., 2010; Ozbek and Rixen, 2012) monitored the dynamic response of a
2.5 MW HAWT using conventional instrumentation (strain gauges) and optical measurement systems
(laser interferometry and photogrammetry). The wind turbine was tested under parked (both with
strain gauges and laser interferometry) and rotating (both with strain gauges and photogrammetry)
conditions. Although good results were achieved, two important drawbacks associated with the use of
optical measurement systems are referred: laser interferometry is not suitable for rotating conditions
testing; and the available camera systems used for the photogrammetry testing had a reduced memory,
allowing to record only small measurement segments.
Also Marinone et al (Marinone, Cloutier et al., 2014) performed both experimental and operational
modal analysis on two 225 kW wind turbines, in parked condition. A very complete instrumentation
was used, with 80 accelerometers installed on the blades, hub, nacelle and tower and 6 seismic
accelerometers installed at the foundation. The authors of the study were able to identify 18 vibration
modes in the 0 – 15 Hz frequency range and the correlation between the results from the EMA and
OMA was very good.
Apart from the entire wind turbine structure testing, blades properties are important features that
need to be identified (Yang and Sun, 2013). (Thomsen, Petersen et al., 2000) tried to identify damping
of edgewise blade modes of a 600 kW wind turbine under operation. For that purpose, the authors
installed eccentric rotating masses in the nacelle to excite the desired modes. Modal identification tests
are also often conducted under laboratory conditions (Griffith, Smith et al., 2006; Griffith and Carne,
2010; White, Adams et al., 2010).
Yang and Allen (Yang and Allen, 2012) also performed a modal analysis on a small wind turbine blade
based on a laser scanning technique. The test was performed under parked conditions and the authors
accomplished to identify some vibration modes (frequencies and mode shapes).
Another important field of application of OMA techniques on wind turbines is in offshore situations,
since different foundation designs lead to different structural behaviours (section 3.1.1.5). Moreover,
the contribution of the different types of damping is also difficult to predict (see section 4.5). Thus, the
extraction of the modal parameters under operating conditions reveals as an essential tool for the
correct comprehension of the dynamic behaviour of these structures.
Some work in this area has already been published. In (Devriendt, Magalhães et al., 2014), the modal
identification of a 3.0 MW offshore wind turbine with monopile foundation is introduced. This study
analyses the acceleration data recorded during 2 weeks (under non-operating conditions). The modal
results obtained were then confronted with collected environmental data, such as tidal level and wind
speed. This wind turbine is also study in section 5.9 of the present work.
In (Shirzadeh, Devriendt et al., 2013), the same wind turbine is experimentally tested. The main goal of
the test was to calculate the damping value of the fundamental fore-aft (FA) vibration mode. For this
purpose, both ambient and rotor-stop tests were performed. Rotor-stop tests (sometimes named
“overspeed stop tests”) are a type of forced vibration test, in which the turbine is operating at a normal
situation until a sudden pitch-out of the blades (Figure 5.1). In this procedure the wind turbine is
highly excited by the sudden absence of aerodynamic loads on the blades. Furthermore, the
131
Operational Modal Analysis of Wind Turbines
aerodynamic damping contribution from the rotor almost does not interfere in the decay response of
the structure, since the blades are pitched-out. The authors were able to identify the first pair of
vibration modes (in FA and SS directions) and the obtained value of the damping in the FA direction
was similar in both ambient and forced vibration tests. An aeroelastic model was also developed, in
which the two tests were simulated with good results.
0.8 90
Fore-Aft Acceleration [m/s2]
0.6 80
0.4 70
a) b)
Figure 5.1 – Example of fore-aft acceleration a) and blade pitch variation b) on a rotor-stop test (adapted from
(Damgaard, Ibsen et al., 2013))
In (Damgaard, Ibsen et al., 2013) the dynamic characteristics of offshore wind turbines with monopile
foundations are also studied. In this study, the value of damping of the fundamental mode for the side-
side direction was investigated. Once again, ambient vibration and rotor stop tests were performed
with the aim of extracting frequency and damping values for this vibration mode. In the analysis of the
results, some considerations are also made concerning the damping introduced by the soil.
Like the previous authors, (Versteijlen, Metrikine et al., 2011) studied the damping effect present in an
offshore wind turbine located in the Burbo Banks wind farm. 12 rotor stop tests were realized in order
to assess the influence of the soil.
In (Ibsen and Liingaard, 2006), a different type of offshore wind turbine is studied. A 3 MW wind
turbine with a prototype suction caisson foundation is tested in three different conditions: idling;
without blades; and without blades and nacelle. For the test, 15 accelerometers were installed along the
tower, and the structure was subjected to ambient excitation. The experimental results were compared
with a numerical model of the structure.
132
Chapter 5
1
10
0
10
r=0
r = 26.42m
-1
10 r = 54.40m
-2
10
0.001 0.01 0.1 1p 2p 3p 4p 1
Frequency [Hz]
Figure 5.2 – Normalized power spectral density of the temporal auto-correlation functions computed for points on the
same blade (for different radii cases) (adapted from (Tcherniak, Chauhan et al., 2010a))
133
Operational Modal Analysis of Wind Turbines
As can be seen, the power spectral density functions (for radii higher than 0) are not flat as in a
common wind spectrum (Kármán, 1948). Moving from the centre of the rotor to the blade tip, peaks
with a frequency equal to the frequency of rotation of the rotor (and its harmonics) become more
pronounced. This is motivated by the same phenomena exposed in Chapter 4:
o Every time a blade passes through a region with a determined wind speed (gust), the action on
the blade is somehow “similar” to the prior passage (considering the gust is long enough) –
this fact justifies the existence of the peaks;
o With the increase of radius of the point on the blade, the ratio between tangential speed of the
point and the mean wind speed increases – this fact justifies the increase of the peaks with the
increase of the radius of the point.
This is the reason why the second assumption presented earlier is violated. On the other hand, the
results obtained for the point with = 0, since it is stationary, do not present any peak. This is the
reason why this assumption is not violated under parked conditions.
The third assumption of OMA (spatial non-correlation of the input excitation) was also investigated in
(Tcherniak, Chauhan et al., 2010a). For this study, the authors computed the coherence ( )
between the wind speed fluctuations at different points ( and ) on the same and different blades:
| |
(5.1)
.
with:
Coherence
The results presented in (Tcherniak, Chauhan et al., 2010a) for points with different radii are shown in
Figure 5.3. It is once again evident that several peaks are presented at rotor frequency (and its
harmonics), leading to the violation of the third OMA assumption.
1
2
γAB (rA = 26. 42m , rB = 54. 40m , same blade)
2
0.8 γAB (rA = 26. 42m , rB = 54. 40m , diff. blades )
2
γAB (rA = 26. 42m , rB = 26. 42m , diff. blades )
AB
2
Coherence γ2
0.4
0.2
0
0 1p 2p 3p 4p 1 6p 7p 8p 9p 2 11p 2.5
Frequency [Hz]
Figure 5.3 – Coherence functions for 2 points under different study scenarios (adapted from (Tcherniak, Chauhan et al.,
2010a))
134
Chapter 5
One important observation that should be noted is the wide shape of the peaks in Figure 5.3 (and also
noted in Figure 5.2). This situation, in which the peaks have “thick tails”, leads to the spread of the
assumption violation for a quite wide frequency band around the harmonics. Contrary to a situation
with narrow peaks, in which the signal is only “polluted” in a very narrow band of frequency around
the harmonic, Figure 5.3 highlights the fact that the band between harmonics in which the coherence
is approximately 0 is very narrow. This fact may increase the difficulty in applying OMA to wind
turbines.
The last assumption, in which is referred that the forces have to be distributed over the entire
structure, is almost fully fulfilled. Although the top of the supporting structure is subjected to a
considerable higher loading when in operation (due to the rotor), the remaining part is also excited by
the wind. Hence, this assumption does not introduce any additional concern in the modal analysis of
wind turbines.
In spite of these difficulties, it is still possible to apply OMA to operating wind turbines. Section 5.4
describes the modal identification algorithms used in this work. Some possible solutions to overcome
the problem of the assumptions violation are presented in section 5.6.
135
Operational Modal Analysis of Wind Turbines
. .
(5.2)
. .
with:
State vector (which contains the displacement and velocity vectors of the system)
Input vector
State matrix
Input matrix
Output matrix
Among all the elements previously presented, a special reference should be made to the matrix since
it characterizes the structural system (it is possible to extract the modal properties of the system from
its eigenvalues and eigenvectors). This is the reason why the main goal of these algorithms is to
estimate this matrix.
136
Chapter 5
In situations in which the inputs are unknown, as is the case of OMA, the terms related to can be
grouped together with the noise terms. In this situation, equation (5.2) can be rewritten as:
.
(5.3)
.
The state-space model from equation (5.2) can also be defined in a modal basis. Considering a special
case of similarity transformation:
Ψ. , (5.4)
with:
, Λ. , Ψ .
(5.5)
. ,
with:
One important property of the modal state-space model is that, due to the diagonal structure of the Λ
matrix, it is possible to decouple the contribution of each vibration mode to the response of the
structure. In the same way, the matrix contains the observable components of the mode shapes in
each column, in accordance with the eigenvalues from the Λ matrix.
Due to these characteristics, it is possible to reduce the order of the state-space model in order to
consider only the desired modes. This property will be applied and discussed in section 5.5.2.3.
137
Operational Modal Analysis of Wind Turbines
The algorithm is based on the principal assumption that the correlation functions of the response of a
structure excited by white noise are defined by a sum of sinusoids with exponential decrement. Since
these sinusoids are related with the impulsive structural response, the dynamic characteristics of the
structure can be then extracted from these functions.
The method starts with the computation of estimates of the output correlation matrix for 2. -1 lags,
being an input of the method to be tuned by the analyst.
These correlation matrices are organized in a Toeplitz matrix:
…
…
| (5.6)
… … … …
. . …
with:
Refers to the reference sensors. Reference sensors are a subset of sensors that should present relevant
∗
modal ordinates of the modes to be identified. All sensors might be assumed as reference sensors.
Correlation matrix evaluated at time lag , between all sensors and reference sensors
Considering that the inputs are idealized as realizations of white noise processes and that the state
vector can be characterized by a stationary stochastic process with zero mean, each correlation matrix
can be factorized according to (Overschee and Moor, 1996):
. . (5.7)
with:
Considering the application of the property presented in equation (5.7) to the Toeplitz matrix, one
obtains:
.
| . . … . .Γ (5.8)
…
.
with:
138
Chapter 5
Computing the singular value decomposition of the same Toeplitz matrix, the following factorization
is obtained:
0
| . . . . . . (5.9)
0 0
Comparing now equations (5.8) and (5.9), it is possible to define O and Γ matrices in a different way:
. (5.10)
Γ . (5.11)
Considering that matrix A characterizes the dynamic behaviour of the structural system, and observing
equation (5.8), it is possible to state that, with the equalities from equations (5.10) and (5.11), the
identification problem is solved. Since it is known that the matrix is defined by the first rows of the
matrix ( represents the number of measured outputs), the matrix can be obtained through several
ways. In this work, the following formulation was implemented:
. .
. . . .
. ⟺ . (5.12)
… … … …
. . . .
with:
∗ Moore-Penrose pseudo-inverse
Computing the eigenvalues ( ) and eigenvectors decomposition of matrix , the results in discrete
time are obtained:
A Ψ. Λ. Ψ (5.13)
with:
139
Operational Modal Analysis of Wind Turbines
The natural frequency values and modal damping ratios for each vibration mode are then obtained
taking into account the conversion of the eigenvalues from discrete time to continuous time model:
ln
∆ (5.14)
2.
ln
∆
ln (5.15)
∆
with:
∗ Absolute value of *
∗ Real part of *
With the values of the frequency and damping identified, the only missing result is the mode shape of
each vibration mode. The observable components of each mode shape are thus obtained with:
. (5.16)
with:
Matrix with the mode shape configurations at the measured degrees of freedom
Some further considerations should be mentioned concerning the application of the SSI-COV
algorithm to real structures. Considering the system identification of theoretical structures, the
application of the Singular Value Decomposition to the Toeplitz matrix, leads to singular values equal
(or very close) to zero after a certain order (see equation (5.9)). In other words, there should be a clear
gap between singular values that defines the order of the model. However, in real structures, due to the
noise contained in the signals and due to the fact that the used correlation functions are estimates,
there is not a clear gap between singular values. Thus, it is common to calculate the results for different
orders of the matrices and . With this method, a group of modal results is obtained for each order.
Then, these results are compared with the results from the previous model order by analysing the
variation between the estimated modal parameters. If the variation is within a previously defined limit,
the result (or pole) is considered stable. The results obtained for different orders are usually presented
in a stabilization diagrams as illustrated in Figure 5.5.
Example
The SSI-COV algorithm was applied to the acceleration time series obtained with the numerical model of
the NREL 5MW wind turbine introduced in section 4.7. For this analysis, the three sensors were defined
140
Chapter 5
as reference and 257 points were considered for calculation of the correlation functions (2. -1 = 257). The
obtained correlation functions are presented in Figure 5.4.
-3 R11 -3 R12 -4 R13
x 10 x 10 x 10
3 5
2
5
Amplitude
Amplitude
Amplitude
1
0 0
0
-1
-2
-5
-3 -5
0 1 2 3 4 5 0 1 2 3 4 5 0 1 2 3 4 5
Time [s] Time [s] Time [s]
Amplitude
Amplitude
1
0
0
0
-5 -1
-5
0 1 2 3 4 5 0 1 2 3 4 5 0 2 4
Time [s] Time [s] Time [s]
-4 R31 -3 R 32 -3 R33
x 10 x 10 x 10
2 4
5
1 2
Amplitude
Amplitude
Amplitude
0 0 0
-1 -2
-5
0 1 2 3 4 5 0 1 2 3 4 5 0 1 2 3 4 5
Time [s] Time [s] Time [s]
141
Operational Modal Analysis of Wind Turbines
The modal damping of the stable poles is shown in Figure 5.6. In this figure, the poles related with the two
considered vibration modes are also highlighted by red boxes.
40
Model Order
30
20
10
0
0 0.5 1 1.5 2 2.5 3
Frequency [Hz]
Figure 5.5 – Stabilization diagram obtained with the SSI-COV method (average spectra at the background)
8
Damping ratio [%]
0
0 0.5 1 1.5 2 2.5 3
Frequency [Hz]
Figure 5.6 – Modal damping ratio estimates for the stable poles
The identified modal properties from the two vibration modes are shown in Table 5.1.
142
Chapter 5
The results obtained for the natural frequency are in a good agreement with the reference values. On the
other hand, the identified damping ratio for the 1st vibration mode is considerably higher than the value
obtained with the HAWC2 eigensolver. This discrepancy is due to the fact that the identified damping is
related to the effective damping of the mode during the 600 s. Thus, besides the structural damping, the
identified value also includes the aerodynamic component of the damping. The results obtained with the
SSI-COV algorithm clearly show the influence of the aerodynamic damping on the 1st vibration mode of
the tower. This effect is not relevant for the 2nd vibration mode of the tower.
. . (5.17)
. . . . (5.18)
. . . . . . . . (5.19)
with:
The Kalman filter state sequence organizes the Kalman filter state estimates obtained with equations
(5.17) to (5.19):
… (5.20)
143
Operational Modal Analysis of Wind Turbines
The version of SSI-DATA algorithm implemented in this work is presented in (Peeters, 2000), which
in turn is based on (Overschee and Moor, 1996).
The SSI-DATA algorithm starts with the organization of the data in an Hankel matrix:
…
…
… … … …
1 ↕ . " "
… |
(5.21)
√ … | .
↕ . " "
…
… … … …
. . … .
with:
It is common to divide the Hankel matrix into a past reference and a future part, the upper and lower
part of the matrix, respectively. Considering this division, the projection operation can then be defined
as:
/ (5.22)
The definition of the is a very important aspect of the algorithm. In fact, the SSI-DATA algorithm
relies on fact that the can be factorized as product of the extended observability matrix and the
Kalman filter state sequence :
.
. . … (5.23)
…
.
The most computational efficiency way to calculate is through the application of the 1
. (5.24)
1
In a factorization, the is an orthonormal matrix ( . . ) and is a lower triangular matrix.
144
Chapter 5
Omitting the columns filled only with zeros in the matrix, the and matrices can be organized as
follow:
. 1
↔ ↔ ↔ ↔ ↔
. ↕ 0 0 0 ↕ .
↕ 0 0 ↕
. (5.25)
↕ 0 ↕
1 ↕ ↕ 1
0
. . . . . . (5.26)
0 0
/
. (5.27)
. (5.28)
On the other hand, the projection of future row spaces into past row spaces can be expressed with the
help of the and submatrices:
. (5.29)
. (5.30)
The extended observability matrix is obtained after deleting the last rows of , while is
given by:
. (5.31)
145
Operational Modal Analysis of Wind Turbines
At this point, the system matrices can be estimated through the application of the system of equations
from the state-space model (see equation (5.3)):
. (5.32)
|
where the | matrix can be defined resorting, once again, to the and submatrices:
0
| . (5.33)
Once all matrices all calculated, equation (5.32) can be used to estimate matrices and . Since it is an
overdetermined problem, it can be solved through:
. (5.34)
|
Once state-space matrices and are obtained, the modal properties of the structural system can be
calculated with the already presented equations (5.13) to (5.16).
An important property of the SSI-DATA algorithm is that it allows to recover the noise covariance
matrices , and from the state-space model equations:
. (5.35)
Like in the case of the SSI-COV, it is common to calculate the state-state matrices ( and ) for
different orders. Then a similar strategy is followed, where the results from the different orders are
compared in order to identify the stable poles.
Example
The simulated acceleration time series from the NREL 5MW wind turbine were used to illustrate the SSI-
DATA algorithm. For this analysis, three sensors were considered as reference and a value of = 80 was
used.
The stabilization criteria defined in the SSI-COV example were also used in this example. The obtained
stabilization diagram is presented in Figure 5.7. The damping values of the stable poles are shown in
Figure 5.8.
146
Chapter 5
40
Model Order
30
20
10
0
0 0.5 1 1.5 2 2.5 3
Frequency [Hz]
Figure 5.7 – Stabilization diagram obtained with the SSI-DATA method (average spectra at the background)
8
Damping ratio [%]
0
0 0.5 1 1.5 2 2.5 3
Frequency [Hz]
Figure 5.8 – Modal damping ratio estimates for the stable poles
The values of natural frequencies and damping ratios of the two founded vibration modes are shown in
Table 5.2. The values are very similar to those previously obtained with the SSI-COV algorithm. Once
again, the contribution of the aerodynamic damping is noticeable for the 1st vibration mode.
147
Operational Modal Analysis of Wind Turbines
5.4.2 P-LSCF
The third and last identification method covered in this work is the poly-reference Least Squares
Complex Frequency Domain method (p-LSCF), also known as Polymax. This is a parametric,
frequency-domain method which was initially developed in order to identify the modal characteristics
of a system from its frequency response functions (Guillaume, Verboven et al., 2003; Peeters,
Auweraera et al., 2004). Later, the previously developed methodology was adapted by (Peeters and
Auweraer, 2005) to work based on the output half-spectrum functions, an output-only version of this
method. In this work, only this last version is presented.
5.4.2.1 Half-Spectrum
Prior to the application of the p-LSCF routine, the half spectrum matrix should be estimated.
Considering the input excitation as white noise, the input spectrum ( ) is constant and not
dependent on frequency, and the output spectrum is obtained by:
. . (5.36)
Considering the modal decomposition of the frequency response function (Heylen, Lammens et al.,
2007),
∗
. .
. ∗ (5.37)
.
with:
∗
System poles ( , . . 1 . )
it is possible to define the output spectrum as a summation of the contribution of structural vibration
modes (Peeters, 2000):
∗ ∗
. . . .
. ∗
. ∗ (5.38)
. .
with:
148
Chapter 5
As can be observed, the previous equation has four poles for each mode which impose the use of
models with twice the order of the system defined by equation (5.37). For this reason, the positive (or
half) spectrum is usually used:
∗
. .
. ∗ (5.39)
.
Since the frequency response function (5.37) and the half spectrum (5.39) present similar modal
decomposition, they both can be parameterized in exactly the same way (Peeters and Auweraer, 2005).
In this work, the Correlogram approach, restricted to the positive time lags of the correlation function,
is used for the estimation of the half spectrum:
0 . . .
.Δ (5.40)
2
with:
With the Correlogram approach, an exponential window should be applied to the correlations before
the use of equation (5.40). This window is important mainly for two reasons: to reduce the effect of
leakage and to reduce the influence of the higher time lags, which have a larger variance (Peeters and
Auweraer, 2005). The used exponential window follows the equation:
. .
, for 1 (5.41)
with:
The application of the exponential window leads to biased values of damping for the computed poles.
For this reason, the damping values should be corrected according to (Cauberghe, 2004):
(5.42)
149
Operational Modal Analysis of Wind Turbines
parameterizing the transfer function as a right division of two polynomial matrices and
(Reynders, 2009):
(5.43)
. . .
with:
, Polynomial matrices
Complex variable
For the implemented version of the p-LSCF method, it is of great importance to refer the conversion
from a RMRD model to a state-space model (Reynders, 2009). Assuming that matrices and
have the same order ( ) and that the degree of the polynomial with the determinant of s is equal
to . (where refers to the number of structure inputs), the matrices , , and from the state-
space model are obtained through (the subscript ∗ is dropped for simplification):
. . … . .
0 … 0 0
⋮ ⋱ … ⋮ ⋮
0 0 … 0
(5.44)
0
. . … . .
with:
Identity matrix
150
Chapter 5
. . . . . .
. . . . (5.45)
Bearing in mind that these matrices contain the modal parameters of the dynamic system, the main
goal of this method involves the calculation of and that reduce the approximation error
associated to equation (5.45). The nonlinear least squares problem can be simplified to a linear
problem according to (Guillaume, Verboven et al., 2003):
. . . . . .
. . . (5.46)
Considering the polynomial basis functions evaluated at each frequency (ω ) organized in one row
with ( 1) components:
Ω Ω Ω … Ω . . . . . .
… . . . (5.47)
Ω . Ω . Ω . … Ω . . (5.48)
⋮ ⋮
with:
Line of the , or matrix (which varies from 1 to the number of measured outputs - )
151
Operational Modal Analysis of Wind Turbines
⋮ (5.49)
, with 1, 2, … , (5.50)
⋮
, . (5.51)
where:
Ω
⋮ (5.52)
Ω
Ω … Ω ⊗
… , with 1, 2, … , (5.53)
Ω … Ω ⊗
with:
Expanding equation (5.51) with all the parameters from matrices α and β, one obtains:
0 … 0
0 … 0
⋮ ⋮ ⋱ ⋮ ⋮ . ⋮ . (5.54)
0 0 …
152
Chapter 5
The solution for the identification problem can be solved through . 0. However, this is very
computationally demanding. For this reason, a different approach is usually followed. It consists on
determining the model parameters through a linear least squares cost function obtained by adding all
the squared elements of the matrix evaluated at each discretized frequency:
∗
, . , (5.55)
with:
, , . , . .
(5.56)
. ; . ; .
with:
,
2 . . 0 (5.57)
,
2 . . 0 (5.58)
Eliminating the unknowns , the size of the system of equations can be reduced:
2 . . 0⇔ . . (5.59)
2 . . 0⇔ . 0 (5.60)
153
Operational Mod
dal Analysis of Wind
W Turbines
. 0 ⇔ . 0 ⇔ . ⇔
. ⋮ ⋮
0 (5.61)
. ⇒
.
Once α is d
defined, it iss possible to compute thhe matrixx using equattion (5.59), w
which defines the
RMFD mod del.
The next steep in the algo
orithm is thee definition oof the companion matrix:
. . … . .
0 … 0 0
(5.62)
⋮ ⋱ … ⋮ ⋮
0 0 … 0
If matrix from equatiion (5.44) is compared too matrix from equatiion (5.62) it is easily obsserved
their similarrity. Thus, representss the matrix from the state-space
s model.
m The reemaining maatrices
from the moodel can be computed
c through equatiion (5.44).
With the sttate-space model
m defined
d, the modall parameters of the dynaamic system can be calcu
ulated
using, once again, the eq
quations (5.1
13) to (5.16).
As referred in the case of
o the SSI-CO OV algorithm
m, there is no wledge about what model order
o prior know
( ) will prooduce the beest results. For
F this reasson, several orders are computed
c annd the resullts are
usually pressented in a sttabilization diagram.
d
Exam
mple
The aapplication of the p-LSCF algorithm
a to tthe simulated
d acceleration signals obtainned for 3 degrrees of
freedoom of the NR REL 5MW wiind turbine m model was peerformed from m correlation functions with 256
pointss. Then, the half-spectrum
h functions werre obtained foor subsequentt estimation oof matrices anda .
Thesee functions aree shown in Figgure 5.9.
154
Chapter 5
-4 -4 -5
Amplitude
Amplitude
Amplitude
10 10 10
-5 -5 -6
10 10 10
-6 -6 -7
10 10 10
0 1 2 3 4 5 0 1 2 3 4 5 0 1 2 3 4 5
Time [s] Time [s] Time [s]
-4 -4 -5
Amplitude
Amplitude
Amplitude
10 10 10
-5 -5 -6
10 10 10
-6 -6 -7
10 10 10
0 1 2 3 4 5 0 1 2 3 4 5 0 1 2 3 4 5
Time [s] Time [s] Time [s]
-5 -5 -4
Amplitude
Amplitude
Amplitude
10 10 10
-6 -6 -5
10 10 10
-7 -7 -6
10 10 10
0 1 2 3 4 5 0 1 2 3 4 5 0 1 2 3 4 5
Time [s] Time [s] Time [s]
The stabilization diagram for models with order between 2 and 35 was created considering the same
stabilization criteria defined for the SSI-COV and SSI-DATA algorithms (Figure 5.10). The modal
damping ratio results associated with the stable poles are shown in Figure 5.11.
30
25
Model Order
20
15
10
0
0 0.5 1 1.5 2 2.5 3
Frequency [Hz]
Figure 5.10 – Stabilization diagram obtained with the p-LSCF method (average spectra at the background)
155
Operational Modal Analysis of Wind Turbines
8
Damping ratio [%]
0
0 0.5 1 1.5 2 2.5 3
Frequency [Hz]
Figure 5.11 –Modal damping ratio estimates for the stable poles
The final results obtained with p-LSCF algorithm are presented in Table 5.3. The results are similar to
those previously obtained with the other two identification algorithms.
156
Chapter 5
. .
(5.63)
.
with:
State vector
Innovation sequence
The innovation sequence is a white noise vector, with the covariance matrix defined as:
. . (5.64)
The forward innovation model is thus defined with the matrices and , in addition to the state-
space matrices A and C already defined. These matrices can be obtained through the solution of the
discrete Riccati equation [(Peeters, 2000)]:
. . . . . . . . (5.65)
with:
157
Operational Modal Analysis of Wind Turbines
After calculation of the forward state covariance matrix, the covariance matrix of the innovation
sequence is obtained as:
. . (5.66)
. . (5.67)
Σ . Σ. (5.68)
. Σ. (5.69)
. Σ. (5.70)
Once these matrices are obtained, it is possible to define a forward innovation model (equation (5.63))
through the application of equation (5.65) to (5.67).
158
Chapter 5
The forward innovation model from equation (5.63) can then be defined in the modal basis:
, Λ . , .
(5.71)
. ,
with:
The advantage of using the model in a modal basis is that, since Λ is a diagonal matrix, ,
represents the separate contribution of each vibration mode to the total response signal. Rearranging
equation (5.71), it can be written as:
, Λ . , .
(5.72)
. ,
At this point, it is possible to use the system of equations (5.72) as a state-space model in order to
estimate the modal state sequence , and the innovation sequence from the measured
acceleration . The measured acceleration can then be reconstructed as the sum of an estimated
response part ( ) with an estimation error part ( ):
. , (5.73)
In the previous equation, represents the estimated contribution to the response from the vibration
modes, while represents the estimation error, i.e., the part of the response that cannot be explained
by the stochastic state-space model.
In addition, the acceleration signal estimation can be also thought as a sum of contributions from
single vibration modes. The estimated modal response can be split in single modal responses as:
. , (5.74)
with:
Column vector with the observable components of the mode shape of the -th vibration mode
, Array vector from the modal state vector corresponding to the -th vibration mode
159
Operational Modal Analysis of Wind Turbines
Naturally, is a complex modal response. However, if the complex modal response of the pair of
complex-conjugated poles from the same vibration mode is combined, the real modal response is
obtained.
. (5.75)
|
. (5.35)
160
Chapter 5
In order to evaluate the stability of the poles and exclude the unstable poles (i.e. associated with
negative modal damping ratios), the unit circle can be used. The unit circle limits the stable region of
the poles in the -domain. The radius of pole in the unit circle is given by:
. . (5.76)
with:
(imaginar y axis)
j i stable pole
t
unstable pole e j j
e i i t
t (real axis)
t
e i i
j e j j
Figure 5.12 – Identification of a stable and unstable pole in the unit circle
Since it is desired to only account for the stable poles (and the respective vibration modes), it is
necessary to eliminate the unstable poles from the and state-space model matrices.
In this work, a truncation based methodology of the modal state-space model is used. The identified
and matrices are initially transformed into the modal basis:
Λ Ψ . .Ψ (5.77)
.Ψ (5.78)
As referred in section 5.4.1.1, the Λ is a diagonal matrix with the eigenvalues of the system and is a
matrix with the observable components of the vibration modes in the columns. After the
transformation from equations (5.77) and (5.78), the eigenvalues corresponding to the unstable poles
can now be removed, deleting the corresponding lines and columns from matrix Λ. Similarly, the
columns from the matrix can now be removed in accordance to the deleted eigenvalues. This
truncation operation is illustrated in Figure 5.13.
161
Operational Mod
dal Analysis of Wind
W Turbines
…
Λ V
uncation is complete,
After the tru c thee state-space model can be
b recovered from the moodal basis through
the inversioon of the equ
uations (5.77
7) and (5.78)). After this step,
s it is possible to folllow the proccedure
presented foor the SSI-CO
OV algorithmm in section 5.5.2.2.
The validatiion of this prrocedure is illlustrated thrrough a simp
ple example in
i the Appenndix A.
Exam
mple
With the aim of illlustrate the methodology
m presented to
o decompose the acceleratiion time seriees into
modaal responses, thhe procedure was applied too the accelerattion time seriees numericallyy generated with
w the
NREL L 5MW wind turbine modeel. For that puurpose, the reesults obtained d with the appplication of th
he SSI-
DATA A algorithm (presented in n the examplle of the secttion 5.4.1.4) were used aas reference for f the
validaation of the po
ost-processingg tool applied w
with the otherr identification
n algorithms.
A statte-space modeel of order 50,, obtained withh the SSI-DATA, was used to define the forward inno ovation
modeel. Since the reesonance response of the w wind turbine att the harmoniics frequenciees was identifieed as a
regulaar vibration mode
m (Figure 5.7), it is alsoo possible to estimate
e the response
r of thhe structure at these
frequeencies.
Figuree 5.14 presentts the modal responses
r of toop sensor S3 for
f the two traacked vibratioon modes (1 FAF and
2 FA) and the two first
f harmoniccs (3Ω and 6Ω ). The responnses obtained illustrate
i importance off the 1st
the im
F and the 3Ω harmonic eexcitation to th
vibrattion mode (1 FA) he motion at the
t top of the w wind turbine tower.
162
Chapter 5
1 FA 1 FA
0.2 0.05
SSI-DATA SSI-DATA
0.15 SSI-COV
p-LSCF
0.1 0.025
0.05
Acc. [m/s 2 ]
Acc. [m/s 2 ]
0 0
-0.05
-0.1 -0.025
-0.15
-0.2 -0.05
0 100 200 300 400 500 600 375 380 385 390 395 400
Time [s] Time [s]
3Ω 3Ω
0.2 0.15
0.15
0.1
0.1
0.05
0.05
Acc. [m/s2]
Acc. [m/s2]
0 0
-0.05
-0.05
-0.1
-0.1
-0.15
-0.2 -0.15
0 100 200 300 400 500 600 375 380 385 390 395 400
Time [s] Time [s]
6Ω 6Ω
0.2 0.05
0.15
0.1 0.025
0.05
Acc. [m/s2]
Acc. [m/s2]
0 0
-0.05
-0.1 -0.025
-0.15
-0.2 -0.05
0 100 200 300 400 500 600 375 380 385 390 395 400
Time [s] Time [s]
2 FA 2 FA
0.2 0.01
0.15
0.1 0.005
0.05
Acc. [m/s2]
Acc. [m/s2]
0 0
-0.05
-0.1 -0.005
-0.15
-0.2 -0.01
0 100 200 300 400 500 600 375 380 385 390 395 400
Time [s] Time [s]
Figure 5.14 – Modal responses of the top sensor S3 (only the 1 FA and 2 FA modes and the 3Ω and 6Ω harmonics were
considered)
163
Operational Modal Analysis of Wind Turbines
(5.73)
with:
Equation (5.73) can be also written as a sum of the estimated vibration mode acceleration:
(5.73)
with:
Time series of the estimated modal response of the -th vibration mode
Multiplying each element from the previous equation by the transpose of and retaining only the
diagonal elements of the matrices, the following is obtained:
. . . ⇒
(5.79)
. . . ⋯ . .
with:
164
Chapter 5
In the previous equation, the diagonal elements of . and . represent, respectively, the
covariance between the estimated modal response from mode and the original signal and the
covariance between the estimation error and the original signal. Normalizing equation (5.79), one
obtains:
. . . . . . . . . ⇒
(5.80)
1 Δ Δ ⇒ 1 Δ Δ ⋯ Δ Δ
with:
Δ Vector with the contribution of the estimated modal response of the -th vibration mode
The procedure presented so far is referred to an acceleration signal from a single sensor. Thus, the
contribution of each vibration mode to the original signal is only referred to that single sensor.
However, it is common to have more than one sensor installed on the structure. In that case, a global
assessment of the modal contribution can be obtained with the mean value of the Δ and Δ vectors
obtained from the l installed sensors:
1
Δ (5.81)
1
Δ (5.82)
with:
Example
The acceleration signals obtained with the NREL5 MW wind turbine model were once again used to
illustrate the procedure presented in this section.
In that sense, the modal responses obtained with the forward innovation model defined with the state-
space model of order 50 from the SSI-DATA algorithm was used to estimate the modal contribution to
the recorded acceleration from the 3 sensors. Once again, only the 1 FA and 2 FA modes and the two first
rotor harmonics were considered.
Figure 5.15 presents the contribution of each considered modes/ harmonic excitations for the three
sensors, together with the estimated error (Δ ). It should be noted that the sum of the presented
165
Operational Modal Analysis of Wind Turbines
contributions (including error) is not equal to 1, since some residual contributions from the other modes/
harmonic excitations are not presented in the figure.
The acceleration time series recorded by the sensors S1 and S2 are clearly dominated by the 2 FA. It is an
expected result because the mode shape of this mode presents higher modal amplitudes at the positions of
these sensors than the 1 FA mode. On the other hand, the acceleration signal captured by the S3 sensor is
dominated by the 3Ω harmonic excitation and, in a smaller scale, by the 1 FA mode. This is justified by
the fact that motion configuration of this mode/ harmonic presents its higher modal amplitude at the top
of the tower.
0.5
0.45 S1
S2
0.4 S3
0.35
0.3
0.25
0.2
0.15
0.1
0.05
0
1 FA 3Ω 6Ω 2 FA ∆e
Figure 5.15 – Modal contributions from the 1 FA and 2 FA modes, 3Ω and 6Ω harmonics and error to the acceleration
signals obtained for sensors S1, S2 and S3 with the NREL 5MW wind turbine model
166
Chapter 5
′ ,
.
′ ,
(5.83)
with:
Rotation matrix
Angle between the tower sensors referential and nacelle orientation (see Figure 5.16)
This transformation leads to the projection of the collected acceleration time series in the FA and SS
directions (Figure 5.16).
167
Operational Modal Analysis of Wind Turbines
Y
Y’ ≡ SS X’ ≡ FA
Figure 5.16 – Illustration of the angle between the sensors referential and nacelle orientation (θ)
168
Chapter 5
The referred violation can be faced through Multi-Blade Coordinate (MBC) transformation (also
named as Coleman transformation) (Hansen, 2007; Bir, 2008). The MBC transformation helps
integrate the dynamics of individual blades into the whole structural system, converting the motion of
each blade in the rotating blade frame into the ground-fixed frame (the same of the tower coordinate
system). With this transformation, the periodic terms in the equations of motion are eliminated
(Hansen, 2007). The MBC transformation is defined, for the most common case (3-bladed rotor wind
turbines), as (Bir, 2008):
1
3
2
. cos
3 (5.84)
2
. sin
3
with:
Azimuth location of the -th blade: Ω. 1 . (it is assumed that for 0, the blade is
vertically up)
The new degrees of freedom in the ground-fixed frame ( , and ) represent now a cumulative
behaviour of all rotor blades, instead of each individual blade ( ). Two examples of the physical
meaning of the new coordinates is presented in (Hansen, 2007), for operating conditions with a low
value of blades pitch angle:
If → flapwise blade deflection shape (Figure 5.17) If → edgewise blade deflection shape
→ simultaneous flapwise motion of all blades → simultaneous edgewise motion of all blades
b1
a0 a1
Figure 5.17 – Multi-blade coordinate when qi represents flapwise motion (Skjoldan, 2011)
169
Operational Modal Analysis of Wind Turbines
The inverse transformation (from ground-fixed coordinates to the rotating frame) is given by:
In (Tcherniak, Chauhan et al., 2010b), a routine for the application of MBC transformation to records
from sensors located on the rotor is proposed:
1. The signals of points located on the blades and tower (and also other locations) and the
azimuth angle are acquired as time histories;
2. The signals data from the rotating parts are subjected to MBC transformation using the
azimuth data. The signals data from the non-rotating components (tower and nacelle) are not
transformed;
3. The new signals obtained with MBC transformation and the signals from non-rotating
components are the inputs to OMA identification algorithms;
4. The frequencies and damping values are determined;
5. The resulting mode shapes are subjected to inverse MBC transformation (5.85); the results can
be directly animated overlaid by the rotor rotation.
Although MBC is an useful tool to eliminate the periodic effect of the rotor, there are some practical
aspects that may complicate its application (Ozbek, Meng et al., 2013):
o If the measurements are only taken from one blade, the transformation cannot be applied;
o The measurements must be taken at equal radial locations for all the blades;
o The application of the MBC transformation requires the knowledge of the instantaneous
position of the blade (azimuth angle).
Two last comments should be made about the use of MBC transformation to wind turbines. There is
one fundamental assumption of this method (verified for almost all wind turbines): the rotor has to be
isotropic, which means, the all the blades are identical, identically pitched and symmetrical mounted
on the hub (Hansen, 2003). Another important aspect about the effects of utilization of this technique
is that, although it removes the periodicity from the matrices, it does not eliminate the periodicity
character of the excitation forces.
170
Chapter 5
̅
| ,
(5.86)
with:
Expectation value
where > 3, = 3 and < 3 define distributions as leptokurtic (distribution more peaked than
normal distributions), mesokurtic (normal distribution) and platykurtic (more flat than normal
distribution), respectively. It is common to express this measure in the form of excess kurtosis:
∗
̅
3 (5.87)
| ,
∗
With this designation, leptokurtic, mesokurtic and platykurtic distributions will be defined as > 0,
∗
= 0 and ∗ < 0.
This kurtosis can also be used as an indicator of the presence of harmonic components (Brincker,
Andersen et al., 2000):
o It is assumed from the central limit theorem that any linear combination of a large number of
random variable tends to a normal distribution independently of the statistical properties of
the variables. Thus, the stochastic structural response, due to loading from a sufficient large
number of independent sources, will be approximately Gaussian (see Figure 5.18 for the
Gaussian probability function). For this reason, the value of excess kurtosis will be ∗ = 0
(mesokurtic distribution);
o In the case of a pure harmonic, the probability function presents a characteristic shape with
two symmetric peaks approaching infinity (Figure 5.18). As result, it will present a value of
excess kurtosis of ∗ = - 1.5.
171
Operational Modal Analysis of Wind Turbines
PDF
1
Normal
Harmonic
0.8
0.6
0.4
0.2
0
-3 -2 -1 0 1 2 3
Figure 5.18 – Normalized probability density function of a pure structural mode (blue line) and pure harmonic
component (red dashed line)
Based on these properties, some methodologies were proposed (Jacobsen, Andersen et al., 2006;
Andersen, Brincker et al., 2007; Jacobsen, Andersen et al., 2008). However, some difficulties are
expected regarding the application of these methods to the modal identification of wind turbines. In
fact, it is referred that when the harmonic is located closed to the natural frequency of a structural
mode, the results may deteriorate (Jacobsen, Andersen et al., 2008).
log (5.88)
where:
| |. .
(5.89)
with:
172
Chapter 5
and, consequently:
log ln | | . (5.90)
Equation (5.88) will result in the “complex cepstrum”. On the other hand, if the phase of is set to
zero, “real cepstrum” is obtained.
The main steps of this methodology are presented in Figure 5.19. It consists on the combination of the
edited “real cepstrum” (where influence of the harmonics is removed) with the unmodified phase
information. The complex spectrum is then retrieved by exponentiation.
Input FFT + Edited log Exp. Comple x
Phase
signal spectrum spectrum
+ IFFT
Log Tim e
amplitude signal
IFFT
Real
cepstrum
Edit
Edited
cepstrum
FFT
Edited log
amplitude
cepstrum
Figure 5.19 – Algorithm of the Cepstral method used to eliminate the effect of harmonics in signals (adapted from
(Randall and Sawalhi, 2011))
The application of this methodology to vibration signals from wind turbine is described in the
literature. The results indicated that the use of the cepstral method is still not a definite solution to
solve this problem. In some cases (Manzato, White et al., 2013), the application of the cepstral method
led to an improve of the results, while in other situations (Manzato, Devriendt et al., 2014), its
application spoiled the modal identification.
173
Operational Modal Analysis of Wind Turbines
Frequency
Frequency
Frequency
Time Time Time
a) b) c)
Figure 5.20 – Graphical representation: a) sine with constant frequency; b) sine with increasing frequency with constant
Δ ; c) sine with increasing frequency and constant Δ (adapted from (Blough, 2006))
Time synchronous averaging represents an useful tool to identify a varying frequency of excitation
along time. As example, Peeters et al (Peeters, Cornelius et al., 2007) refers a successfully case study in
which this technique was applied to an helicopter.
However, the application of this method for the analysis of vibration signals from wind turbines
presents some drawbacks, including the need to know the position or frequency/ radial speed of the
rotor with very good precision, which implies the installation of additional sensors at the turbine.
Other important difficulties are also referred in literature regarding the application of TSA to wind
turbines. Tcherniak et al (Tcherniak, Chauhan et al., 2010a) refers that it should not work properly
since this method suits better in situations where the harmonics are identified as well-defined sharp
peaks. As shown in section 5.3, the peaks obtained from the harmonic excitation present the
previously introduced “thick tail” characteristic. This difficulty was experimentally confirmed in
(Manzato, White et al., 2013).
Although different methodologies to identify and remove the effect of harmonics in vibration signals
have been developed, their application to wind turbine signals have not revealed as a definite solution
for this problem. In that sense, other types of strategies are usually employed, as shown in Chapter 7.
174
Chapter 5
- Initial seed
E
- Poles
B
- Cluster
A
A C
H D
G
F
175
Operational Modal Analysis of Wind Turbines
The last described strategy is the hierarchical clustering analysis. With this strategy, each initial pole is
considered a cluster. Then, and in each step, the closest poles are aggregated into the same cluster until
all poles are combined in one large cluster (Figure 5.22). The algorithm of a hierarchical clustering
methodology is as follows:
1. Definition of distance between poles. This property will measure the closeness between poles;
2. Calculation of similarity between every clusters. The criteria for similarity calculation has to be
chosen (the more common are single link, complete link, group average link, Ward’s method,
centroid method and median (Gordon, 1999));
3. Linking of the clusters in an hierarchical tree;
4. Definition of a rule to cut the hierarchical tree. This rule will define the final number of
clusters.
It should also be referred that, contrary to what can happens in non-hierarchical methods, a pair of
poles grouped together at a certain step cannot be separated in subsequent steps. Pappa et al (Pappa,
III et al., 1997) is said to be the first author to report the utilization of this kind of strategy for
automatic system identification (Reynders, Houbrechts et al., 2012). Other descriptions are presented
in (Verboven, Guillaume et al., 2003) and (Magalhães, Cunha et al., 2009) with examples of application
with the LSCF and SSI-COV method, respectively.
A,B,C,D,E,F,G,H
cut level - Cluster
- Final Cluster
A,B,C,D E,F,G,H
A,B F,G,H
F,G
C,D
A B C D E F G H
176
Chapter 5
pole property for calculation of similarity between poles was dropped. Thus, the frequency value and
the mode shape (more specifically, the MAC value) were the properties used for distance calculation
between poles (Magalhães, Cunha et al., 2009):
1 , (5.91)
with:
The different parameters of equation (5.91) can be computed with weighting factors, which allows
assigning different importance to each property. However, this possibility was not used in the
monitoring system. As can be concluded from the observation of equation (5.91), the shorter the
distance is, the greater the similarity between poles and the higher the probability that both are
referred to the same vibration mode.
After the calculation of the distance between poles, it is necessary to decide a criterion to assess the
similarity between clusters. In the implemented strategy, the single link was used. This criterion
defines the distance between two clusters as the smallest distance between poles contained in different
clusters. This distance is the same as defined in equation (5.91).
Once the hierarchical tree is constructed, it is necessary to define another criterion for the cut level.
This decision cannot be based on a hard criterion, like a final number of clusters: it is not possible to
know how many clusters will be formed containing spurious poles; besides, there are structural modes
which will not be detected in every setup. These aspects introduce an uncertainty about the final
number of physical and spurious clusters. Thus, the chosen criterion imposes a degree of homogeneity
inside the cluster: a maximum distance between any point and its closest point in the same cluster.
However, this maximum distance has to be carefully evaluated, since it can lead to the split of similar
poles into different clusters (if the distance is too tight) and the aggregation of different poles into the
same cluster (if the distance is too large).
After the application of the previous criterion, it is necessary to pick the final clusters containing
physical poles. Considering that poles referring to physical modes are consistent along the different
orders of the system identification algorithms, an initial cut-off criterion should be implemented to
eliminate clusters with a reduced number of poles. Then, the modal properties of the remaining cluster
can be compared with the reference properties of the vibration modes intended to be tracked. These
reference properties can be defined during an initial training period.
Example
The results obtained with the application of the SSI-COV algorithm to the acceleration time series from
the NREL 5MW wind turbine model were used to illustrate the implemented hierarchical cluster
methodology. In this example, the maximum distance between poles was set to 0.02 and a minimum
number of poles in a cluster in order to be considered was defined as 5. The results obtained with the
177
Operational Modal Analysis of Wind Turbines
cluster analysis are presented in Figure 5.23, where the considered cut level of 5 poles is represented by the
dashed red line).
The natural frequency and damping ratio of the poles found after the consideration of the cut level are
shown in Figure 5.24. In this figure, the small circles represent the properties of the stable poles, while the
large circles indicate the mean values of the poles inside each cluster.
25 9
2 FA 8
20 7
3Ω 5
1 FA 4
6Ω
10
3
2
5
1
0 0
0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3
Frequency [Hz] Frequency [Hz]
Figure 5.23 – Number of poles obtained for each cluster Figure 5.24 – Natural frequency and damping ratio of the
defined according to the implemented hierarchical cluster selected clusters (the small circles represent the properties
analysis (the red dashed line indicates the cut level) of the pole, while the large circles indicate the mean
values of the poles inside each clusters
178
Chapter 5
5.8.1 INTRODUCTION
This section presents the most relevant results from the experimental identification of the dynamic
properties of a Izar Bonus 1.3MW/62 wind turbine. This analysis is integrated in a project to upgrade
the wind turbine with an extension of the blades, hereafter designated as Rotor Blade Extension (RBE).
This project was led by Energiekontor Windpower Improvement GmbH and was developed in
collaboration with the Institute of Science and Innovation in Mechanical and Industrial Engineering
(INEGI).
In the scope of the RBE project, several tasks were developed in order to assess the feasibility of the
rotor upgrade. Among them, the bending stress of the blades, the evaluation of debris in the gear box
oil and the definition of the wind turbine power curve before and after the rotor extension were
analysed. The Laboratory of Vibrations and Monitoring (ViBest) of FEUP was invited to perform
dynamic tests of the wind turbine blades (Oliveira, Magalhães et al., 2014).
In the context of the project, the assessment of the blades dynamic properties before and after the
extension allows to analyse three important aspects:
o Assessment of the natural frequency of the rotor blades modes – it is of the utmost importance
to guarantee that the frequency values of the extended blades are not close the tower modes
and to the frequency of rotation of the rotor;
o Assessment of the damping ratios from the rotor blades modes – it is important to assess if
there is significant change in the structural damping of the blades after the extension;
o Independent assessment of the natural frequency of each blade – this analysis is intended to
check if the RBE does not change the structural similarity between the blades.
Within the RBE project, data collected by two different measurement systems were used: one system
based on fibre Bragg grating sensors and another based on accelerometers.
The results presented in this section are mainly focused on the assessment of the dynamic properties of
the wind turbine blades before and after the rotor extension. However, it was given the opportunity to
performed an ambient vibration test (before the implementation of the RBE), including simultaneous
measurements at the tower and blades. In this context, a rotor-stop test was also performed in order to
assess the damping ratio of the first and second tower bending modes in the FA direction.
179
Operational Mod
dal Analysis of Wind
W Turbines
Figure
e 5.25 – Photo of the Izar Bon
nus 1.3MW/62
a) b)
Figu
ure 5.26 – Blade
e orientation: aa) operating situ
uation; b) parke
ed situation
180
Chapter 5
Sensor A
Sensor C
Sensor K
Sensor E
Figure 5.27 – Position of the sensors of the measurement system based on fibre Bragg grating sensors
181
Operational Modal Analysis of Wind Turbines
In each setup, a different wind turbine blade was monitored, alongside with the instrumentation
installed at the tower. The test was conducted under low wind speed conditions and the nacelle
presented a misalignment of around 110º to the wind direction. The elements monitored in each setup
are identified in Table 5.4.
Table 5.4 – Measurement setups and corresponding instrumented structural elements during the ambient vibration test
S1 to S4 Tower
1 16
S5 and S6 Blade A
S1 to S4 Tower
2 16
S5 and S6 Blade B
S1 to S4 Tower
3 16
S5 and S6 Blade C
Throughout the test, four accelerometers (S1 to S4), were kept fixed at tower positions at the platform
levels 6, 8 and 10, as schematized in Figure 5.28. Both sensors S3 and S4 were placed at the top
platform in order to identify possible torsion components of the vibration modes.
Lateral view Top view
Level 6 Level 8
62 m
y S1 y S2
S3/S4 Level 10 x z x z
S2 Level 8
68 m Level 10
S1 Level 6
S4 S3
y y
x z x z
+ 0.000
Figure 5.28 – Accelerometers position kept fixed at the tower during the three data sets
In each set, a different wind turbine blade was monitored with two accelerometers (S5 and S6). The
blade to be instrumented was oriented in the horizontal position and the two accelerometers were
placed inside the parked blade in the positions characterized in Figure 5.29 (the mechanical brake was
activated and the blades were pitched with an angle of 97º). The position of the sensors at the blade for
the three setups is presented in Figure 5.30.
182
Chapter 5
a) b))
Figuree 5.29 – Position of the instrum
mented blade d
during the reco
orded setups: a)
a position for tthe blade A and
d C (data sets 1
and 3); b) p
position of the blade B (data set
s 2)
Top view
4.00 m 6.30 m
x S6 x S5
z z
Blade A
y y
5..30 m 4.70
0m
S5 S6
y
x
y
z
Blade B
x z
3.30 m 4.60 m
x S6 x S5
y z y z Blade C
Figure
e 5.30 – Positio
on of the sensors inside the different blades
Figurre 5.31 preseents an illusstration of thhe monitoreed elements of the windd turbine. The blue linee
repreesents the in
nstrumented tower, whille the red lin ne denotes thet portion of the bladee which wass
moniitored. The crosses
c stand
d for the possitions of thee installed seensors. On thhe other han
nd, the blackk
dasheed lines repreesent the parrt of the windd turbine thaat was not insstrumented.
183
3
Operational Modal Analysis of Wind Turbines
a) b)
Figure 5.31 – Instrumented elements of the wind turbine: a) isometric view; b) front view
The performed rotor-stop test was conducted at the end of the day, when the wind speed was around
4.5 m/s. The test consisted in the sudden stop of the rotor, when the wind turbine was operating at
normal conditions, with a rotor speed of 13 rpm. The stop was accomplished through the activation of
the mechanical brake and the change of the blades pitch angle. With this operation, it was possible to
create an impulse at the top of the wind turbine in the FA direction.
In the scope of the rotor-stop test, only the acceleration at the tower was recorded. For that reason,
only the sensors S1 to S4 from Figure 5.28 were considered.
184
Chapter 5
Blade A Blade A
Flapwise Edgewise Flapwise Edgewise Flapwise Edgewise Flapwise Edgewise
25 20
20 15
15
10
10
5 5
με
με
0 0
-5 -5
-10
-10
-15
-20 -15
-25 -20
0 200 400 600 800 1000 988 990 992 994 996 998 1000 1002
Time [s] Time [s]
Blade B Blade B
Flapwise Edgewise Flapwise Edgewise Flapwise Edgewise Flapwise Edgewise
25 15
20
15 10
10
5
5
με
με
0 0
-5
-10 -5
-15
-10
-20
-25 -15
0 200 400 600 800 1000 988 990 992 994 996 998 1000 1002
Time [s] Time [s]
Blade C Blade C
Edgewise Flapwise Edgewise Edgewise Flapwise Edgewise
15
25
20
10
15
10 5
5
με
0
με
0
-5
-5
-10
-15
-10
-20
-25 -15
0 200 400 600 800 1000 988 990 992 994 996 998 1000 1002
Time [s] Time [s]
Figure 5.32 – Strain measured by the fibre Bragg grating sensors: a) complete time series; b) zoom of a period of 14
seconds
As preliminary analysis, the averaged normalized power spectrum density of the time series collected
by each sensor of the three blades was computed. It is presented in Figure 5.33. It is visible how several
resonance peaks can be identified in this figure.
185
Operational Modal Analysis of Wind Turbines
-4
10
-6
10
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Frequency [Hz]
Figure 5.33 – Averaged normalized power spectrum density from each sensor (the dashed lines indicate the main peaks
of the spectra)
The identification of the modal parameters was based on the application of two different parametric
identification algorithms: the SSI-COV (see section 5.4.1.2) and p-LSCF (see section 5.4.2). In this
analysis, the 11 collected signals were processed all together.
The results obtained with the two algorithms were very similar and in accordance with the spectra
from Figure 5.33. The stabilization diagram obtained with the p-LSCF algorithm is presented in Figure
5.34.
Stabilization diagram (p-LSCF)
40
35
30
Model order
25
20
15
10
0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Frequency [Hz]
Figure 5.34 – Stabilization diagram obtained with the p-LSCF algorithm for the data obtained with the measurement
system based on fibre Bragg grating sensors
From the analysis of the stabilization diagrams obtained with both methods, it was possible to estimate
the modal parameters of the 13 vibration modes presented in Table 5.5, including three pairs of closely
spaced modes (around 0.40 Hz, 1.14 Hz and 3.76 Hz). The results obtained with the two algorithms
were very similar. Notwithstanding, the use of the p-LSCF algorithm allowed the identification of three
additional vibration modes.
186
Chapter 5
Table 5.5 – Comparison of results obtained with the SSI-COV and p-LSCF methods
SSI-COV p-LSCF
1 - - 0.441 0.38
2 - - 0.456 0.17
9 - - 2.108 1.00
Since the available layout of sensors only assessed the modal curvature at the blades root, a complete
description of the mode shapes is not possible. Nevertheless, considering that the shape of the first
modes of wind turbine rotors usually follows a typical configuration, an estimation of the first modes is
presented. The modal curvatures at the blades root from each identified vibration mode, alongside
with an illustration of the theoretical shapes of the first 8 modes, are presented in Figure 5.35. Since the
data from channel E from blade C was not available, the symmetric value obtained for the sensor A of
the same blade was used for purposes of illustration (identified with the white bar).
The mode shapes from the first two vibration modes do not resemble any typical rotor mode shape
configuration. These two modes present natural frequency values in accordance with typical first tower
bending modes in the FA and SS directions. Nevertheless, this fact was only confirmed with the
analysis of the results obtained with the ambient vibration test (section 5.8.4.2).
187
Operational Modal Analysis of Wind Turbines
= 0,44 Hz = 0,46 Hz
Mode 1 Mode 2
Blade A Blade B Blade C 1 SS Blade A Blade B Blade C 1 FA
1 Tower Bending 1
Tower Bending
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0 0
-0.2 -0.2
-0.4 -0.4
-0.6 -0.6
-0.8 -0.8
-1 -1
E A C K E A C K E A C K E A C K E A C K E A C K
Flapwise Flapwise Flapwise Flapwise Flapwise Flapwise
Edgewise Edgewise Edgewise Edgewise Edgewise Edgewise
= 0,61 Hz = 1,13 Hz
Mode 3 Mode 4
Blade A Blade B Blade C In plane Blade A Blade B Blade C
1 1 In plane
(symmetric)
0.8 0.8 (vertical)
0.6 0.6
0.4 0.4
0.2 0.2
0 0
-0.2 -0.2
-0.4 -0.4
-0.6 -0.6
-0.8 -0.8
-1 -1
E A C K E A C K E A C K E A C K E A C K E A C K
Flapwise Flapwise Flapwise Flapwise Flapwise Flapwise
Edgewise Edgewise Edgewise Edgewise Edgewise Edgewise
= 1,15 Hz = 1,33 Hz
Mode 5 Mode 6
Blade A Blade B Blade C In plane Blade A Blade B Blade C Out of plane
1 (horizontal) 1
(yaw)
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0 0
-0.2 -0.2
-0.4 -0.4
-0.6 -0.6
-0.8 -0.8
-1 -1
E A C K E A C K E A C K E A C K E A C K E A C K
Flapwise Flapwise Flapwise Flapwise Flapwise Flapwise
Edgewise Edgewise Edgewise Edgewise Edgewise Edgewise
= 1,45 Hz = 2,04 Hz
Mode 7 Mode 8
Blade A Blade B Blade C Out of plane Blade A Blade B Blade C Out of plane
1 1
(tilt) (symmetric)
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0 0
-0.2 -0.2
-0.4 -0.4
-0.6 -0.6
-0.8 -0.8
-1 -1
E A C K E A C K E A C K E A C K E A C K E A C K
Flapwise Flapwise Flapwise Flapwise Flapwise Flapwise
Edgewise Edgewise Edgewise Edgewise Edgewise Edgewise
Figure 5.35 –Amplitude of the modal curvature obtained for the identified vibration modes and respective illustration
(the rotor configuration during the measurement period may not exactly match the illustration)
188
Chapter 5
It is noticed that the illustrations presented for the rotor vibration modes are referred to the wind
turbine under parked condition. Under this condition, the flapwise motion occurs in the rotor plane,
while the edgewise motion is related to motion out of plane of the rotor. This configuration is opposite
to the one when the wind turbine is operating and, for that reason, these illustrations should not be
extrapolated for that condition.
Furthermore, the third identified vibration mode ( = 0.61 Hz) is probably dominated by the drive
train inertia and by the support condition of the rotor (the mechanical brake was activated).
189
Operational Modal Analysis of Wind Turbines
0.4
-2
10
Acceleration [m/s 2]
0.2
Amplitude
-4
0 10
-0.2
-6
10
-0.4
-8
-0.6 10
0 200 400 600 800 1000 0 1 2 3 4 5
Time [s] Frequency [Hz]
0.1
-2
10
Acceleration [m/s 2]
0.05
Amplitude
-4
0 10
-0.05
-6
10
-0.1
-8
-0.15 10
0 200 400 600 800 1000 0 1 2 3 4 5
Time [s] Frequency [Hz]
0.1 -2
10
Acceleration [m/s 2]
0.05
Amplitude
-4
0 10
-0.05
-6
10
-0.1
-8
-0.15 10
0 200 400 600 800 1000 0 1 2 3 4 5
Time [s] Frequency [Hz]
a) b)
Figure 5.36 – Recorded data sets from the accelerometers placed at the blades, during the ambient vibration test: a) time
domain; b) normalized power spectral density
The preliminary analysis of the data sets recorded with the sensors located at the tower was performed
considering two different signals combination for the tower top sensors S3 and S4: half-sum and half-
190
Chapter 5
difference of the horizontal accelerations. With this strategy, it is possible to identify modes with
relevant bending (with the half-sum) and torsion (with the half-difference) components of the tower.
The averaged normalized power spectral density of the acceleration time series in the FA direction
recorded by the sensors placed at the tower top is presented in Figure 5.37 a). On the other hand, the
averaged normalized power spectral density of the accelerations in the SS direction is presented in
Figure 5.37 b). Both spectra clearly evidence one important resonance peak around 0.4 Hz. In addition,
the spectra referred to the SS direction evidence a second important peak around 4.0 Hz. It is also
visible, for the half-difference signal combination in the SS direction, a resonance peak at 3.3 Hz that is
not seen for the half-sum combination. Thus, this peak is probably due to a vibration mode with an
important torsion component.
FA SS
Plat. 6 Plat.10 (Half-sum 6 Plat.10 (Half-sum
0 Plat. 8 Plat.10 (Half-difference) 0 Plat. 8 Plat.10 (Half-difference)
10 10
-2 -2
10 10
Amplitude
-4 Amplitude -4
10 10
-6 -6
10 10
-8 -8
10 10
0 1 2 3 4 5 0 1 2 3 4 5
Frequency [Hz] Frequency [Hz]
a) b)
Figure 5.37 – Averaged normalized power spectral density of the of the acceleration time series recorded by the sensors
placed at the tower in the a) FA and b) SS direction
The identification of the modal parameters was also based on the application of the SSI-COV and p-
LSCF algorithms. In this analysis, each data set was analysed separately and all signals from tower and
blades were considered together. With the purpose of illustrating the obtained results, Figure 5.38
shows the stabilization diagram associated with the acceleration time series collected during setup 2.
Stabilization diagram (p-LSCF)
30
25
20
Model order
15
10
0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Frequency [Hz]
Figure 5.38 – Stabilization diagram obtained with the p-LSCF algorithm for the data collected with the measurement
system based on accelerometers during setup 2
191
Operational Modal Analysis of Wind Turbines
The analysis of the three data sets led to the identification of 16 vibration modes of the wind turbine.
Table 5.6 summarizes the modal properties from these vibration modes identified with the p-LSCF
algorithm. The obtained results present a good agreement for the different analysed data sets. It is
observed that, apart from some modes, the majority of them simultaneously evidence motion of the
rotor blades and tower. From the identified modes, only the 1st, 2nd, 15th and 16th mode presented a
dominant tower motion (without important motion of the blades). For that reason, it was concluded
that they represent the 1st and 2nd pair of tower bending modes.
The obtained results are in good agreement with the ones estimated with the measurement system
based on fibre Bragg grating sensors. However, the modal properties of 3rd vibration mode present an
important deviation between the two measurement systems. This may be due to a worst identification
of this mode with the accelerometers since the resonance peak of this mode is almost not noticeable in
Figure 5.36.
192
Chapter 5
Table 5.6 – Natural frequencies ( ) and modal damping ratios ( ) of the identified vibration modes with the data recorded
by the accelerometers and comparison with the results obtained with the fibre Bragg grating measurement system
(F.B.G.) using the p-LSCF algorithm
3 0.520 10.95 0.530 10.30 0.542 9.41 0.610 8.51 Blade flapwise
Blade flapwise
4 1.128 0.28 1.134 0.54 1.120 0.47 1.13 1.93 Tower SS
Blade flapwise
5 1.140 0.52 1.140 1.18 1.156 1.18 1.146 2.51
Tower SS
6 1.326 0.76 1.324 0.89 1.318 0.64 1.329 0.59 Blade edgewise
Blade flap/edgewise
7 - - 1.474 0.41 1.492 0.61 1.447 0.47 Tower FA
8 1.994 1.56 2.016 0.79 2.020 0.91 2.029 0.54 Blade flap/edgewise
9 2.078 0.81 2.062 1.02 2.062 0.86 2.108 1.00 Blade flapwise
Blade edgewise
10 3.260 0.90 3.330 1.02 3.310 0.86 3.399 1.14
Tower torsion
Blade flapwise
11 3.485 0.33 3.498 0.35 3.490 0.42 3.512 0.55
Tower FA
Blade flap/edgewise
12 3.735 0.29 3.742 0.36 3.738 0.38 3.776 0.42
Tower FA
The mode shapes of some of the identified vibration modes are shown in Figure 5.39, following the
illustration introduced in Figure 5.31 (only the instrumented elements are plotted). Although the
measurement system based on accelerometers allows identifying two additional vibration modes
(when compared to the fibre Bragg grating system), the interpretation of the mode shapes with
important rotor motion presents greater difficulties.
193
Operational Modal Analysis of Wind Turbines
= 0.45 Hz
70 70
30
60 60
20
50 50
10
SS direction
40 40
0
30 30
-10
20 20
-20
10 10
-30
0 0
-20 0 20 -20 0 20 -30 -20 -10 0 10 20 30
FA direction SS direction FA direction
= 0.46 Hz
70 70
30
60 60
20
50 50
10
SS direction
40 40
0
30 30
-10
20 20
-20
10 10
-30
0 0
-20 0 20 -20 0 20 -30 -20 -10 0 10 20 30
FA direction SS direction FA direction
= 0.53 Hz
70 70
30
60 60
20
50 50
10
SS direction
40 40
0
30 30
-10
20 20
-20
10 10
-30
0 0
-20 0 20 -20 0 20 -30 -20 -10 0 10 20 30
FA direction SS direction FA direction
Figure 5.39 – Mode shapes of the first three vibration modes identified with the measurement system based on
accelerometer
194
Chapter 5
Rotor-stop Test
In addition to the ambient vibration test, a rotor-stop test was also performed with the aim of
identifying the damping ratio of the first tower bending mode in the FA direction. The acceleration
time history recorded by the sensors oriented in the FA direction is shown in Figure 5.40. A maximum
value around 0.6 m/s2 was achieved. The motion of the wind turbine after the sudden stop is mainly in
the FA direction; however movement in the orthogonal direction is also noticed (Figure 5.40).
Rotor-stop test Sensor S4
0.8 0.2
S1-FA
0.6
S2-FA
0.4 S3-FA 0.1
SS acceleration [m/s 2]
Acceleration [m/s2]
S4-FA
0.2
0 0
-0.2
-0.4 -0.1
-0.6
-0.8 -0.2
0 50 100 150 200 250 300 350 -0.2 -0.1 0 0.1 0.2
Time [s] FA acceleration [m/s 2]
a) b)
Figure 5.40 – a) Time history from the recorded acceleration by the sensors oriented to the FA direction. b) Movement
seen from above from the accelerations at S4 sensor (top sensor)
The main modes excited with the rotor-stop test are the 2nd vibration mode (1st tower FA mode) and
the 8th vibration mode (the identified rotor symmetric out of plane mode) in Table 5.6. The spectra
obtained with the recorded acceleration time series are presented in Figure 5.41, where the resonance
peaks of the referred modes are evidenced.
-2
10
S1-FA
S2-FA
-4
10 S3-FA
S4-FA
Amplitude
-6
10
-8
10
-10
10
0 1 2 3 4 5
Frequency [Hz]
Figure 5.41 – Power spectral density from the acceleration time series recorded during the rotor-stop test
195
Operational Modal Analysis of Wind Turbines
In order to determine the damping level of the 1st FA tower bending mode, the signals recorded by
sensors were filtered and an exponential function was adjusted to the envelope of the free decay signal.
The used exponential function is defined according to:
. . . .
. (5.92)
For this analysis, a low-pass band filter with a cut of frequency of 1 Hz was used. Considering that the
filtered response does not present a perfect theoretical free decay response, three different amplitude
ranges were considered to adjust an exponential window: the initial part, the last part and the whole
decay. Figure 5.42 a) to c) presents the fittings performed to the acceleration time series recorded by
sensor S4, considering the three signal ranges.
The fitting obtained with the exponential function considering the first part of the signal is clearly
misfit. It is visible from Figure 5.42 a) that the function does not follow the progress of the signal. On
the other hand, the other two fittings seems to properly envelope the free decay signal.
Mode 2 Mode 2
0.3 0.3
S4-FA S4-FA
0.2 0.2
0.260*e-0.0077*2*π*0.453*t 0.201*e -0.0032*2*π*0.453*t
Acceleration [m/s 2]
Acceleration [m/s 2]
0.1 0.1
0 0
-0.1 -0.1
-0.2 -0.2
-0.3 -0.3
0 100 200 300 400 0 100 200 300 400
Time [s] Time [s]
a) b)
Mode 2 Mode 8
0.3 0.15
S4-FA S4-FA
0.2 0.1
0.238*e -0.0038*2*π*0.453*t
0.079*e-0.0066*2*π*2.044*t
Acceleration [m/s 2]
Acceleration [m/s 2]
0.1 0.05
0 0
-0.1 -0.05
-0.2 -0.1
-0.3 -0.15
0 100 200 300 400 0 50 100 150 200
Time [s] Time [s]
c) d)
nd th
Figure 5.42 – Identification of the modal damping ratio of the 2 and 8 vibration mode through filtering and fitting of
an exponential function to the envelope of the free decay response
196
Chapter 5
The same procedure was also applied to the 8th vibration mode. This mode is characterized by a
symmetric in-phase bending of the three blades in the FA direction and, for that reason, was clearly
excited during the test. For this mode, a band-pass filter for the frequency range 1.5 – 2.5 Hz was
considered since the influence of other modes in this range seems negligible. The filtered response
obtained for this mode presents an almost perfect free decay shape and, for that reason, only one
adjustment was performed. Figure 5.42 shows the fitting operation performed for the 8th vibration
mode.
In addition to the referred analysis, a second procedure was also used to estimate the damping value of
the 2nd and 8th modes. This procedure is based on the application of the SSI-COV algorithm
(introduced in section 5.4.1.2). This algorithm is adapted to consider the free decay signals as input of
the method, instead of the correlation functions calculated from the ambient responses (Magalhães,
Cunha et al., 2010). This procedure was applied to the same range of points previously presented for
the two modes.
The results obtained with both procedures are shown in Table 5.7. For the 2nd vibration mode (1st FA
mode), the fitting obtained with the initial part of the decay, associated with higher vibration levels,
provided an higher value, which can be an indication that the damping may increase with the
increasing of the vibration amplitude. The other two fittings performed for this mode represent a good
estimation of the damping value for lower vibration amplitudes with a good agreement between the
two used procedures. This value is considerable higher than the value obtained with the ambient
vibration test, which is consistent with hypothesis of the damping dependency with the vibration
amplitude
The damping value obtained for the 8th mode is coherent with the values obtained with the ambient
vibration test.
It should be noted that damping values associated with operating condition can be considerable
different, and generally higher, than these ones identified with parked conditions.
Table 5.7 – Natural frequencies (݂) and modal damping ratios (ߦ) of the 2nd and 8th vibration modes identified with the
rotor-stop test with the filtering and fitting procedure and with the SSI-COV method
197
Operational Mod
dal Analysis of Wind
W Turbines
5.8.5 ANA
ALYSIS AFTER
R THE INSTAL THE RBE
LLATION OF T
The RBE coonsists on a blade tip thhat is attacheed to the oriiginal bladess in order too extend the rotor
blades lenggth. With thee increase of
o the swept rotor area, this is a sollution to im mprove the energy
e
production on wind farmms. An illusttration of thee RBE is pressented in Figu
ure 5.43.
The RBE in nstallation coonsists on thhe overlappiing of the original rotorr blade tip bby two half shells.
These two shells are thhen fixed witth adhesive resin. With this solution n, the lengthh of each blade is
extended wiith 1.5 m len
ngth tips, inccreasing the sswept area off the rotor in
n 8.6 %. Eachh blade tip weights
w
45 kg, whicch representss an increase of 1 % of the blade mass.
m The maain propertiees of the RB BE are
summarized d in Table 5.88.
For the anaalysis of the dynamic
d chaaracteristics oof the wind turbine
t bladees after the R
RBE, only the data
collected byy the measureement system
m based on fi fibre Bragg grrating system
m was considdered.
Upd
dated diameterr 65 m
198
Chapter 5
Amplitude -4
10
-6
10
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Frequency [Hz]
Figure 5.44 – Averaged normalized power spectrum density from each sensor (the dashed lines indicate the main peaks
of the spectra)
The decrease in the natural frequency values is more evident when the signals collected before and
after the RBE are confronted. Figure 5.45 shows the averaged normalized power spectrum density
functions of the signals collected before and after the extension of the blades. Although, the global
shape of the spectra is kept, there is an evident decrease of the frequency values of the resonance peaks.
0 Flapwise -1 Edgewise
10 10
Before Before
After After
-2
10
-2
10
Amplitude
Amplitude
-3
10
-4
10
-4
10
-6 -5
10 10
0 1 2 3 4 5 0 1 2 3 4 5
Frequency [Hz] Frequency [Hz]
Figure 5.45 – Averaged normalized power spectrum density of the three blades before and after the RBE
The identification of the modal properties of the wind turbine after the RBE was performed with the
same output-only algorithms. Once again, the results obtained with both algorithms were similar. In
order to illustrate these results, the stabilization diagram obtained with the p-LSCF algorithm is
presented in Figure 5.46.
199
Operational Modal Analysis of Wind Turbines
35
30
25
Model Order
20
15
10
0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Frequency [Hz]
Figure 5.46 – Stabilization diagram obtained with the p-LSCF algorithm for the data obtained after the RBE
Several vibrations modes were again identified. However, due to the difficulty in the interpretation of
their mode shapes, only the first 8 modes were considered. The results are presented in Table 5.9. The
correspondence between the modes identified before and after the RBE was done considering the type
of mode (in plane or out of plane) and the proximity between natural frequencies
From the results, it can be seen that the first two modes (tower bending modes) are only slightly
affected by the RBE. However, the accuracy on the identification of these modes is lower than for the
others since the main element of motion (the tower) is not instrumented.
The remaining modes, mainly related with motion of the rotor, present a consistent decrease on the
value of the natural frequencies. It was noted that for modes 6 and 7 there is a change on the mode
configuration (from yaw to tilt and vice versa). This might be motivated by a different position of the
blades in the rotor plane during the two testing periods.
200
Chapter 5
Table 5.9 – Natural frequencies ( ) and modal damping ratios ( ) from the first 8 identified vibration modes before and
after the RBE (p-LSCF algorithm)
6 1.329 0.59 Out of plane (yaw) 1.286 0.75 Out of plane (tilt) -3.2
7 1.449 0.47 Out of plane (tilt) 1.425 0.76 Out of plane (yaw) -1.7
5.8.6 CONCLUSIONS
This section presented the work carried out in the experimental identification of the dynamic
properties of an Izar Bonus 1.3MW/62 wind turbine. In this study, different measurement systems
were used to collect vibration signals from the tower (accelerometers) and blades (fibre Bragg grating
sensors and accelerometers).
The analysis of the data allowed to identify 14 vibration modes, including the first two pairs of tower
bending modes and the 6 first order rotor blades modes. The results obtained with both measurement
systems revealed to be consistent.
After the installation of the Rotor Blade Extension, and the consequent increase of blades mass in 1 %,
the decrease of the rotor modes natural frequencies was clearly identified, showing the ability of the
used tools to identify small changes of the structural elements of the wind turbine.
A rotor-stop test was also performed, enabling the analysis of the modal damping ratio of the 1st tower
bending mode in the FA direction and the 1st symmetric flapwise mode with greater precision. In that
sense, two alternative procedures were used. A good coherence of results was obtained with both
methods.
201
Operational Mod
dal Analysis of Wind
W Turbines
5.9.1 INTTRODUCTION
N
This section
n presents th
he results obtained with the analysis of 80 acceleeration data ssets collected
d by a
dynamic monitoring syystem installeed at a Vestaas V90-3.0MMW offshore wind turbinne. This stud dy was
developed wwith the aim
m of verifying the suitabbility of the developed
d toools when appplied to offfshore
wind turbinnes.
The data ussed in this analysis
a was kindly proviided by the Offshore Wiind Infrastruucture Appliccation
Lab (OWI-LLab - www.o owi-lab.be/).
5.9.2 WIN
ND TURBINE DESCRIPTION
The Vestas V90-3.0MWW wind turbinne is part of the Belwind d wind farm, 46 km off thhe Belgian co
ost, on
the Bligh B
Bank, North Sea (Figure 5.47). This offshore wiind farm is composed bby 55 Vestas V90-
3.0MW turbbines foundeed on monop
piles, leadingg to a total po
ower output capacity
c of 1 65 MW.
• Belwind
B
BLIG H
•B
ZA
ANDBANK
46 km
• Ze
eebrugge
• Brugge
• Oost en
nde
Figure 5.47
5 – Location of the Belwind
d wind farm (Be
elwind Offshore Energy, 20099)
The studied
d Vestas wind d turbine is characterized
c d by a 90 m diameter, up
pwind rotor, with a rated
d rotor
speed of 166.1 rpm. Thee pitch anglee of the threee rotor blad
des is contro
olled by a piitch actuatorr. The
power curvee of the turbiine is introdu
uced in Figurre 5.48.
202
Chapter 5
3000
2600
2200
Power [kW]
1800
1400
1000
600
200
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30
Wind sp
peed [m/s]
72 m
17 m
LAT 25 m
22.9 m
20.6 m
203
3
Operational Modal Analysis of Wind Turbines
Figure 5.50 – Position of the accelerometers at the different levels of the wind turbine (Devriendt, Magalhães et al., 2014)
204
Chapter 5
09/05/2012 03:57
Rotor speed
0.01 0.4
[rpm]
Acc. [g]
0.2
0 0
0 5 10 15 20
-0.01 Index
+69 FA 88.6
Pitch [º]
+69 SS 88.4
0 100 200 300 400 500 600
+41 FA 88.2
Time [s] 0 5 10 15 20
+41 SS Index
-3 +27 FA 168
Yaw [º]
x 10 RMS +27 SS 166
5
+19 FA 164
4 0 5 10 15 20
+19 SS Index
Acc. [g]
Wind speed
3 8
[m/s]
2 6
1 4
0 5 10 15 20
Index
0 5 10 15 20
Data set
Figure 5.51 – Example of measured acceleration time series (top Figure 5.52 – SCADA data of the 20 data sets from
plot) and RMS values of the 20 data sets from Case 1 (bottom plot) Case 1
The remaining 3 cases are referred to production periods under different conditions. The data sets
from Case 2 consist of acceleration time series during periods when the turbine is operating close to
the rated wind speed, with the highest wind speed values from the 4 cases (Figure 5.54). The
acceleration levels from the 20 data sets of this case present also the largest values (Figure 5.53). Under
these conditions, with the rotor orientated to the main wind direction and the blades with low pitch
angles, the turbine support structure vibrates mainly in the FA direction. It is also interesting to note
that the measurement level with the highest vibration level is not the top level (+ 69 m) like in the data
from Case 1, but the + 41 m level. In fact, it is observed that the top level present similar vibration
levels to the +27 m level. This phenomenon is analysed in Chapter 7 in another case study.
01/05/2012 02:16
Rotor speed
0.05
16.04
[rpm]
Acc. [g]
16.02
0
16
0 5 10 15 20
-0.05 Index
+69 FA 20
Pitch [º]
+69 SS 0
0 100 200 300 400 500 600
+41 FA 20
Time [s] 0 5 10 15 20
+41 SS Index
RMS +27 FA 80
Yaw [º]
0.02 +27 SS 60
+19 FA 40
0.015 0 5 10 15 20
+19 SS Index
Acc. [g]
Wind speed
20
0.01
[m/s]
15
0.005 10
0 5 10 15 20
Index
0 5 10 15 20
Data set
Figure 5.53 – Example of measured acceleration time series (top Figure 5.54 – SCADA data of the 20 data sets from
plot) and RMS values of the 20 data sets from case 2 (bottom plot) case 2
Case 3 consists of data sets with varying rotor speeds (Figure 5.56). Similarly to case 2, the turbine also
presents the highest acceleration levels of the support structure at +41 m level (Figure 5.55).
205
Operational Modal Analysis of Wind Turbines
04/05/2012 01:06
0.04
Rotor speed
0.02 16
[rpm]
Acc. [g]
14
0 12
0 5 10 15 20
-0.02 Index
+69 FA -2
Pitch [º]
-0.04
0 100 200 300 400 500 600 +69 SS -2.5
Time [s] +41 FA -3
0 5 10 15 20
+41 SS Index
-3 +27 FA 300
Yaw [º]
x 10 RMS
+27 SS 280
8 +19 FA 260
0 5 10 15 20
+19 SS
Acc. [g]
6 Index
Wind speed
10
[m/s]
4
8
2 6
0 5 10 15 20
0 5 10 15 20 Index
Data set
Figure 5.55 – Example of measured acceleration time series (top Figure 5.56 – SCADA data of the 20 data sets from
plot) and RMS values of the 20 data sets from case 3 (bottom plot) case 3
Lastly, Case 4 comprises the time series under operating conditions with the lowest acceleration levels
(Figure 5.57). This case is referred to data sets in which the turbine is operating with a nearly constant
rotor speed of 10 rpm (Figure 5.58). The only exception is the 6th data set, where a stoppage of the
operation seems to occur.
04/05/2012 09:16
0.02
Rotor speed
15
[rpm]
Acc. [g]
10
0 5
0 5 10 15 20
Index
+69 FA 10
Pitch [º]
-0.02 +69 SS 0
0 100 200 300 400 500 600
+41 FA -10
Time [s] 0 5 10 15 20
+41 SS Index
-3 +27 FA 350
Yaw [º]
Wind speed
3 6
[m/s]
2 4
1 2
0 5 10 15 20
Index
0 5 10 15 20
Data set
Figure 5.57 – Example of measured acceleration time series (top Figure 5.58 – SCADA data of the 20 data sets from
plot) and RMS values of the 20 data sets from case 4 (bottom plot) case 4
The colour map shown in Figure 5.59 illustrates the distribution of energy in the analysed frequency
range (0 – 4.5 Hz) using the first singular value spectra of the 80 acceleration data sets within the
frequency range 0 - 4.5 Hz. The frequency values corresponding to the harmonics 3Ω to 24Ω are also
represented by dashed lines. From an initial analysis of the figure, a vertical alignment with high
energy around 0.36 Hz is immediately noticed throughout the 4 studied cases. This alignment is
probably due to the 1st pair of support structure bending modes, since its value is within the common
values of these modes. It is also interesting to observe the large scattering of energy around the
frequency values below this alignment. This situation is probably due to the wave excitation. As
206
Chapter 5
introduced in Chapter 4, the range of frequencies of wave excitation is usually spread along a relative
wide range of frequencies below the 1st pair of bending modes.
The presented colour map also highlights an important aspect in the analysis of the data from this
offshore wind turbine. The energy scattering of the 4 cases is very different from each other. The first
20 data sets (referred to Case 1), present a clean spread of energy, with three well-defined alignments
around 1.20, 1.44 and 1.50 Hz and three additional not so well-defined alignments around 3.24, 3.60
and 3.90 Hz, besides the aforementioned alignment around 0.36 Hz. However, for the remaining 3
cases (corresponding to operating conditions), only the alignment around 0.36 Hz is clearly identified,
due to the important influence of the rotor harmonics in the spectra. Notwithstanding, a frequency
range of high energy is still seen around 1.50 Hz, although with a considerably higher dispersion than
in Case 1. Alignments of high energy corresponding to the possible modes around 3.24, 3.60 and 3.90
Hz are also hard to identify in Figure 5.59 in these 3 operating cases.
3Ω 6Ω 9Ω 12Ω 15Ω 18Ω 21Ω 24Ω
80
70
60
50
Data set
40
30
20
10
Figure 5.59 – Colour map with the 1st singular value spectra of the 80 data sets (the regions with the hotter colours
represent the highest energy)
Figure 5.60 illustrates the difference in the spectra from operating and non-operating conditions. As
expected, the presence of harmonics is clearly evident. The 3 alignments around 1.20, 1.44 and 1.50 Hz
visible in Figure 5.59 are also easily identified in Figure 5.60 for non-operating conditions. Under
operating conditions, the obtained spectrum is apparently changed in this frequency range. The peak
corresponding to the vibration mode around 1.20 Hz is not present in this frequency. At the same
time, a new (but smaller) peak with a slightly lower frequency value (around 1.17 Hz) is found for the
turbine under operating conditions. This may be a sign that this mode corresponds to a rotor vibration
mode, changing its frequency value according to the pitch angle of the blades. Two very close peaks
around 1.42 Hz and 1.46 Hz are found in the spectra corresponding to operating conditions, while
only the latter is identified in the spectra from non-operating conditions. In addition, the resonance
peak around 1.50 Hz found under non-operating conditions is apparently not identified under
operating conditions due to the interference of the 6Ω harmonic in this frequency range.
The three modes around 3.24, 3.60 and 3.90 Hz, identified in the colour map referred to the Case 1 are
also identified in Figure 5.60 for both situations. However, it is expected that some harmonics may
affect the quality of the identification process of these modes. It is interesting to note that these 3
modes do not present a well identified resonance peak as the other modes do. This is the reason why
the energy alignments of these modes in Figure 5.59 are not as clear as the others.
207
Operational Modal Analysis of Wind Turbines
Lastly, the spectra in Figure 5.60 also allow assessing the influence of the wave excitation in the
dynamic response of the wind turbine support structure. This excitation is spread in small peaks with
frequency values below the first resonance peak (around 0.36 Hz).
0 3Ω 6Ω 9Ω 12Ω 15Ω
10
Non-operating
Operating
-2
10
Amplitude
-4
10
-6
10
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5
Frequency [Hz]
Figure 5.60 – ANPSD from two different setups: under non-operating conditions (rotor speed = 0 rpm; wind speed = 6
m/s) and operating conditions (rotor speed = 16.0 rpm; wind speed = 17 m/s). The vertical dashed lines indicate the
frequency values of the rotor speed 3Ω and its harmonics
50
50
40 40
Model Order
30 30
20 20
10 10
0 0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 0.3 0.35 0.4
Frequency [Hz] Frequency [Hz]
Figure 5.61 – Stabilization diagram obtained with the p-LSCF algorithm in a data set corresponding to non-operating
conditions
208
Chapter 5
Figure 5.62 introduces the stabilization diagram obtained for a data set recorded during operating
conditions with the p-LSCF algorithm. From this figure, it is visible that the rotor harmonics are
identified as alignments of stable poles, just like real vibration modes. Notwithstanding, both the first
pair of bending modes and the modes around 3.24, 3.60 and 3.90 Hz have clear alignments of stable
poles.
The right plot of Figure 5.62 shows a zoom of the stabilization diagram between 1.1 and 2.2 Hz. From
this figure, it is possible to see the existence of stable alignments around 1.17, 1.75 and 2.10 Hz. While
the former is close to a resonance peak identified during non-operating conditions (around 1.20 Hz),
the last two are only identified when the turbine is operating. In addition, two close alignments of
poles are also visible around 1.41 and 1.44 Hz. This situation is different from the one obtained with
the turbine under parked/ idling conditions, where only one vibration mode was identified in this
frequency range. As will be seen ahead, these two modes present a similar mode shape.
Operating conditions Poles Stab. freq Stab. damp Stab. MAC ANPSD
6Ω
3Ω 6Ω 9Ω 12Ω 15Ω
50 50
40 40
Model Order
30 30
20 20
10 10
0 0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 1.2 1.4 1.6 1.8 2 2.2
Frequency [Hz] Frequency [Hz]
Figure 5.62 – Stabilization diagram obtained with the p-LSCF algorithm in a data set corresponding to operating
conditions
Since only a limited number of data was analysed, it is not possible to clearly attest the nature of some
vibration modes, as well as the correspondence of some modes between parked/ idling and operating
conditions. Thus, it is not possible to elaborate a clarifying Campbell diagram, which could help to
better interpret the identified vibration modes. If a proper Campbell diagram could be obtained, it
would be possible to check which modes are related to the blades rotor motion as these modes are
highly dependent on the rotor speed when fixed sensors are used, as explained in Chapter 4.
Notwithstanding, some modes are clearly identified and their nature is indubitably understood. This is
the case of the first pair of support structure bending modes, which natural frequencies are kept fairly
constant during non-operating and operating conditions.
The mode shapes of the 8 vibration modes identified under non-operating conditions are presented in
Figure 5.63. These modes are identified by the index “non” (from non-operating conditions). The two
first modes are indubitably the first pair of bending modes (in FA and SS direction). The 3rd, 4th and 5th
identified modes present a second-order configuration for their mode shapes in the FA, SS and FA
direction, respectively. However, it is noticed that 3rd mode presents practically null displacement at
the top sensor, while the other two show some motion at this level. The last three identified modes also
present a similar configuration, potentially a 3rd order shape configuration. However, since the first
209
Operational Modal Analysis of Wind Turbines
60 60 60
40 40 40
20 20 20
0 0 0
60 60 60
40 40 40
20 20 20
0 0 0
60 60
40 40
20 20
0 0
-20 -20
-1.2 0 1.2 -1.2 0 1.2
Figure 5.63 – Tower mode shapes of the identified vibration modes under non-operating conditions
210
Chapter 5
60 60 60
40 40 40
20 20 20
0 0 0
60 60 60
40 40 40
20 20 20
0 0 0
60 60 60
40 40 40
20 20 20
0 0 0
, . = 3.92 Hz (SS)
80
60
40
20
-20
-1.2 0 1.2
Figure 5.64 – Tower mode shapes of the identified vibration modes under operating conditions
211
Operational Modal Analysis of Wind Turbines
Figure 5.64 introduces the mode shapes of the 10 identified vibration modes during operating
conditions, identified with the index “op.”. The two first modes are the 1st pair of support structure
bending modes, also identified during non-operating conditions. The 3rd, 6th and 7th are modes
vibrating in the SS direction that were not identified with the data corresponding to Case 1. Thus,
these are modes potentially linked to asymmetric edgewise blades rotor motion, since this type of
modes creates a reaction at the top of the tower in the SS direction, due to the configuration of the
blades with low pitch angles (contrary to the configuration in standstill conditions, where the pitch
angle is usually around 80º - 90º). On the other hand, the configurations of 4th and 5th modes (both in
FA direction) resemble the ones from the 3rd and 5th modes under parked/ idling conditions. However,
a mode with a configuration similar to the 4th mode (in the SS direction) was not identified in the
analysed data sets, turning it difficult to assess the nature of these modes. Lastly, the 8th, 9th and 10th
modes show a similar configuration to the 6th, 7th and 8th modes identified under non-operating
conditions. However, it is interesting to note that all of them changed their direction of vibration. The
reasons behind this phenomenon are probably related to the interference of the blades pitch angle in
the motion of these modes. A more profound comprehension of these modes would require the
analysis of a more extended period of data.
After the main vibration modes were preliminary analysed, all the data was automated processed in
order to obtain the modal properties of the introduced modes. After the application of the output-only
identification algorithms to the recorded acceleration time series, the cluster algorithm described in
section 5.7.1 was applied to the obtained stable poles. A maximum distance of 0.02 was defined for the
single linkage criterion. Figure 5.65 shows the computed clusters from the stabilization diagrams
represented in Figures 5.61 and 5.62, quantifying the average frequency value of the poles included in
each cluster and the number of poles of each cluster. In this analysis, it was decided to exclude the
clusters with a number of poles lower than 6, as identified by the red dashed line.
Non-operating conditions Operating conditions
50 50
40 40
30 30
Nº of poles
Nº of poles
20 20
10 10
0 0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5
Frequency [Hz] Frequency [Hz]
Figure 5.65 – Characterization of the clusters: average frequency and number of poles included in each cluster
A representative value of the modal properties from each cluster (frequency value, damping ratio and
mode shape) was kept. Figure 5.66 shows the representative frequency value of the considered clusters
for each data set. In this figure, each selected results was defined as FA or SS, according to the main
direction of vibration of its mode shape. This figure is illustrative of the increased difficulty in the
identification of the vibration modes under operating conditions. While in Case 1 the considered poles
present a very stable and clear evolution of their frequency values, a part from some disturbance
212
Chapter 5
around the 1st pair of bending modes due to the wave excitation, the same does not occur for the
remaining cases. Under operating conditions, the appearance of relevant clusters due to the presence
of the harmonics is evident. Besides the harmonics, this figure allows to confirm some previous
referred considerations. For example, it is evident that the last 3 modes detected for both under
parked/ idling and operating conditions inverted theirs main direction of vibration. Furthermore, both
the 6, . and 7, . modes are clearly distinguished once the turbine is operating (although not in
every case). Lastly, the two closely spaced modes 4, . and 5, . are also identifiable only in cases 2, 3
and 4.
FA
SS
4.5
4 24Ω
3.5 21Ω
3 18Ω
Frequency [Hz]
2.5 15Ω
2 12Ω
1.5 9Ω
1 6Ω
0.5 3Ω
0
0 10 20 30 40 50 60 70 80
Data set
Figure 5.66 – Considered poles from each data set according to the main direction of vibration of the mode shape (the
dashed lines refer to the frequency value of the harmonics of each data set)
In order to identify the selected reference vibration modes from the poles obtained with the
identification algorithms, the modal properties presented in Figures 7.63 and 7.64 were used as
reference properties of the modes. Thus, for each data set, the clusters with a frequency value close to a
reference natural frequency value were selected. The frequency range adopted to consider a pole varied
between 10 % (for the 1st pair of bending modes) and 20 % (for the others). Within the considered
poles, the one with highest correlation with the reference mode shape (i.e. the highest value of MAC
coefficient) is selected. Only modes with MAC values higher than 0.80 were considered.
This procedure was applied together with the p-LSCF and SSI-COV modal identification algorithms,
as already stated. For the p-LSCF method, positive time lags of the correlation with 1024 points were
used to calculate the spectra, together with an exponential window with a factor of 0.1. A maximum
model order of 50 was considered. In order to assess the stability of a pole, the criteria exhibited in
Table 5.10 were followed. These settings were already used for the results previously presented in this
section.
213
Operational Modal Analysis of Wind Turbines
Frequency Δf ≤ 1 %
Damping Δξ ≤ 5 %
Figure 5.67 presents the results obtained after the comparison of the selected cluster properties with
the properties of the reference modes. From this figure, it can be seen that the modes were successfully
identified throughout all the cases. Only the identification of the 3, . mode shows a poor success
rate, being only identified in case 2. In this figure, the colours selected for the tracked modes from Case
1 purposely does not match with the colours from the other Cases since it was not intended to link the
identified vibration modes from non-operating conditions with the ones from operating conditions.
Case 1 Case 2 Case 3 Case 4 Case 1 Case 2, 3, 4
4.5
1,non (FA) 1,op. (FA)
4 2,non (SS) 2,op. (SS)
3,non (FA) 3,op. (SS)
3.5
4,non (SS) 4,op. (FA)
3 5,non (FA) 5,op. (FA)
Frequency [Hz]
0.5
0
0 10 20 30 40 50 60 70 80
Data set
The identified modal properties of the vibration modes with the p-LSCF algorithm throughout the 4
cases are summarized in Table 5.11. From these results, it is interesting to note that, under non-
operating conditions, the 2nd mode (the 1st bending mode in the SS direction) presents higher damping
values than its pair in the FA direction. However, under operating conditions, this situation is
inverted, being the mode in the FA direction the one with the highest damping values. This situation is
caused by a possible deviation of the rotor from the wind direction and to the high pitch value of the
blades when the turbine is parked/ idling (Case 1). Under this configuration, the blades present a
larger opposition to the air in the SS direction, leading to the appearance of some contribution of
aerodynamic damping in this direction. When the turbine is operating (Case 2, 3 and 4), the rotor is
oriented to the main wind direction and the blades present a low pitch angle. Under these
circumstances, the main opposition to the air by the rotor is presented in the FA direction, vanishing
the aerodynamic contribution in the SS direction. Since the aerodynamic damping component
increases with the increase of the wind speed (assuming similar blades pitch angle), the damping
values of the mode in the FA direction shows its highest values in Case 2, corresponding to the data
sets with the highest wind speed.
214
Chapter 5
Another interesting aspect of the results obtained is related with the damping values of the last 3 mode
shapes of both parked/ idling and operating conditions. It is noted that for non-operating conditions,
the SS vibration mode (7th mode) presents the highest damping values, while the other 2 modes
vibrating in the FA direction present similar damping values. When the turbine is operating, the two
SS modes present similar damping values (8th and 10th modes), while the 9th mode (in the FA direction)
shows the highest damping ratio. This correspondence reinforces the idea of these modes found in
Case 2, 3 and 4 may be linked to the last three modes found in 1 case.
Table 5.11 – Mean values of natural frequency and damping ratio obtained with the p-LSCF algorithm
The same strategy was also followed with the SSI-COV algorithm. For this method, 512 points of the
correlation functions were used. The state-space models were defined with a maximum order of 100.
Table 5.12 sums up the main results obtained with this algorithm. Although similar, the results
attained with this algorithm were not as good as the ones obtained with the p-LSCF, for the analysed
data. This is particularly relevant due to non-identification of 3rd and 7th vibration mode during
operating conditions.
215
Operational Modal Analysis of Wind Turbines
Table 5.12 – Mean values of natural frequency and damping ratio obtained with the SSI-COV algorithm
5.9.6 CONCLUSIONS
This section presented the results obtained with the analysis of vibration data collected during a short
period of time from a Vestas V90-3.0MW offshore wind turbine. In this analysis, four different
scenarios of operation were studied, including idling and power production conditions.
The analysis undertaken permitted to identify 8 vibration modes during non-operating conditions and
10 modes during power production conditions. It was noticed that the dynamic properties of the wind
turbine are considerably different depending if the system is operating or not.
An automated procedure was successfully applied to the data to automatically track the modal
properties of the identified vibration modes throughout the different cases. The results obtained gave a
good indication about the possibility of development of a generic dynamic monitoring system to be
installed in both onshore and offshore, as exposed in (Oliveira, Weijtjens et al., 2014).
Due to the restricted period of analysis, it was not possible to attest the accuracy of the monitoring
system to detect small damages on the support structure of the wind turbine. Notwithstanding, work
developed on the same wind turbine using a long period of data revealed a good accuracy in the
detection of structural changes in the foundation (Weijtjens, Verbelen et al., 2015).
216
Chapter 6
6
FATIGUE ASSESSMENT OF WIND
TURBINES
217
Fatigue Assessment of Wind Turbines
15 0
17 -2
11 -4
100 200 300 400 500 600 700
Time [s]
Figure 6.1 – Mean wind speed and turbulent component of the wind speed at the hub height (simulated in the context of
the NREL 5MW wind turbine example)
The turbulent component of the wind can thus be seen as the component responsible for the dynamic
excitation of the wind turbine. It is considered a non-deterministic variable and, for that reason, it is
usually defined by statistical means. A common way to define the turbulence intensity of the wind
in the flow direction is through the ratio between the standard-deviation of the wind speed and its
mean value:
(6.1)
In addition to the turbulence in the mean wind direction, also the lateral and vertical turbulence are
defined similarly.
The turbulence is usually represented in the spectral form. This representation allows understanding
the frequency regions with higher excitation energy. The most commonly used turbulence spectra
218
Chapter 6
models are the von Kármán and the Kaimal spectra (Burton, Sharpe et al., 2001). Figure 6.2 presents
the von Kármán turbulence autospectral density function for several mean wind speeds. As can be
seen, the energy is essentially concentrated at the very low frequency values.
For common wind turbine structures, whose lowest vibration mode is usually around 0.30 – 0.40 Hz,
the energy from the lowest part of the spectrum is mainly responsible for the quasi-static response of
the structure, i. e., with high amplitude displacement cycles with large periods of occurrence. On the
other hand, the remaining frequency range of the spectrum is responsible for the resonance driven
motion of the wind turbine. This means that, in this frequency range, the motion of the wind turbine is
commanded by resonance of the main support structure vibration modes.
The mean component of the wind may be also important for fatigue analysis of some structural details.
For example, the assessment of fatigue damage of some materials (such as composite or concrete) or
elements (such as prestressed bolted connections) is dependent on the mean value of the stress cycles.
Notwithstanding, other elements (such as steel towers) do not require any consideration about the
mean stress level for fatigue analysis according to the main standards/ guidelines (section 6.3.2).
2 Turbulence autospectral density function (longitudinal)
10
0
10
Amplitude
28 m/s
24 m/s
20 m/s
-2 16 m/s
10
12 m/s
8 m/s
-4 4 m/s
10
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Frequency [Hz]
Figure 6.2 – Von Kármán longitudinal turbulent spectra for various mean wind speeds (I = 10 %)
The rotor operation is another source of excitation of wind turbines. As stated in section 4.6, the
rotation of the rotor introduces cyclic loads at integer multiples of the rotor speed. For the wind
turbine support structure, the energy of this excitation will be mainly located at the frequency of 3Ω
and its integer multiples. For a common variable-speed wind turbine (with a rotor operating regime
between 8 and 20 rpm), the 3Ω harmonic excitation presents a frequency between 0.4 Hz and 1.0 Hz.
Figure 6.3 illustrates the change introduced in the aerodynamic rotor thrust by the rotor rotation using
data from the NREL 5MW wind turbine example. In this figure, the (theoretical) aerodynamic force
that would be obtained if the harmonic effect is vanished (“aerodynamic force”) is compared with the
actual aerodynamic rotor thrust calculated by the HAWC2 code (“aerodynamic rotor thrust”). These
forces represent the load transmitted by the rotor blades to the support structure of the wind turbine.
The effect introduced by the rotor rotation is clearly visible, with well-defined peaks at the harmonic
frequencies. It is thus concluded that, during operating periods, the wind load really applied on the
support structure does not follow a turbulence spectrum like the one from Figure 6.2 but one similar to
the presented in Figure 6.3.
219
Fatigue Assessment of Wind Turbines
0
10
Aero. force
-2 Aero. rotor thrust
10
Normalized amplitude
-4
10
-6
10
-8
10
-10
10
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Frequency [Hz]
Figure 6.3 – Averaged normalized spectrum of the aerodynamic force and aerodynamic rotor thrust
At offshore conditions, wave excitation is another important source of dynamic loading of the wind
turbine support structure. Waves are mainly originated as result of the wind action along the surface of
the sea. Thus, just like the wind, waves also present an apparent random behaviour (wave loading is
considered a non-deterministic process too).
The formation of waves can be seen as a combination of several regular waves, defining the sea surface.
It is thus helpful to define the energy distribution in the frequency domain. In that sense, two
important statistical properties should be referred: the significant wave height ( ) and the peak
period ( ). The significant wave height is related to the magnitude of the waves. It is approximately
four times the standard deviation of the sea surface height (Manwell, McGowan et al., 2010). The peak
period is the wave period associated with the most energetic waves (which can be also defined in terms
of frequency - ).
The wave spectrum is an essential tool to assess the loading environment to which the wind turbine
will be subjected. The two most commonly used wave spectra are the Pierson-Moskowitz and the
JONSWAP spectra.
The shape of the Pierson-Moskowitz wave spectrum was fitted to measurements from the Atlantic
Ocean during long periods of constant environment conditions (Pierson and Moskowitz, 1964). This
spectrum only considers the wind speed as input. Later, this spectrum was updated to consider and
as input.
The JONSWAP spectrum corresponds to a modified version of the Pierson-Moskowitz spectrum
(Hasselmann, Barnett et al., 1973). It is based on measurements performed in the North Sea. The
JONSWAP spectrum is characterized by a higher and narrower peak (when compared to the Pierson-
Moskowitz spectrum). For that reason, the JONSWAP spectrum is often used for extreme event
analysis (Manwell, McGowan et al., 2010).
Figure 6.4 presents the 22 representative sea state conditions for the 20 m water depth Offshore Wind
farm Egmond aan Zee (OWEZ), located in the Dutch North Sea (Tempel, 2006). For this figure, the
Pierson-Moskowitz wave spectrum was used. It is seen that the highest energy is related to low
frequency values, below the common values of natural frequency of the 1st support structure bending
mode. Notwithstanding, it is also visible that, although the peak may present a wide spread of energy
220
Chapter 6
for low values of wave height, the energy is almost zero for very low frequencies. This is an indication
that the wave loading may not induce a quasi-static response of the wind turbine.
14
12 P=0.39 %
8
P=0.15 %
6 P=2.34 %
4 P=5.28 %
P=1.79 %
2 P=3.26 %
P=10.48 %
0
0 0.1 0.2 0.3 0.4 0.5 0.6
Frequency [Hz]
Figure 6.4 – Wave spectra from the OWEZ site (the probability of occurrence of the first 7 states with higher wave height
is shown)
221
Fatigue Assessment of Wind Turbines
a) b) c)
Figure 6.5 – a) Bolted connection between flanges. b) Welded connection between cylindrical sheet and the flange. c)
Door opening at the bottom of the tower
The connection between the tower and the foundation is also another sensitive location for fatigue
(Currie, Saafi et al., 2015). Considering onshore wind turbines, the constant vibration of the structural
system can lead to cracking of the concrete around the ring imbibed in the foundation (see Figure 6.6).
The problem of fatigue is even more severe in offshore wind turbines. The grouted connection of the
transition piece is a very delicate location for fatigue and is subjected to very high stresses (see Figure
3.9). In addition, also some offshore foundation models (such as the tripod and jacket foundations)
need an accurate analysis of fatigue efforts, due to their complicate geometry and high number of
joints.
Figure 6.6 – Detail of the connection between the insert ring and the slab foundation
222
Chapter 6
2
10
Threshold defined in
EC3
1
10
4 5 6 7 8 9
10 10 10 10 10 10
Number of cycles (N)
In addition, corrections to the S-N curves are proposed in some standards/ guidelines, considering
thickness, corrosion and material effects, among others.
The consideration of the mean stress in fatigue analysis is also referred in the standards/ guidelines,
since different structural details required different approaches:
o The fatigue analysis of concrete elements, such foundations and grouted joints, requires the
knowledge of the mean stress;
o The study of fatigue state of the pre-loaded bolted joints also requires the consideration of the
mean stress level. This is due to the non-linear relation between the stress state of the tower
and the forces in the bolts (Sørensen and Sørensen, 2011);
o For the analysis of steel elements from the support structure, the mean value of the stress
cycles is usually disregarded. Some standards/ guidelines present a correction coefficient to
consider the mean stress effect. However, this consideration only introduces a positive effect
when a portion (or the whole range) of the stress range is in the compression side. Thus, in
this context, the disregarding of this coefficient (and consequently, of the mean stress effect) is
conservative.
These documents also define partial safety factors to account for the uncertainties about material,
loads and consequences of failure (although some documents do not define each value separately).
223
Fatigue Assessment of Wind Turbines
Both IEC, GL and EC 3 consider the safety factors that increase the loads. On the other hand, DNV
defines a safety factor (named Design Fatigue Factor - DFF) which is considered as a magnification
factor of the accumulated damage . As example, the safety factors defined in the referred documents
for the steel elements of an onshore tower are summarized in Table 6.1.
Standard/ Consequences of
Load Material Total
Guideline failure
DNV* - - - 2.00
224
Chapter 6
(6.2)
with:
Initial velocity
Initial displacement
225
Fatigue Assessment of Wind Turbines
1
y0 = 0.2
.
y0 = 0.06
0.5
Displacement
0
y0 = 0
.
y0 = 0
-0.5
-1
0 2 4 6 8 10 12 14
Time
Figure 6.8 – Results from double integration of acceleration functions with different initial conditions
Notwithstanding, there are some methodologies published regarding the estimation of the stress time
history from accelerometers. However, they are focused on the dynamic component of the
displacements, i. e., the quasi-static displacements, which depend on the initial conditions and on the
static loads, are disregarded.
In (Hjelm, Brincker et al., 2005) and (Aenlle, Skafte et al., 2013), an interesting methodology is
presented. This procedure uses the measured accelerations to estimate the dynamic properties of the
structure. Alongside, the measured accelerations are integrated twice to obtain an estimation of the
displacements. Using the estimated displacements and the experimentally identified vibration modes,
the modal coordinates can be estimated according to:
, . , (6.3)
with:
Estimated displacements
, Modal coordinates
Equation (6.3) can be solved directly if the number of mode shapes equals the measured degrees of
freedom. If the measured degrees of freedom exceed the number of vibration modes considered, the
equation can be solved with the least square method.
226
Chapter 6
Once the dynamic characteristics are known, a finite-element model of the structure is developed,
which is calibrated and updated based on the experimental results. With the numerically obtained
mode shapes, it is possible to estimate the displacements at any point of the structure through:
, . , (6.4)
with:
Considering that the relationship between bending moment and curvature at a given section of the
structure is defined by:
. . (6.5)
with:
(6.6)
with:
Bending stress
Bending moment
227
Fatigue Assessment of Wind Turbines
the bending stress at any point along the height of the structure ( axis) is then calculated according to
the combination of equations (6.5) and (6.6):
,
, . (6.7)
with:
This formulation implies that the function describing each vibration mode has to be derived twice (in
order to the height) to obtain the stress at any point of the structure. An example of application of the
methodology to a reduced two story building model is presented in (Aenlle, Hermanns et al., 2013).
The described procedure is developed with the support of a finite-element model. However, its
application to wind turbines presents some important drawbacks. Firstly, the need of a properly
calibrated finite-element model of a wind turbine can be an obstacle, since geometrical and material
characteristics of the nacelle and rotor are usually not provided by the manufacturer. In addition, the
direct use of the experimentally obtained mode shapes in equation (6.7), avoiding the need of a finite-
element model, may lead to important errors in the double derivative operation. Lastly, the quasi-static
component of the displacements, mostly due to wind action, is not considered.
In order to overcome these obstacles, a procedure to assess the stress condition at any location of a
wind turbine support structure is presented in next section.
228
Chapter 6
Once the modal acceleration responses are computed, the modal displacements can be estimated
through a double integration process. In this work, the acceleration time series are transformed into
the frequency domain and thereafter double integrated. The modal displacement of the structure
motivated by the -th vibration mode at one of the sensors position is thus defined as:
(6.8)
with:
The obtained displacements are then transformed back to the time domain.
After the displacement field for each vibration mode is achieved, it is possible to compute equivalent
forces imposing the same (modal) deformation. The equivalent forces acting along the height of the
support structure at the time step from the -th vibration mode are obtained according to:
. (6.9)
with:
Time step
Equation (6.9) requires the definition of the stiffness matrix of the wind turbine support structure.
Contrary to the nacelle and rotor, the support structure can be defined through very simple models.
The tower structure can be assumed as a simple cantilevered beam, once the material and structural
characteristics (variation of the bending stiffness along the height) are known.
The total equivalent forces acting along the height of the support structure at the time step are
obtained as a sum of the contribution from the identified vibration modes:
(6.10)
Equation (6.9) defines equivalent loads at the measurements points (where the modal amplitudes of
the mode shapes are known) imposing the same modal deformation of mode . Thus, if the mode
shape is defined by a limited number of points (as is usual the case, since the number of accelerometers
placed along the structure is reduced), equation (6.9) will conduct to large concentrated loads at few
229
Fatigue Assessment of Wind Turbines
points of the structure. In this case, the stress condition obtained for the support structure will not
correspond to reality.
In order to overcome this problem, an interpolation of the experimentally found mode shapes is used
(e.g. spline interpolation) to obtain a more detailed estimation of the modal amplitudes of the mode
along the height of the support structure. As referred in section 5.5, the information about the mode
shapes is contained in the matrix from the forward innovation model (see equation (5.71)). When
the mode shapes are interpolated, a new matrix is defined, named , which can be used to
estimate the accelerations at the interpolation points:
. , (6.11)
Once the acceleration time series at the interpolation points are obtained, the presented procedure can
be used to estimate equivalent loads at these points. Considering that a fine resolution is used in the
interpolating operation, the loads will resemble a distributed load, leading to a stress field in
accordance with the reality.
The interpolating operation has the advantage of avoiding the need of a well calibrated finite-element
model. However, the suitability of the interpolation must be carefully analysed. In situations where the
number of sensors is limited and vibration modes with high order modal configurations are important
for the displacement response, this operation may introduce important errors. In the case of support
structures of wind turbines, the displacement response is mainly conditioned by the first bending
mode and low order harmonics. These mode shapes (and operating deflection shapes) present smooth
configurations. Thus, a large number of sensors is not required for their correct definition.
Figure 6.9 illustrates the application of the presented methodology to a wind turbine support structure,
using three levels of measurement. In this figure, the red circles represent the identified modal
amplitudes from the considered mode shapes, which are used to obtain the interpolation presented by
the thick line. The stress at any point of the structure is then obtained as a sum of the stress
contribution from each vibration mode. However, it is important to refer that the equivalent forces
computed with equation (6.9) should not be confused as real loads acting on the structure, since they
are just a gimmick used to recreate the deformed structure.
230
Chapter 6
= + +
Figure 6.9 – Illustration of the proposed methodology to estimate the equivalent forces along the tower height
One critical step when using accelerometers for fatigue assessment of structures is the double
integration operation. From equation (6.8), it is noticed that the result is prone to errors for frequency
values close to zero. Figure 6.10 illustrates a typical result obtained with the double integration of
acceleration time series from wind turbines support structures. It presents the estimation of the
displacement at the top position (computed according to equation (6.8)) from a simulated acceleration
(sensor S3) of the HAWC2 example. The displacement directly obtained with the numerical analysis,
from which the mean value was subtracted, is also presented. The comparison between the two signals
evidences a clear distortion of the estimated signal in the time domain. The estimated displacement is
governed by an erroneous large period cycle, masking the real displacement time series. This illustrates
the difficulty in estimating displacements from accelerations.
Sensor S2 - Displacement 5 Sensor S2 - Displacement
2 10
Original
1.5
Estimated
0
1 10
Displacement [m]
0.5
Amplitude
-5
0 10 0 0.15 0.3
-0.5
-10
-1 10
Original
-1.5
Estimated
-15
-2 10
100 200 300 400 500 600 700 0 1 2 3 4 5
Time [s] Frequency [Hz]
Figure 6.10 – Time and frequency representation of the displacement (from which the mean value was subtracted)
obtained with the HAWC2 model (original) and estimated displacement computed through the double integration
process without any filtering window
231
Fatigue Assessment of Wind Turbines
It is thus comprehensible that this procedure is not capable of estimating the stress associated with the
very low frequency range of the spectrum. This part of the response spectrum, usually related to the
quasi-static component of the wind loading, will be referred in this work as “quasi-static” component
of the stress.
On the other hand, the described procedure is adequate to estimate the stress condition associated
with the remaining part of the spectrum, which includes the contribution of the vibration modes and
sinusoidal excitations (such as waves and rotor rotation driven harmonics). This component will be
referred as the “dynamic” component of the stress.
A complementary approach is thus required to estimate the quasi-static component of the stress. This
approach takes advantage of the fact that most of the wind action is applied on the support structure as
a thrust force at the hub level. Thus, it can be assumed that the quasi-static deformation of the
structure is similar to a cantilevered beam with a concentrated load at the top (Figure 6.11). The
stiffness matrix defined for the estimation of the dynamic component of the stress can then be used to
estimate this deformation.
Once this deformation shape is known, the knowledge of the quasi-static condition of a reference
point along the support structure is enough to estimate the stress condition along the entire structure.
This reference quasi-static condition can be the stress (measured with strain gauges), the displacement
(measured with GPS antenna) or the curvature (measured with inclinometers) evolution along time at
an adequate section of the support structure.
The use of a single measurement point to estimate the condition of the whole support structure is
illustrated in Figure 6.11. Assuming the deformed shape of the structure and a known reference
condition, the bending moment from the quasi-static component (and, consequently, the bending
stress) can be extrapolated for the rest of the structure.
Figure 6.11 – Thrust force applied to the wind turbine and illustration of the equivalent deformation and bending
moment of the support structure due to the quasi-static component of the loading
232
Chapter 6
1
Low-pass
0.8 High-pass
Anplitude
0.6
0.4
0.2
0
0 fc1 fc2
Frequency
The filter parameters ( and ) should be adjusted according to each situation. The main
constraints to define the parameters are the lowest natural frequency of the wind turbine, the first
detected harmonic excitation, the wave loading spectrum (if the turbine is located offshore) and the
eventual frequency limits for the linear behaviour of the accelerometers. The filter should then be
defined in order to consider these effects in the dynamic component of the displacements.
233
Fatigue Assessment of Wind Turbines
The main steps of the proposed procedure for fatigue assessment of the wind turbine support structure
are summarized in Figure 6.13. As previously described, the procedure starts with the computation of
the low-pass filtered quasi-static signal from the quasi-static input data (stress, displacement or
curvature). Similarly, the dynamic acceleration data is obtained with the application of the referred
high-pass filter to the collected acceleration signals. From this point, two independent procedures are
applied in order to compute the two contributions to fatigue.
The quasi-static contribution to fatigue is obtained assuming the applied forces acting solely on the
rotor blades (and then transmitted to the tower through the hub). Since the deformed shape of a
cantilever beam with a concentrated force at the top is known, the stress condition along the support
structure can be estimated based on the quasi-static condition of a reference point (installed sensor).
The procedure to estimate the dynamic contribution to fatigue requires a more complex methodology.
It starts with the decomposition of the “dynamic” acceleration data into modal (and harmonic, if the
turbine is operating) responses. These acceleration time series are integrated to obtain modal (or
harmonic) displacement responses. Then, the modal (or harmonic) displacement responses are
estimated at non measured locations through an interpolation of the experimentally found mode
shapes. Lastly, equivalent forces are obtain, for each vibration mode (or harmonic), imposing the same
(modal) deformation.
Once the quasi-static and dynamic contributions to stress are estimated, the fatigue condition at an
arbitrary position of the wind turbine support structure can be obtained by summing these two
components.
234
Chapter 6
Input Data
Quasi-static Dynamic
contribution to fatigue contribution to fatigue
Known deformed
Acceleration
shape Time
Acceleration
modal response
Acceleration
Time
Acceleration
Time
Double
Displacement
integration
Time
Time
Vibraon
Mode 1
Mode 1 Harmonics
Modes
Displacement
Time
Interpolation of the
displacements along
Displacement
p
Time
tower height
Displacement
p
Time
Displacement
p
Time
Displacement
Time
Mode 1
Vibraon
p
Time
Mode 1 Harmonics
Displacement
Modes
p
Time
Displacement
Computation of
equivalent forces
Equivalent forces
Mode 1
Vibraon
Mode 1 Harmonics
Modes
Figure 6.13 – Chain of events included in the proposed procedure for fatigue assessment at any position of the wind
turbine support structure
235
Fatigue Assessment of Wind Turbines
Example
In order to illustrate the proposed methodology, an additional sensor was introduced in the equipment
layout shown in Figure 4.21. This sensor is required to obtain the reference quasi-static condition. It is
assumed that the displacement at the position of sensor S1 is known.
For the analysis, the filter parameters ݂ଵ and ݂ଶ were defined as 0.10 Hz and 0.15 Hz, respectively. The
measured displacement and the signal obtained after the application of the low-pass filter are presented in
Figure 6.14.
0.04
Measured disp.
Filtered disp.
0.035
Displacement [m]
0.03
0.025
0 100 200 300 400 500 600
Time [s]
Figure 6.14 – Measured displacement and low-pass filtered displacement time series at S1 sensor position
Once the displacement at the S1 position is known, it is possible to estimate the quasi-static condition at
any position along the support structure. Figure 6.15 shows the comparison between the real (“original” -
which was filtered with the same low-pass window), and estimated quasi-static displacements at the S2
and S3 sensors position. As expected, the agreement between the signals is very good.
Sensor S2 - Low-pass filtered displacements Sensor S3 - Low-pass filtered displacements
0.16 0.32
Original Original
0.15 Estimated Estimated
0.3
Displacement [m]
Displacement [m]
0.14
0.28
0.13
0.26
0.12
0.24
0.11
0.1 0.22
0 100 200 300 400 500 600 0 100 200 300 400 500 600
Time [s] Time [s]
Figure 6.15 – Original and estimated low-pass filtered displacements at S2 and S3 sensors position
Together with the stiffness matrix, the low-pass filtered signal of the measured displacements at S1
position is used to estimate quasi-static component of the bending moment at the foundation level. The
obtained results are shown in Figure 6.16. An almost perfect correlation is observed between the
estimated and the original quasi-static base bending moment.
236
Chapter 6
4 4
x 10 Base bending moment x 10 Low-pass filtered base bending moment
5 5
Original
4.8 Original
Estimated (filt). 4.8
Estimated
4.6 4.6
Bending moment [kN.m]
3.4 3.4
3.2 3.2
0 100 200 300 400 500 600 0 100 200 300 400 500 600
Time [s] Time [s]
a) b)
Figure 6.16 – Comparison of the estimated quasi-static component of the base bending moment with: a) original
bending moment; b) low-pass filtered bending moment
After the estimation of the quasi-static component, it is necessary to compute the dynamic component of
the bending moment. In this example, the SSI-COV modal identification algorithm was used. Using the
modal acceleration responses estimated with the algorithm presented in section 5.5, the dynamic
component of the displacements at the sensors position was computed. The results were then compared
to the displacements obtained with the HAWC2 simulation after application of the high-pass filter. Figure
6.17 shows the very good agreement between both results.
237
Fatigue Assessment of Wind Turbines
-3 -3
x 10 S1 - High-pass filtered displacements x 10 S1 - High-pass filtered displacements
4 4
Original Original
3 3
Estimated Estimated
2 2
Displacements [m]
Displacements [m]
1 1
0 0
-1 -1
-2 -2
-3 -3
-4 -4
0 100 200 300 400 500 600 100 110 120 130 140 150
Time [s] Time [s]
Displacements [m]
0.005 0.005
0 0
-0.005 -0.005
-0.01 -0.01
-0.015 -0.015
0 100 200 300 400 500 600 100 110 120 130 140 150
Time [s] Time [s]
Displacements [m]
0.01 0.01
0 0
-0.01 -0.01
-0.02 -0.02
-0.03 -0.03
0 100 200 300 400 500 600 100 110 120 130 140 150
Time [s] Time [s]
Figure 6.17 – Original and estimated high-pass filtered displacements at the sensors position (the right hand plots are a
zoom of the left hand figures)
During the integration process in the frequency domain, numerical errors are introduced in the
displacement signal due to the application of the filter window over the acceleration time series. In order
to minimize these errors, a window is applied to the acceleration signal (in time domain) to attenuate the
amplitude of the signal in the beginning and end, prior to the integration. Figure 6.18 compares the
dynamic displacement obtained without and with the application of the attenuate window. It is visible
that, if the window is not used (Figure 6.18 a)), an error is present in the signal. Although the use of the
window may distort even further the initial and last part of the signal, it allows to confidently control the
length of these erroneous segments. With this strategy, it is thus possible to disregard these two time
segments when computing the bending stress for fatigue analysis purposes.
238
Chapter 6
-3 -3
x 10 S1 - High-pass filtered displacements x 10 S1 - High-pass filtered displacements
4 4
Original Original
3 3
Estimated Estimated
2 2
Displacements [m]
Displacements [m]
1 1
0 0
-1 -1
-2 -2
-3 -3
-4 -4
0 5 10 15 20 25 30 0 5 10 15 20 25 30
Time [s] Time [s]
a) b)
Figure 6.18 – Estimated dynamic displacements at sensor S1: a) without using the attenuate window; b) using the
attenuate window
The estimated dynamic displacements are then used to compute the dynamic component of the bending
moment at any position of the support structure. In this example, the foundation level was chosen.
Through the procedure introduced in section 6.4.1, the dynamic component of the bending moment at
the foundation level was estimated. The result is presented in Figure 6.19 alongside with the high-pass
filtered bending moment obtained as results of the numerical analysis. In this figure, the effect of the
applied attenuating window is seen in the beginning and end of the estimated time series. These two
segments should then be disregarded.
High-pass filtered base bending moment
6000
Original
4000 Estimated
Bending moment [kN.m]
2000
-2000
-4000
-6000
0 100 200 300 400 500 600
Time [s]
Figure 6.19 – Comparison of the estimated dynamic component of the base bending moment with the original high-pass
filtered bending moment
With both quasi-static and dynamic components estimated, the total bending moment can be obtained
with a sum between the two components. The total bending stress at the foundation level is shown in
Figure 6.20, where the initial and last transient segments were already eliminated.
239
Fatigue Assessment of Wind Turbines
600
Nº occurrence
60
500
55
400
50 300
200
45
100
40 0
100 200 300 400 500 0 5 10 15 20
Base bending moment [kN.m] Bending stress [MPa]
Figure 6.20 – Comparison between the original and Figure 6.21 –Histogram with the number of cycles from
estimated base bending stress the original and estimated bending stress at the
foundation level
At this point, the histogram of the loading cycles can be computed. This histogram is compared with the
one corresponding to the original bending moment in Figure 6.21 (from which the same initial and last
parts were also not considered).
It would be interesting to characterize the relative contribution of each component (quasi-static and
dynamic) to the total fatigue condition of a wind turbine. In that sense, numerical analyses were
performed to assess each contribution for different wind conditions. However, it was noticed that the
contributions are dependent on the control system (mainly, on the pitch control) and, for that reason,
they are not generalizable to any wind turbine.
240
Chapter 6
The characterization of the mode shapes during a training period was used in this work. With this
approach, a suitable number of temporary sensors is installed at the wind turbine support structure
during a short period of time (e.g. between two scheduled maintenance operations). Throughout this
period, the modal tracking of the wind turbine modes is accomplished with all the sensors (permanent
plus temporary), allowing to characterize the modal amplitudes at the measurement positions, under
different operating conditions. In addition, the characterization of the operational deflection shapes of
the wind turbine vibrating at harmonics frequencies along the operating regime of the turbine is also
necessary. The number of installed sensors during this period has to be sufficient to obtain a
satisfactory characterization of these vibration configurations.
After the training period, the mode shapes (and operational deflection shapes) are well characterized at
the positions of the permanent and temporary sensors. At this point, the continuous monitoring of the
wind turbine is performed with the single sensor. As result, the identified modes shapes are
characterized by a single modal amplitude per direction (at the permanent sensor position). These
modal amplitudes are used to scale the contribution of mode shapes (and operational deflection
shapes) characterized during the training period. With this step, a new matrix is defined and the
estimated acceleration time series at the unmeasured locations are obtained with:
. , (6.12)
Example
The methodology based on virtual sensors was applied to the numerical NREL 5MW wind turbine
example. In this example, the recorded acceleration time series from one sensor was used to predict the
acceleration at the location of the other two sensors. The mode shapes obtained with the analysis in which
all sensors were considered were adopted to define the matrix, simulating the results from a training
period.
Three different analyses were then performed, considering in each one a different single sensor. Initially,
the single acceleration time series was used as input for the output-only modal identification algorithm.
For this example, the SSI-COV algorithm was selected. From the identification results, a suitable model
order was considered to define the state-space model.
Considering the selected state-space model, the forward innovation model was obtained as introduced in
section 5.5.2. Then, the acceleration time series at the location of the non-considered sensors were
estimated through equation (6.12).
Figure 6.22 illustrates the results obtained with this analysis. In this figure, each column is referred to an
analysis where a single sensor was used to estimate the acceleration from the other two. For example, the
first column is referred to the analysis where the S1 sensor was used as single sensor. The predicted
acceleration to the considered single sensor is also presented in each column to show the error associated
with application of the model (see equation (5.73)).
The results presented in this figure show a good agreement between the predicted and original time series
at the unmeasured places. However, it is noticed that the consideration of the top sensor S3 as single
sensor produces the worst results. This situation is due to the fact that the acceleration response at the S1
and S2 location is mainly dominated by vibration modes with reduced modal amplitude at the tower top.
241
Fatigue Assessment of Wind Turbines
Consequently, the identification of these modes will be worse when a sensor at the tower top is considered
as single sensor, leading to a poorer agreement between the predicted and the real acceleration signal.
Virtual S3 Virtual S3 S3
Virtual S2 S2 Virtual S2
S1 Virtual S1 Virtual S1
Original
Prediction
Sensor S1 Virtual Sensor S1 Virtual Sensor S1
0.4 0.4 0.4
Acceleration [m/s2]
0 0 0
0 0 0
0 0 0
Figure 6.22 – Comparison between the recorded and estimated acceleration at the sensors position when a single sensor
is used to estimate the time series of the other two sensors. Each column is referred to the results obtained with the
consideration of a different single sensor
242
Chapter 6
243
Fatigue Assessment of Wind Turbines
244
Chapter 7
7
STRUCTURAL MONITORING OF
WIND TURBINES
245
Structural Monitoring of Wind Turbines
o The continuous assessment of the structural condition and operation of the support structure
might give a good support to decide about the extension of operation after the initial
predefined period of life or, in the other hand, if the structure should be dismantled (or
reinforced) before the expected period of life due to unforeseen loads and/ or degradation.
Considering specifically offshore wind turbines, the installation of a monitoring system at the support
structure is even more advantageous. Being this type of structures fairly recent, the degree of
knowledge is not as mature as for onshore systems. Mainly for this reason, some additional
motivations should be considered for the installation of a monitoring system:
o The effects of the sea acting at the structure (mainly waves) are still not mastered. For this
reason, there is some uncertainty when dealing with the fatigue life of these structures.
Furthermore, some structural aspects, like foundations and the grouted joint in the transition
piece (Rücker, 2007), are prone to rapid wear, which may require specific structural health
assessment procedures;
o One specific problem of offshore wind turbines is scour around the foundation piles. This
problem increases the flexibility of the entire structure and leads to reduction of the support
structure natural frequency (Zaaijer, Tempel et al., 2002; Prendergast, Gavin et al., 2015). This
situation may lead to resonance problems if the natural frequencies of important support
structure vibration modes are reduced to values close to the harmonics of the rotor speed.
Figure 7.1 shows the evolution of the 1st and 2nd natural frequency values of a monopile
support structure. It is seen that these modes (especially the 2nd mode) are very sensitive to this
problem;
o Conversely, the increasing of the foundation stiffness might also represent a problem when the
correspondent increase of the frequency values brings them close to harmonics. This is the
case of soft-stiff solutions. A real case study in which a progressive increase of the modes
frequency values was noted is presented in (Weijtjens, Verbelen et al., 2015). It is referred that
this phenomenon may be related to soil stiffening due to cyclic loading.
1
S - Scour depth
0.95 D - Monopile diameter
Relative natural frequency
0.90
0.85
0.80
Figure 7.1 – Variation of natural frequencies of offshore wind turbines with scour (adapted from (Zaaijer, Tempel et al.,
2002) and (Prendergast, Gavin et al., 2015))
Thus, the implementation of a monitoring system in wind turbines allowing the assessment of the
actual condition of the components and support structure would introduce considerable
246
Chapter 7
247
Structural Monitoring of Wind Turbines
Particle count
SCADA
(temperature)
Life time
Figure 7.2 – Detection of damage by different systems defined by GL Guidelines (Germanischer Lloyd (GL), 2003)
248
Chapter 7
LEVEL 1 – DETECTION (qualitative assessment that damage is present (or not) in the structure)
LEVEL 4 – CONSEQUENCE (information about the actual safety of the structure, prognosis of the
remaining lifetime)
SHM systems have been applied for years in different types of civil engineering structures. Magalhães
et al (Magalhães, Cunha et al., 2009) implemented a monitoring system in concrete arch bridge. Hu et
al (Hu, Moutinho et al., 2012) implemented a similar SHM system in a footbridge. Brownjohn
(Brownjohn, 2007) presents several examples of SHM applications in a variety of structures, among
which dams, offshore installations, nuclear installations and tunnels. Some examples of very complete
SHM systems installed in bridges are presented in (Boller, Chang et al., 2009).
Lately, the application of SHM to wind turbines has been one of the main topics of study. Blades have
been the preferred substructure of study due to their high propensity for failures. However, recently
also the support structure (tower and foundation) have focused some attention concerning safety
issues. This is mainly due to the increasingly higher dimensions of the tower, alongside with larger
loads and masses at the top, due to larger rotor diameters, and due to the lack of knowledge of the
environmental situations at sea and greater wear of structures in offshore installations.
Several methods of SHM for application in wind turbines have already been developed. An exhaustive
list of methods (referred to blades) is introduced in (Boller, Chang et al., 2009) and (Ciang, Lee et al.,
2008). Also in (Wymore, Van Dam et al., 2015), a general overview of the state of development of
health monitoring systems is presented. When considering SHM for support structures, essentially two
249
Structural Monitoring of Wind Turbines
methods are used: vibration-based and strain measurement. The present work deals with the
application of SHM to the support structure of wind turbines through the application of vibration-
based methods.
Some numerical studies about the feasibility of using the frequency values of the vibration modes to
identify structural damage in wind turbines are present in the literature. Adams et al (Adams, White et
al., 2011) simulated three different types of damage in different locations in a numerical model: blade
root section, low-speed shaft and yaw joint. Through a sensitivity analysis, the variations of the first 14
natural frequencies were calculated for each damage scenario. Also, some developments have been
made in order to assess the evolution of the frequency values of the support structure with scour
(Zaaijer, Tempel et al., 2002; Sørensen and Ibsen, 2013; Prendergast, Gavin et al., 2015).
A good example of the use of vibration-based methodologies for damage detection on the support
structure is given in (Pelayo, López-Aenlle et al., 2011). In this work, high levels of vibration, together
with cracks at the foundation, were detected in two wind turbines. For diagnosing the cause of this
anomaly, operational modal tests were performed in these two turbines, alongside with a similar but
healthy wind turbine. For these tests, accelerometers were placed on the foundation and on the lower
part of the tower. The results showed that both damaged wind turbines presented lower values of
natural frequencies than the healthy turbine (around 7 % to 3 %). After this assessment, the
foundations of the two wind turbines were reinforced. Then, the tests were once again performed and
the results indicated an increase in the natural frequencies (in line with the healthy wind turbine).
Also, in (Currie, Saafi et al., 2015), problems related to high levels of vibration at the foundation level
of onshore wind turbines, due to damage around the bottom flange of the embedded ring, were
described.
In the experimental field, some cases of implementation of vibration-based monitoring systems are
described in literature. During the HISTWIN project (Veljkovic, Heistermann et al., 2012), a
monitoring system was implemented on a wind turbine for almost one year. For the instrumentation,
accelerometers, strain measurements, temperature sensors and inclinometers were used, together with
some SCADA data. During this period, the modal identification of the tower was performed, as well as
the fatigue assessment of the tower and bolts from the flanged connections between segments.
A different approach for a monitoring system is described by Braam et al (Braam, Obdam et al., 2012).
In this work, a strategy of monitoring an entire wind farm with the installation of additional sensors in
just a few reference turbines is presented. On the reference turbines, fibre optic sensors were installed
on the blades with the aim of measure strains (load indicators) and evaluate the fatigue state of these
turbines. The relations between these load indicators and the parameters measured by the SCADA
system (present in every turbine of the wind farm) were then established, in order to extrapolate the
load accumulation of the non-reference turbines. This methodology allows defining a maintenance
strategy considering the most mechanically loaded wind turbines. In (Guo and Infield, 2012), also a
more economical SHM system is presented, since it only uses built-in SCADA parameters. It consists
in the processing of data from the SCADA system, mainly 10 min. averaged values of the tower top
and drive train accelerometers. This data is then correlated with other SCADA data and analysed in
the context of the different operational regimes of the rotor.
Also in (Hu, Thöns et al., 2015b; Hu, Thöns et al., 2015a), a study about the implementation of a
vibration-based monitoring system at a prototype 5 MW wind turbine with tripod foundation installed
onshore is described in two parts. For this system, 8 accelerometers distributed along the support
250
Chapter 7
structure were used. In the first part, the main results concerning the dynamic characterization of the
wind turbine are presented. Then, in the second part, the modal results obtained during a period of 2
years were used to assess the detection of damage at the tower and foundation.
The offshore Vestas V90-3.0 MW, introduced in section 5.9, is also used in the context of a project
aiming to develop monitoring systems based on the dynamic properties of wind turbines. The results
obtained up to date are described in (Weijtjens, Verbelen et al., 2015).
The assessment of the variation of natural frequencies in order to detect blade damage is also a field of
current study. The majority of published works, both in the experimental (Wang, Liang et al., 2014;
Ulriksen, Tcherniak et al., 2015) and numerical domain (Hoell and Omenzetter, 2014), refer the small
sensitivity of the frequency values of the first order blades vibration modes to common small damages.
Notwithstanding, in (Marulo, Petrone et al., 2015), an experimental study of the sensitivity of the
dynamic properties of a small-scale blade to a very severe damage is described. The blade was tested to
buckling and its modal properties were analysed before and after the buckling. In this case, the
variation was notorious, exposing that the application of this kind of monitoring technique may be
only effective for detection of heavy damages.
251
Structural Monitoring of Wind Turbines
252
Chapter 7
In order to understand and remove the influence of environmental and operational effects on the
modal properties of the structure, three different types of approaches can be followed: physical models,
models whose environmental effects are used as input and models which do not consider these inputs.
The last two models were tested in the context of the dynamic monitoring system described in this
work. These methods are considerable easier and quicker to implement and operate, however they
required a period of training in which a representative variation of the environmental and operational
effects should be included.
The methods considering the measurement of the environmental and operational effects in the
conception of the model are defined as input-output methods. The application of these methods for
removal of the external influences in the context of wind turbines seems, from a conceptual point of
view, very suitable. Since the SCADA system collects a variety of operational and environmental data
related to the wind turbine in which it is installed, the access to these records is easier and do not
require the installation of any additional sensor. In this approach, two methodologies are described
based on Multivariate Linear Regression Models and Dynamic Regression Models.
On the other hand, the approach not requiring the measurement of the operational/ environmental
effects is referred as output-only methods. In the context of wind turbines, output-only methods can
be also used in association with the previously presented input-output methodologies with the aim of
achieving a “cleaner” time evolution of the natural frequencies. This can be obtained since there may
be external factors not recorded by the SCADA system that also influence the modal properties.
Within this scope, the approach based on the Principal Component Analysis was tested (Johnson and
Wichern, 2002). However, it was observed that the application of this method did not improve the
results obtained with the linear regression models. For that reason, this method was not considered in
the context of the monitoring system described in this work.
253
Structural Monitoring of Wind Turbines
The mathematical definition of a Multivariate Linear Regression Model is provided by (Johnson and
Wichern, 2002):
. (7.1)
with:
Matrix with the values of the predictors variables (for operational/ environmental effects)
The vector , which represents the random error due to measurement issues and due to the existence
of variables not explicitly considered in model, presents the following properties (Johnson and
Wichern, 2002):
0
(7.2)
. .
with:
∗ Transpose operator
∗ Standard deviation
Identity matrix
When defining the matrices from equation (7.1), some considerations are employed. The first column
of is filled by ones, so this matrix contains 1 columns. Also, the values of are usually
normalized, being each element subtracted from the mean values and then divided by the standard
deviation of the variable:
(7.3)
with:
Normalized value of
With this configuration, the first element of the vector contains the mean value of the observed
variable.
254
Chapter 7
Before the definition of the model, it is needed to assess which predictors influence the response
variable. For this step, the predictors with high correlations with the response variable and low
correlations with the other predictors should be selected (Magalhães, 2010). A suitable tool for the
evaluation of the correlation between two variables and is the Pearson’s product-moment
correlation coefficient (Johnson and Wichern, 2002):
,
. (7.4)
where values close to zero represent a low linear correlation, while values close to one (in absolute
value) might indicate a good correlation. It should be noted that this tool only considers linear
correlations. Thus, if a zero value is obtained, it does not imply that a non-linear relation is also not
present.
Once the choice of the relevant predictors is complete, the vector β should be estimated. This vector
(named as ) is obtained through application of the least squares method and defined in (Johnson and
Wichern, 2002):
. . . (7.5)
Since the weight factors of the predictor variables are already defined, the vector should then be
evaluated. In that sense, an estimation of the residual random errors is given by the difference between
the real and estimated response variable (Johnson and Wichern, 2002):
̂ . . . . (7.6)
Two ways for assessing the quality of the model adjustment can be used. The first one is through the
computation of the estimation of variance of ̂:
1
. (7.7)
However, this indicator has the disadvantage of being dependent of the absolute values of y variable.
To solve this problem, a coefficient is presented in (Johnson and Wichern, 2002) that avoids this
inconvenience:
∑ ∑
1 (7.8)
∑ ∑
255
Structural Monitoring of Wind Turbines
The coefficient is a ratio of variances that aims to quantify the proportion of variance of
“explained” by the selected predictor variables (it varies from 1 to 0, where 1 holds to situations where
the adjust passes through all data points – a “perfect adjust” -, while 0 corresponds to a total
independence between response and predictor variables).
After the assessment of the residual errors is complete and the results are accepted, the model is ready
to predict future values of the response variable ( ) based only on the predictors not considered for its
definition ( ):
. (7.9)
At this point, the model is capable of predicting natural frequencies (the response variable), based on
the environmental effects (the predictor variables) from the same period. Thus, observed and
forecasted values can be compared. Since errors ( ) from this difference are inevitable, the definition of
a range in which the errors are acceptable is required. In that sense, if the errors are considered as
normally distributed, a confidence interval of 100 1 α % is given by (Johnson and Wichern, 2002):
. 1 . . (7.10)
2
with:
The previous equation set a range ; with a defined confidence level in which
the new values of response variable should lay in. If the previously referred considerations about the
construction of the matrix are met (first column of is filled with ones and the elements of the other
columns are normalized), the confidence interval can be written as (Magalhães, 2010):
. (7.11)
with:
New response variable value (which was not considered for the definition of )
The previous equation can be interpreted in the context of the monitoring situation under study. Once
the training period for the definition of matrix is complete, predictions of the frequency values ( )
from the subsequent periods can be computed. These predictions will be formulated considering the
entire set of environmental effects covered during the training period. For this reason, predicted
frequencies will present variations that are interpreted as normal, i.e., this variations do not refer to
any kind of damage since the training period is assumed to have occurred during a healthy stage of the
structure. When comparing the predicted and real values of the natural frequencies from the
256
Chapter 7
subsequent periods, their difference should be evaluated. If it is abnormally high (i.e., the value is
outside the confidence interval defined in equation (7.11)), the variation cannot be explained by the
environmental effects considered during the training period, and an abnormal condition affecting the
structure, such as damage, may be present.
The generalization of the presented method is easily performed for several dependent variables
(frequency values of several vibration modes). This can be achieved by formulating the previous
methodology as many times as response variables or through the matrix equation:
. (7.12)
with:
One last property from these models should also be referred: Multivariate Linear Regression Models
are not valid for extrapolations. For that reason, a period of training with a wide range of values from
the environmental effects should be used.
257
Structural Monitoring of Wind Turbines
statistical process control, are used to evaluate if one or more variables are kept within predefined
limits, which indicates that the process is occurring without abnormalities. For this reason, control
charts are a very suitable technique to monitor possible frequency deviations due to damage.
A typical control chart is presented in Figure 7.4. It consists of a graphical display, with horizontal
development (usually associated with time), in which observations are plotted. There are also some
control parameters: a centre line, which represents the average value of the quality characteristic under
study; an upper (UCL) and lower control line (LCL) defining an interval in which the observations are
considered “in-control”. If an observation has a value considered as abnormal, it falls outside this
interval and is considered as “out-of-control”.
Out-of-control
Upper control limit
Sample quality characteristic
Center line
A common control chart to assess the evolution of a single variable is the Shewhart -chart. In the
process control, the points (in the plot) under control represent a single observation or the mean of a
group of observation (subsample). For a -chart, the definition of the upper and lower control lines is
given by (Montgomery, 2009):
̅ .
(7.13)
̅ .
with:
Standard deviation. When each point represent one single observation, is the standard deviation
from all observations; if each point represents the mean of a subsample with elements, is given by
the quotient between the standard deviation from all observations and √
“Distance” between the control lines and the centre line. If 3 is defined in a distribution
considered as normal, the control limits correspond to a confidence level of 99.7 %
When dealing with a situation with more than one variable (like vibration modes), it is required a
slightly evolution from a univariate technique (such as the -chart) to a multivariate technique. A
similar technique of the Shewhart -chart for monitoring more than one parameter is the control
258
Chapter 7
chart. The control chart is a simple tool to detect deviations in the characteristics under evaluation.
Although very similar to -charts, control charts present two distinct characteristics: the vertical
axis does not represent the quality characteristic under evaluation, but a statistic test named ; and
the LCL is always zero.
This multivariate statistical parameter control can be also applied to individual observations or
subgrouped data (subsamples). When every observation is checked (i.e. each point of the plot
represents one observation), the - statistic and the UCL are computed according to (Montgomery,
2009):
. ̅ . . ̅
1
(7.14)
1 .
. ,
3
with:
Covariance matrix
If, instead of checking all the observations, subsamples with observations are verified, the -
statistic and the UCL are computed according to (Montgomery, 2009):
. ̅ ̅ . . ̅ ̅
. 1 . 1 (7.15)
. , .
. 1
with:
̅ Subgroup average
̿ Process average
259
Structural Monitoring of Wind Turbines
Support structures, as seen in Chapter 3, are a fundamental component of wind turbines that cannot
fail at any circumstance, otherwise the whole system is jeopardized. Furthermore, their design is
usually very conservative with respect to fatigue, which may lead to real fatigue life considerably higher
than 20 years (the usually considered expected fatigue lifetime). Thus, the implementation of a fatigue
monitoring system for assessment of the real solicitation of the support structure can introduce several
advantages. If the monitoring system indicates that damage is not present in the structure, the fatigue
monitoring system can be extremely useful in the following aspects:
o Estimation of the evolution of real fatigue condition of the support structure and its suitability
with the expected lifetime (usually 20 years);
o Since an estimation of the real condition of fatigue damage is known, the results of the fatigue
monitoring system could be an essential element for a decision about extending the lifespan of
the structure, leading to a higher profitability of the investment;
o Possibility of repowering or overpowering of the system, while keeping the support structure.
If the fatigue monitoring system indicates a considerable margin of safety at the end of the
period of life initially stipulated, together with a situation of non-damage detection by the
dynamic monitoring system, one of these two options could be considered. For that purpose, a
careful inspection of the structure would have to be carried out on-site. However, the results
obtained with both components of the monitoring system, could provide a factual evidence of
the overall good condition and low level of structural loading during the first period of life.
The decision about repowering and overpowering could greatly reduce the cost of tower +
foundation portion, which contributes with 18 % and 23 % of the capital cost for onshore and
offshore installations, respectively (see section 3.1.2).
The description of fatigue monitoring systems and publication of results from long periods of data
acquisition is not very common in literature. In the context of the HISTWIN project (Veljkovic,
Heistermann et al., 2012), a 2.1 MW wind turbine tower was monitored with strain rosettes, among
other sensors. Results from two periods of 139 and 159 days for the most unfavourable structural detail
lead to the conclusion of a considerable overdesign of the fatigue strength.
Another work considering the problem of fatigue in the support structure of a wind turbine is reported
in (Pollino and Huckelbridge, 2012). A 100 kW wind turbine was monitored with strain gauges during
approximately one year. The authors studied the principal structural details of the tower and also
concluded that the fatigue life of these elements were significantly higher than 20 years.
An interesting study is described by Thies et al (Thies, Johanning et al., 2014) about the fatigue damage
assessment of the mooring lines of an offshore floating marine energy converter. This equipment, that
also comprises a 1.2 m diameter wind turbine, was monitored with load cells to measure the tension
forces in the mooring system.
Outside the scope of wind turbine support structures, there is a greater range of studies about the
assessment of fatigue and prediction of the remaining period of life on steel structures. As an example,
Ye et al (Ye, Ni et al., 2012) presented a study about a long-term monitoring data of dynamic strain on
a steel bridge (the Tsing Ma Bridge).
260
Chapter 7
. (7.16)
with:
However, this extrapolation assumes that the monitored period is representative of the loading
scenarios from the whole period of operation of the wind turbine. Considering that the monitored
period may not be representative of the environmental loading along the estimated period of life, a
more complex approach should be followed.
An alternative approach can be obtained if the information about the environmental data is recorded.
In that case, loading scenarios should be defined according to the simultaneous occurrence of the
various environmental parameters. Important parameters such as mean wind speed, turbulence
intensity and wind direction should be considered. Also, the significant wave height and peak period
should be included for offshore installations. The variability of the parameters is usually reduced into
bins with a representative magnitude in order to ease the computation of probability. Figure 7.5 shows
part of the scatter diagram from the OWEZ site, where the probability of occurrence of loading
scenarios with a mean wind speed of 10 m/s combined with various wave conditions is shown. In this
figure, the wave loading condition is defined by the significant wave height (introduced in section
6.2) and the mean zero-crossing period (instead of the peak period ). This period is referred to
the mean value of the zero up-crossings of a point at the sea surface (Tempel, 2006).
261
Structural Monitoring of Wind Turbines
6
5.5
5
4.5
4
3.5
Hs 3
2.5 0.000038 0.000038
2 0.000342 0.000342
1.5 0.017727 0.001674 0.019401
1 0.003918 0.089550 0.093468
0.5 0.048009 0.004831 0.052840
0 0.000266 0.000038 0.000304
0 1 2 3 4 5 6 7 0.166394
0.166394
Vw = 10m/s Tz
Figure 7.5 – Part of scatter diagram corresponding to the mean wind speed of 10 m/s (the values represent the
probability of occurrence of the loading scenarios) (Tempel, 2006)
. , ,… . (7.17)
with:
The monitoring period of the structure (which might be much shorter than the monitoring period of
the loading conditions) should include all the important load conditions for the quantification of D .
262
Chapter 7
(SSI-COV, SSI-DATA and p-LSCF) to the acceleration signals. These algorithms were introduced in
section 5.4. The results obtained are then analysed through a cluster analysis (section 5.7.1), in which
the poles with similar modal properties are gathered into the same clusters. Lastly, these clusters are
compared with reference modal properties (frequency values and mode shapes) of the vibration modes
intended to be tracked in order to separate the clusters referred to these modes from the other ones.
These reference properties are defined during a training period prior to the automated processing.
Once they are defined, these properties are saved in a database, being used in every processing cycle.
From the moment the modal tracking is complete, it is necessary to remove the influence of the
operational and environmental effects on the frequency values of the identified modes. In that sense,
multivariate linear regression models have to be defined during a representative period (again, prior to
the automated process). Once these models are set, they are applied to the identified frequency values
in order to reduce their variability. This process was presented in section 7.2.1.1.
The last step of the module is referred to the assessment of important deviations of the frequency
values (damage detection). The implemented procedure, based on control charts, was introduced in
section 7.2.1.2. A caveat should be made at this level that, if subsamples with more than one
observation are used, the damage detection step is only processed after the required number of
observations is achieved.
The module related to the fatigue assessment of the wind turbine support structure starts after the
application of the output-only identification algorithms to the acceleration signals. At this point, it is
necessary to select an order of the state-space model. This selection should take into consideration the
order with the highest number of identified modes with important contribution to the support
structure motion. For that reason, the results from the modal tracking represent an important
information. Furthermore, the selected order should contain poles related to the low order harmonics
(when the turbine is operating). Thus, the data from the SCADA system is also important at this point.
Once the state-space model order is selected, it is possible to define the forward-innovation model.
With the transformation of the model into the modal basis, it is thus possible to obtain the modal state
vector, essential information to estimate the acceleration at unmeasured locations (this procedure was
presented in section 5.5). At this point, a detailed estimation of the mode shapes along the height of the
support structure is required to estimate the acceleration time series at every location. Two options are
available:
o If the number of installed sensors at the structure is enough to estimate the mode shapes of the
most important modes along the height of the support structure, an interpolation of the
experimentally found mode shapes can be used. This methodology was referred in section
6.4.1;
o On the other hand, if a reduced number of sensors is used, reference support structure mode
shapes of the important modes should be used. In addition, operational deflection shapes of
the harmonics should also be used for situations under operation. Both mode shapes and
operational deflection shapes must be defined after a period of training and should consider
the different operating conditions of the wind turbine. This procedure was described in section
6.4.2.
263
Structural Monitoring of Wind Turbines
With the possibility of estimate the acceleration time series at every position of the wind turbine
support structure, the fatigue damage of the structure is assessed with the procedure proposed in
section 6.4.1. During this process, the stiffness matrix of the support structure is required and should
be previously calculated according to the mechanical properties of the support structure.
Lastly, the fatigue damage estimations from the 10 min. time series should be continuously summed in
order to obtain a cumulative damage of the support structure. Once representative damage
accumulation events from the various environmental loading scenarios are recorded, it is possible to
perform an estimation of the damage condition at the end of the designed fatigue life and, thus, predict
the real lifetime of the structure. This procedure was introduced in section 7.2.2.1.
Signal Decimation
Coordinate Transformation
Automated Modal Analysis
Fatigue Assessment
Definition of the
Modal Identication
State-Space Model
Definition of the
Cluster Analysis
Forward-Innovation Model
Database Database
Removal Operational/
Multivariate Regression Model Fatigue Damage Stiffness Matrix
Environmental Conditions
Accumulated Damage/
Damage Detection
Lifetime Prediction
Figure 7.6 – Chain of events included in the dynamic monitoring system (the dashed box is only referred to situations in
which a reduced number of sensors is used)
264
Chapter 7
7.3.1 INTRODU
UCTION
The SSenvion MM
M82 is a 2.0 MW
M onshore wind turbin ne. This wind
d turbine starrted operatin
ng in 2007 att
the T
Torrão wind farm,
f in the north
n of Porttugal (Figuree 7.7).
265
5
Structural Monittoring of Wind Tu
urbines
Figu
ure 7.8 – Senvio
on MM82 wind turbine installe
ed at the Torrão wind farm
266
Chapter 7
Model Senvion MM 82
General
Rotor
Nº blades 3
Horizontal axis
Rotor orientation
Up-wind rotor
Rotor speed range [rpm] 8.5 – 17.1 (±16 %)
82.0
Rotor diameter [m]
(Blade length: 40 m)
Control systems Pitch and rotor speed control
Machinery
Variable-speed
Generator type
asynchronous generator
3 stage Planetary/ helical gear
Gearbox
(ratio: 105.4)
Yaw system Active system with disc brake
Tower
Aiming to obtain an initial approach of the modal properties of the Senvion MM82 wind turbine, two
numerical models were developed. Initially, the structural behaviour of the tower was modelled with
the ANSYS software (Ansys Inc., 2011), according to technical drawings made available by the
manufacturer. Four-node shell elements with six degrees of freedom at each node were used to model
the shell components of the tower, while the flanged bolt connections were simulated with beam
elements. For the nacelle and rotor blades components, a point element with a representative
concentrated mass was employed on their centre of gravity. The connection between tower and
foundation was defined as fixed. Figure 7.9 presents a global view and a detail from the door opening
of the finite element model.
267
Structural Monitoring of Wind Turbines
1 1
ELEMENTS ELEMENTS
NOV 21 2012 NOV 21 2012
MAT NUM 15:10:03 MAT NUM 15:15:14
PLOT NO. 1 PLOT NO. 1
Y
X
Z
Z X
Figure 7.9 – Global view and detail of the door opening of the Senvion MM82 finite element model
The second numerical model of the Senvion MM82 wind turbine was developed using the HAWC2
aeroelastic code (Larsen and Hansen, 2007). For this model, an improvement on the previous model
developed using ANSYS was introduced. In addition to modelling the tower structure according to
technical drawings, the nacelle and rotor blades were also modelled. Since the characteristics of these
elements were not provided, the characteristics from the NREL 5MW reference wind turbine
(Jonkman, Butterfield et al., 2009) were scaled down to meet the characteristics of the Senvion MM82
wind turbine. The connection between tower and foundation was also defined as fixed.
The results obtained after the modal analysis of both models for the frequency values of the first 2 pairs
of tower bending modes are shown in Table 7.2. It is observed that the results achieved with both
models are coherent, giving a good indication about the expected frequency values of these modes.
Table 7.2 – Natural frequency values of the first 2 pairs of tower bending modes obtained with the numerical models
developed using ANSYS and HAWC2
1 SS 0.378 0.366
1 FA 0.378 0.361
2 FA 2.963 2.771
2 SS 2.967 2.830
The mode shapes of the referred modes are illustrated in Figure 7.10. It is seen that second pair of
bending modes includes an important participation of rotor blades motion, while this contribution is
imperceptible in first pair.
268
Chapter 7
1 SS
1
NODAL SOLUTION
MX NOV 21 2012
STEP=1 14:32:38
SUB =2 PLOT NO. 1
FREQ=.378684
USUM (AVG)
RSYS=0
DMX =.002921
SMX =.002921
Y
Z X
MN
0 .649E-0
3 .001298 .00194
7 .002597
.325E-03 .974E-03 .001623 .002272 .002921
Senvion MM 82
1 FA
1
NODAL SOLUTION
NOV 21 2012
STEP=1 14:30:44
SUB =1 PLOT NO. 1
FREQ=.378154
USUM (AVG)
RSYS=0
DMX =.002921
SMX =.002921
Y
Z MN X
0 .649E-0
3 .001298 .00194
7 .002597
.325E-03 .974E-03 .001623 .002272 .002921
Senvion MM 82
2 FA
1
NODAL SOLUTION
NOV 21 2012
STEP=1 14:32:53
SUB =3 PLOT NO. 1
FREQ=2.96267
USUM (AVG)
RSYS=0
DMX =.004349
SMX =.004349
MX
Y
Z MN X
0 .966E-0
3 .001933 .00289
9 .003865
.483E-03 .00145 .002416 .003382 .004349
Senvion MM 82
2 SS
1
NODAL SOLUTION
NOV 21 2012
STEP=1 14:32:14
SUB =4 PLOT NO. 1
FREQ=2.96982
USUM (AVG)
RSYS=0
DMX =.004357
SMX =.004357
MX
Y
Z X
MN
0 .968E-0
3 .001936 .00290
5 .003873
.484E-03 .001452 .00242 .003389 .004357
Senvion MM 82
Figure 7.10 – Mode shapes of the first 2 pairs of tower bending modes obtained with the numerical models (left column:
ANSYS; right column: HAWC2)
269
Structural Monitoring of Wind Turbines
1500
Power [kW]
1000
500
0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28
Wind speed [m/s]
Figure 7.11 also indicates the various operating conditions of the wind turbine. The first region is
upper-limited by the cut-in speed of the turbine and is characterized by very low wind speeds. In this
region, the wind speed is not enough to create a consistent rotor torque and, thus, the turbine does not
operate and the rotor is usually parked or idling. While in this region, the control strategy of the
turbine involves the analysis of the wind condition in order to evaluate a feasible start of the turbine
operation.
When the wind speed is higher than the cut-in value, the turbine enters in Region 2. In this region, the
wind turbine starts the power production regime in which the rotor operates close to an optimal tip-
speed ratio. This regions corresponds to line BC’ in Figure 3.39 and characterized by an important
variation of rotor speed.
Once the Region 3 is achieved, the wind turbine enters in a regime defined by line C’D’ in Figure 3.39.
This is a transition condition in which the rotor torque achieves its rated value while the rotor speed
slightly increases. During both Regimes 2 and 3, the pitch angle of the blades is kept constant on its
minimum value and the system acts on the yaw actuator to keep the rotor orientated to the main
direction of the wind.
Region 4 is defined by the start of operation of the blades pitch actuator. In this region, the rotor
reaches its rated speed and the blade pitch angle is adjusted to keep the rotor speed under defined
values.
Once the turbine is operating at its rated power, the Region 5 is reached. In this region, the main goal
of the control system is to keep constant this production level. For that goal, the blade pitch angle is
controlled over this region, increasing its value with the increase of the wind speed. The actuation of
the control system over the blade pitch permits to control the rotor speed, keeping the production
constant, as well as to avoid exceeding electrical and mechanical load limits.
270
Chapter 7
The last region of Figure 7.11 is referred for wind speeds higher than the cut-out value. At this point,
the wind load acting on the system is excessive and the production is stopped. In order to avoid
extreme loads, the turbine is shutdown, the blades are pitched out of the wind.
These six regions define, in a simplistic way, the control system of a wind turbine for different wind
speed conditions. Thus, if it is desired to track the modal properties of the wind turbine over its period
of life, the variations occurring within each region need to be considered. For that reason, the data
recorded by the SCADA system has an important role in the monitoring system.
The different operating regions are also evident in the evolution of the rotor speed with the wind speed
(Figure 7.12). In this figure, the rotor speed range from the variable-speed generator system is evident.
During the period in analysis, Region 6 was not detected.
Region Region Region Region Region
1 2 3 4 5
18
16
14
Rotor speed [rpm]
12
10
0
0 5 10 15 20 25
Wind speed [m/s]
The behaviour of the pitch angle with the rotor speed is depicted in Figure 7.13. In this figure, the
different operating conditions of the wind turbine system are also visible. Up to the wind cut-in speed,
the pitch angle is kept with high values (around 70º to 90º). Under this condition, the blades are only
able to slowly rotate, which is desire since the wind speed is too low for production. From the moment
the wind speed is higher than the cut-in speed, the pitch decreases and the generator starts operating
(from around 8.8 rpm). The pitch angle is kept low until the Region 4 is reached. Since the pitch
actuator acts practically at the rated rotor speed, the increase of the pitch angle in Region 4 and 5 is not
completely perceptible in the plots of Figure 7.13.
271
Structural Monitoring of Wind Turbines
100 25
80 20
15
Pitch angle [º]
40 10
20 5
0 0
0 5 10 15 8 10 12 14 16
Rotor speed [rpm] Rotor speed [rpm]
a) b)
Figure 7.13 – Pitch angle vs rotor speed: a) all regimes; b) only operating regimes
The illustration of evolution of the blades pitch angle actuator with the wind speed (Figure 7.14) is
helpful to understand the pitch mechanism in Region 4 and 5. In this figure, the start of the actuator
around 11 m/s is clearly visible. From the observation of Figure 7.14 b), it is also possible to attest that
the increase rate of the pitch angle with the wind speed is lower in Region 5 than in Region 4.
Region Reg. Region Reg. Region
1 2 3 4 5
100 25
80 20
15
Pitch angle [º]
60
40 10
20 5
0 0
0 5 10 15 20 25 0 5 10 15 20 25
Wind speed [m/s] Wind speed [m/s]
a) b)
Figure 7.14 – Pitch angle vs wind speed: a) all regimes; b) only operating regimes
The analysis of the recorded SCADA data also permitted to conclude that the main wind incidence
direction is around 110º. It is concluded, from the wind rose of Figure 7.15, that the higher wind speed
regimes are also obtained for this direction.
272
Chapter 7
0º
NORTH
15%
10%
N
5% W E
Wind speed
110º
270º 90º [m/s]:
WEST EAST S
22 - 24
20 - 22
18 - 20
16 - 18
14 - 16
12 - 14
10 - 12
8 - 10
6-8
4-6
SOUTH 2-4
0-2
180º
Figure 7.15 – Wind rose histogram of mean wind speed Figure 7.16 – Local orography on the wind turbine site
The local orography at the wind turbine site is represented in simplified form in Figure 7.16. In this
figure, the origin of the referential represents the wind turbine location. It is interesting to note that at
the direction of 110º, the terrain presents an apparent depression.
S1
S3
S9 S4
+ 74.988
S8 S2
S6
+ 48.392
S5 S7 + 48.392 + 74.988
S4 S6
S3 + 21.772 S7
S9
S5 S8
S1 S2
+ 0.000
a) b)
Figure 7.17 – Position of the accelerometers at the different levels of the wind turbine: a) front view; b) top view of the
instrumented sections
273
Structural Monittoring of Wind Tu
urbines
a) b)
c) d)
wer flange; and c) foundation flange. d) Centtral
Figure 7.18 – Accelerometers installed at: a) tower fla nge; b) top tow
acquisition
a systtem with the CR
R-5P unit
274
Chapter 7
The dynamic monitoring system is complemented by the SCADA system of the wind turbine. This
system records the mean, maximum and minimum value from 10 min. period of several operational
and environmental parameters. Among them, some were important for the context of this structural
health monitoring project:
o Wind speed and direction;
o Rotor speed;
o Yaw angle;
o Blades pitch angle;
o Outdoor temperature.
Since it was not possible to connect the central acquisition system to an external GPS antenna, the data
recorded by the CR-5P was manually synchronized to the SCADA data. This operation was tuned
using the date information of startup and shut-down events of the turbine recorded by the SCADA
data, which were correlated to events of sudden increase or decrease of the tower vibration levels.
The dynamic monitoring system was installed at the wind turbine on the 22th of July of 2013.
Unfortunately, problems related to electrical power supply of the central acquisition system prevented
a continuous operation. Moreover, it was only possible to have access to the equipment during
maintenance periods, which annulled the possibility to make rapid interventions once the problems
were noticed. For that reason, it was not possible to avoid some periods with a low rate of recorded
setups (or even inactivity of the system). A total number of 37 685 data sets were collected during this
period. The percentage of recorded setups for each month between 22/07/2013 and 21/07/2014 is
evidenced in Table 7.3.
Month 7 8 9 10 11 12 1 2 3 4 5 6 7
Recorded
100 100 63 0 23 89 89 100 100 94 67 59 53
Setups [%]
275
Structural Monitoring of Wind Turbines
1 0.4 0.2
Acc. [m/s2]
Acc. [m/s2]
Acc. [m/s 2]
0 0 0
-1 -0.4 -0.2
-2 -0.8 -0.4
0 200 400 600 0 200 400 600 0 200 400 600
time [s] time [s] time [s]
Figure 7.19 – Acceleration time series Figure 7.20 – Acceleration time series Figure 7.21 – Acceleration time series
from a setup with the highest during a startup event during a shutdown event
observed mean wind speed (23.3 m/s)
The variation of the vibration levels during the monitored period is shown in Figures 7.22 and 7.23 for,
respectively, the RMS and maxima values from each recorded 10 min. time series. The maximum
acceleration recorded in FA direction was 3.73 m/s2 while in the SS direction was 3.10 m/s2. As
expected, the vibration levels at the foundation are considerably lower than in the tower.
RMS
0.6
+74.988 FA
Acc. [m/s 2]
0.4 +74.988 SS
+48.392 FA
0.2 +48.392 FA
+48.392 SS
+21.772 FA
0 +21.772 SS
0 0.5 1 1.5 2
Index 4
x 10
-3
x 10 RMS
5
Acc. [m/s 2]
2.5 Found. 6
Found. 7
0
0 0.5 1 1.5 2
Index 4
x 10
Figure 7.22 – Time evolution of the RMS values of the acceleration time series collected during the monitored period
276
Chapter 7
Maxima
4
3 +74.988 FA
Acc. [m/s 2] +74.988 SS
2 +48.392 FA
+48.392 FA
1 +48.392 SS
+21.772 FA
0 +21.772 SS
0 0.5 1 1.5 2
Index 4
x 10
Maxima
0.03
Acc. [m/s 2]
0.02
Found. 6
0.01 Found. 7
0
0.5 1 1.5 2
Index 4
x 10
Figure 7.23 – Time evolution of the maxima values of the acceleration time series collected during the monitored period
It is interesting to observe that the higher values of acceleration do not occur at the tower top but at
+48.392 m level. This is a common situation when the turbine is operating (the majority of the time).
However, when the turbine is parked or idling, the highest vibration levels were recorded by the top
sensors (+ 74.988 m). This behaviour was also noticed in the analysis of the offshore Vestas V90 wind
turbine (section 5.9).
Figure 7.24 shows the RMS values of acceleration in the FA direction for operating and non-operating
situations. It can be seen that, when the turbine is operating, the acceleration level at the +48.392 m
height is clearly higher than at the other two levels. Indeed, the vibration levels at the top (+ 74.988 m)
and +21.772 m are similar. On the other hand, when the turbine is not operating, the level of vibration
usually increases from the lowest part to the top of the tower structure. Nevertheless, the vibration at
the +48.392 m sometimes exceeds the value from the top sensor. The causes for this behaviour will be
further analysed in section 7.3.6.4.
Operation Parked/ Idling
0.5 0.07
+74.988 FA +74.988 FA
+48.392 FA 0.06 +48.392 FA
0.4 +21.772 FA +21.772 FA
0.05
Acc. [m/s 2 ]
Acc. [m/s 2 ]
0.3 0.04
0.2 0.03
0.02
0.1
0.01
0 0
4000 4100 4200 4300 4400 4500 1250 1300 1350 1400 1450 1500 1550 1600
Index Index
a) b)
Figure 7.24 – Time evolution of the RMS values of the acceleration time series during: a) operating conditions and; b)
parked or idling conditions
277
Structural Monitoring of Wind Turbines
The wind excitation is, naturally, the main driver of wind turbine vibration. Figure 7.25 presents the
time evolution of the RMS values at the +74.988 m level in the FA direction and of the wind speed. As
expected, there is a clear coherence between acceleration level and wind speed.
The relationship between the RMS values of acceleration and the wind speed is also illustrated in
Figure 7.26. From the figure, it seems that the acceleration continuous increases with the increase of
the wind speed.
0.25 25 0.6
+74.988 FA
+48.392 FA
0.5
0.2 20 +21.772 FA
0.4
Acc. [m/s 2]
Acc. [m/s 2]
0.15 15
0.3
0.1 10
0.2
0.05 5
0.1
0 0 0
8600 8800 9000 9200 9400 9600 0 5 10 15 20 25
Index Wind speed [m/s]
Figure 7.25 – RMS values of acceleration of the +74.988 m Figure 7.26 – Correlation of the RMS values of the sensors
sensor in the FA direction vs the wind speed in the FA direction with the wind speed
The variation of the vibration amplitude with the yaw angle is an important analysis to check the
directions with greater wear of the support structure. As expected, the vibration level is highly
dependent on the rotor orientation (which is conditioned by the wind direction). Figure 7.27 presents
the RMS values of the acceleration in FA direction according to the yaw angle. As can be seen, this
figure is coherent with the main wind directions illustrated in the wind rose of Figure 7.15.
0.6
+74.988 FA
+48.392 FA
0.5
+21.772 FA
0.4
Acc RMS [m/s 2]
0.3
0.2
0.1
0
0 40 80 120 160 200 240 280 320 360
Yaw angle [º]
Figure 7.27 – RMS values of the acceleration in the FA direction according to the yaw angle
Figure 7.27 clearly indicates the directions according to which the structure vibrates with the highest
levels. However, as the wind speed is not evenly distributed across all the directions (as shown in
Figure 7.15), it is not possible to make any consideration about the heterogeneity of the support
structure just with this observation.
278
Chapter 7
In order to compare the vibration amplitude for different angles of nacelle orientation, only the
recorded setups from a narrow range of wind speeds, under operating conditions, were selected. With
this consideration, it can be assumed that the wind excitation is roughly similar across all directions.
Still, differences can exist in terms of wind turbulence.
Figure 7.28 presents the box plots obtained for the RMS values of acceleration in the FA direction for
the top sensor considering only setups when the turbine was operating and the wind speed was
between 6 and 7 m/s. For the analysis, the yaw angles were grouped in 18 ranges of 20º, from 0º to
360º. In this figure, the line connects the median value of each group, while the edges of the box
represent the 25th and 75th percentiles. The whiskers were extended to the most extreme points not
considered as outliers. In the background, the data considered for the analysis is plotted in light grey.
This figure clearly indicates that, for roughly similar wind conditions, the structure presents higher
vibration levels for the ranges 100º – 140º and 180º – 240º than for remain directions. The first range
coincides with the main direction of the wind.
Wind speed: 6 - 7 m/s 0.12
0.2 Wind speed [m/s:]
3-4
0.1 4-5
5-6
0.15 6-7
0.08
7-8
Acc. [m/s 2]
Acc. [m/s 2]
0.1 0.06
0.04
0.05
0.02
0 0
0 40 80 120 160 200 240 280 320 360 0 40 80 120 160 200 240 280 320 360
Yaw angle [º] Yaw angle [º]
Figure 7.28 – Box plots of the RMS values of acceleration of Figure 7.29 – Median values of the RMS values of
the top sensor in the FA direction according to the yaw acceleration of the top sensor in the FA direction
angle, considering setups when the turbine was operating according to the yaw angle, considering setups when the
and the wind speed was in the 6 – 7 m/s range turbine was operating and several wind speed ranges
This analysis is further extended to ranges of wind speed whose number of recorded setups is
representative of all yaw sectors. For that reason, ranges with the largest wind speeds were not
considered since they mainly occur in a few yaw sectors. Figure 7.29 presents the results obtained with
this analysis, considering wind speed ranges between 3 and 8 m/s. In this figure, only the median
values are shown. The results obtained are in line with the conclusion already introduced for Figure
7.28. It is interesting to note that, for example, the highest values obtained with the wind speed range 4
– 5 m/s are similar to the values obtained for the 7 - 8 m/s in the 300º - 360º region. From this analysis,
it seems that the foundation is less stiff along the directions that are more excited by the wind.
However, it should be noticed that this analysis should have taken into account the turbulence of the
wind flow, which also influences the dynamic excitation of the support structure. Since this
information is not provided by the SCADA system, it is not possible to confirm the presented
considerations.
It is also interesting to assess the evolution of RMS values of acceleration with the rotor speed (Figure
7.30). As expected, the vibration levels tend to increase with the increase of the rotor speed.
279
Structural Monitoring of Wind Turbines
0.6
+74.988 FA
+48.392 FA
0.5
+21.772 FA
0.4
Acc. [m/s 2]
0.3
0.2
0.1
0
0 2 4 6 8 10 12 14 16
Rotor speed [rpm]
The colour map presented in Figure 7.31 illustrates the frequency content of the acceleration signals
during a period of one month. This figure is a top view of the first singular value spectra of the
spectrum matrices obtained from each recorded setup. The regions with the hotter colours represent
the highest energy. Two vertical alignments with high energy are clearly visible around 0.35 Hz and
2.80 Hz (indicated by arrows). These frequency values are coherent with the results obtained for the
first and second pairs of tower bending modes of the numerical models described in section 7.3.2.
Albeit with less energy, two additional alignments are also visible around 1.30 Hz and 1.80 Hz (also
indicated by arrows).
In the same figure, the rotor speed frequencies from the 1Ω, 3Ω and 6Ω harmonics are represented by
dashed lines. As expected, the excitation introduced by the rotor rotation is visible, mainly for the 3Ω
harmonic. In a small scale, the 6Ω harmonic is also detected and seems to cross the 1.30 Hz alignment
several times. The 1Ω harmonic alignment is very tenuous and is only visible in a few setups.
1Ω 3Ω 6Ω
Time
Figure 7.31 – Colour map with the variation of the signals frequency content during the 15/02/2014 and 15/03/2014
280
Chapter 7
-2
10
Amplitude
-4
10
-6
10
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Frequency [Hz]
Figure 7.32 – ANPSD of two different setups: under non-operating conditions (rotor speed = 0 rpm; wind speed = 1.9 m/s)
and operating conditions (rotor speed = 16.4 rpm; wind speed = 10.7 m/s). The vertical dashed lines indicate the
frequency values of the rotor speed and its harmonics
It is thus important to perform a preliminary analysis of the properties of the most relevant vibration
modes and their variation throughout the various operating conditions. In that sense, the recorded
setups were initially processed with the modal identification algorithms introduced in section 5.4.
Then, an automatic interpretation of the produced stabilization diagrams was performed by the
algorithm based on the hierarchical clustering presented in section 5.7.1. The same datasets used to
plot the ANPSD in Figure 7.32 are used to illustrate this preliminary analysis. The methodology is
similar to the one described in section 5.9 for the case study related to the Vestas V90.
281
Structural Monitoring of Wind Turbines
Figure 7.33 shows the stabilization diagrams associated with the selected datasets provided by the SSI -
COV algorithm, including an averaged spectrum at the background. As expected, the spectrum and
the stabilization diagram associated with the parked condition are clearer. Still, in the other
stabilization diagram several vertical alignments of stable poles can be easily identified.
Poles Stab. freq Stab. damp Stab. MAC ANPSD
60 60
.
50 50
Model Order
Model Order
40 40
30 30
20 20
10 10
0 0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5
Frequency [Hz] Frequency [Hz]
Figure 7.33 – Stabilization diagrams obtained with the SSI-COV algorithm for operating and non-operating conditions
The automatic processing of the stabilization diagrams generated by the identification algorithms (the
three tested methods produce similar stabilization diagrams) is then obtained with the cluster
algorithm. Figure 7.34 characterizes the obtained clusters by the average frequency of the poles
included in each cluster and by the number of poles. At this preliminary stage, it was decided to
exclude just the clusters with a very low number of poles (less than 6, as marked by the red dashed
line).
Non-operating conditions Operating conditions
30 25
25
20
20
15
Nº of poles
Nº of poles
15
10
10
5
5
0 0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5
Frequency [Hz] Frequency [Hz]
Figure 7.34 – Characterization of the clusters: average frequency and number of poles included in each cluster
For each considered cluster, a representative value of frequency, damping and mode shape was
retained. Each cluster was then classified as FA or SS, according to the direction of its representative
mode shape. After this preliminary analysis, it is possible to plot all the considered clusters in a
Campbell diagram, considering the main direction of vibration (Figure 7.35). In this figure, several
282
Chapter 7
alignments of clusters are clearly identified, corresponding to vibration modes. However, it also shows
some clusters located around the diagonal dashed lines (corresponding to the harmonics associated
with the rotor rotation). These clusters, are only present when the turbine is in operation (rotor speed
higher than 8.7 rpm), corresponding to poles motivated by the harmonic excitation.
FA
SS
4.5
4 FA
4 3 SS*
3.5 3 FA/ 3 SS
3 2 SS
Frequency [Hz]
2 FA
2.5
2 2 SS*
1.5
1 1 SS*
0.5 1 SS
1 FA
0
0 2 4 6 8 10 12 14 16 18
Rotor speed [rpm]
Figure 7.35 – Campbell diagram of considered clusters of stable poles. The vertical dashed lines separate the non-
production situation, a transition state and the operating regimes of the wind turbine. The diagonal dashed lines
represent the harmonic frequencies associated with the rotor rotation (multiples of 3Ω)
From this figure, some initial considerations should be highlighted. The two well defined vertical
alignments around 0.35 Hz and 2.80 Hz in Figure 7.31 are clearly visible in the Campbell diagram
along the whole rotor speed regimes. In fact, they both correspond to two pairs of closely spaced
vibration modes in the FA and SS directions, as initially indicated by the numerical models. For the
pair of cluster around 2.80 Hz it is interesting to note that for high values of rotor speed (higher than
15 rpm), the number of SS clusters of this pair tend to decrease. On the other hand, FA poles appear in
the same alignment of the SS poles. This phenomenon is clearly visible in Figure 7.36, where the
clusters are compared with the wind speed. Apparently, after a certain point of operation (Figure 7.37),
the mode vibrating in the SS direction tends to change its main direction of vibration to FA (although
not perfectly aligned to the FA direction as its FA clusters pair). Looking at Figures 7.13 and 7.14, it
can be attested that this point of operation coincides with the start of the blades pitch actuation.
283
Structural Monitoring of Wind Turbines
2.95 2.95
FA FA
SS SS
2.9 2.9
Frequency [Hz]
Frequency [Hz]
2.85 2.85
2.8 2.8
2.75 2.75
2.7 2.7
3 5 7 9 11 13 15 17 19 21 23 25 9 10 11 12 13 14 15 16 17
Wind speed [m/s] Rotor speed [rpm]
Figure 7.36 – Zoom of the clusters around 2.70 Hz and 2.95 Figure 7.37 – Zoom of Campbell diagram around 2.70 Hz
Hz vs. wind speed and 2.95 Hz
In addition, the other two referred alignments (1.30 Hz and 1.80 Hz) are also distinguished as SS
clusters in all operating conditions. It is interesting to note that the frequency value of these two modes
seems to decrease and increase, respectively, with the increase of the rotor speed (Figure 7.38). As
referred in section 4.1.2, this is an expected behaviour of rotor whirling modes when considering a
fixed reference (as is the case of the implemented monitoring system with sensors at the tower).
Nonetheless, this figure also evidences that, with the start of the blades pitch actuation, both modes
tend to decrease their natural frequency, changing their behaviour. It is also interesting to note that
these two modes do not seem to be affected by the rotor harmonics, even when their frequency value is
close (or coincident) with the one from the harmonic.
1.9 1.1
FA FA
1.8 SS 1.05 SS
1
1.7
Frequency [Hz]
Frequency [Hz]
0.95
1.6
0.9
1.5
0.85
1.4
0.8
1.3 0.75
1.2 0.7
9 10 11 12 13 14 15 16 17 0 2 4 6 8 10 12 14 16 17
Rotor speed [rpm] RPM
Figure 7.38 – Zoom of Campbell diagram around 1.20 Hz Figure 7.39 – Zoom of Campbell diagram around 0.70 Hz
and 1.90 Hz and 1.10 Hz
The alignment of clusters around 3.70 Hz is composed by FA and SS clusters (although it is not clearly
visible in Figure 7.35). A detailed analysis of this alignment shows that it is referred to only one mode,
since the FA and SS cluster do not appear at the same time. This phenomenon is analysed in more
detail in the next section.
The analysis of Figure 7.35 is also important to assess the influence of the rotor excitation on the
detection of the vibration modes throughout the operating regimes. It is observed that some stable
284
Chapter 7
horizontal alignments of cluster are only visible until the start of the turbine operation. The alignment
of FA clusters around 0.80 Hz is an example of this phenomenon (Figure 7.39). This alignment may be
related to vibration modes whose influence on the tower motion is too low to be noticed when the
rotor starts spinning. Another example of the harmonic effect is illustrated with the horizontal
alignment of FA clusters around 2.30 Hz. From the rotor speed of 8.7 rpm, it is noticed that this
alignment tend to fluctuate according to the harmonics that cross it.
Considering this initial analysis, 9 modes were decided to be followed in the scope of the dynamic
monitoring system. This decision was made based on the importance of the mode to the wind turbine
motion and on influence of the harmonics over the mode.
A nomenclature is also given for the monitored modes. It is referred to the number of the order of the
mode and to the main direction of vibration. In addition, when the evolution of the identified
frequency value of a mode is clearly dependent on the rotor speed, an “*” is added to the name since it
may be referred to a rotor mode. The vibration mode with frequency value around 3.70 Hz is labelled
as 3 FA/ 3 SS, since, as referred, it seems to vibrate either way.
Considering the results obtained with this preliminary analysis, 6 operating regimes were adopted to
define the reference properties of the monitored vibration modes. For each regime, reference values of
frequency value and mode shape of the 9 considered modes were kept. These regimes are characterized
in Table 7.4.
Table 7.4 – Regimes considered for reference modal properties of the vibration modes
1 Parked or idling (with high pitch angle values higher than 72º)
Regime 1 is referred to setups when the pitch angle of the blades is at its maximum level. As
introduced in Figure 7.14, under certain non-operating conditions, the pitch angle of the blades is set
to values around 90º. Regime 2 is also referred to non-operating situations, but the pitch angle is
defined to values of around 70º.
The region defined as a transition state in Figure 7.35, related to setups where the turbine starts
operating, is named as Regime 3. Since the SCADA system records 10 min. mean values, the rotor
speed values of this regime are between 0 and the lowest operating rotor speed (8.7 rpm).
Regime 4 is regarded to setups in which the rotor speed is higher than 8.7 rpm and the pitch angle is
kept at its lower level, i.e., before the pitch actuator is actively varying the angle of attack of the blades.
285
Structural Monitoring of Wind Turbines
Once the pitch angle starts increasing in order to avoid excessive rotor speed values, it is considered as
Regime 5. This Regime extends to the limit operating conditions of the turbine.
Lastly, Regime 6 is referred to situations when the wind speed is higher than the turbine cut-out wind
speed. However, this regime was not identified during the monitoring period.
With the aim of illustrating the 9 considered vibration modes, the reference mode shapes from Regime
4 are presented Figure 7.40. At the bottom of each mode shape, a polar plot shows the magnitude and
phase angle of the modal amplitude of the sensors orientated to the main direction of vibration. It can
be seen that some are not perfect real modes, since some amplitudes present phase angles slightly
different from 0º or 180º.
From the illustrations of Figure 7.40, it is possible to conclude that the two pairs of closely-spaced
modes around 0.35 Hz (modes 1 SS and 1 FA) and 2.80 Hz (2 FA and 2 SS) are, respectively, the first
and second pairs of tower bending modes, which is in accordance with the results of the numerical
models described in section 7.3.2 (Oliveira, Magalhães et al., 2013).
The remaining modes are certainly related to motion from other structural elements besides the tower,
such as the rotor blades. This fact is in line with the behaviour observed in the Campbell diagram for
the modes 1 SS*, 2 SS* and 3 SS*.
70 70 70
60 60 60
50 50 50
40 40 40
30 30 30
20 20 20
10 10 10
0 0 0
-1.2 0 1.2 -1.2 0 1.2 -1.2 0 1.2
90
90 120 90
120 60 1 60 120
1 60
150 30
150 30 150 30
286
Chapter 7
70 70 70
60 60 60
50 50 50
40 40 40
30 30 30
20 20 20
10 10 10
0 0 0
-1.2 0 1.2 -1.2 0 1.2 -1.2 0 1.2
90 90 90
120 60 120 120
1 1 60 1 60
150 30 150 30 150 30
70 70 70
60 60 60
50 50 50
40 40 40
30 30 30
20 20 20
10 10 10
0 0 0
-1.2 0 1.2 -1.2 0 1.2 -1.2 0 1.2
90 90 90
120 120 60
1 60 1 120 1 60
150 30 150 30 150 30
287
Structural Monitoring of Wind Turbines
Frequency Δ ≤1%
Damping Δ ≤5%
From the obtained clusters, only the ones with a number of poles higher than 6 were considered. Then,
for each reference mode, the clusters presenting an average natural frequency that did not differ more
than a predefined percentage value (10 % to 20 % depending on the mode type) from the reference
natural frequency value were selected. From those, it is selected the one that presents the average mode
shape with the highest correlation with the reference mode shape (evaluated with the MAC
coefficient). Modes with MACs lower than 0.80 are not considered.
The stated parameters were tried on several initial datasets and proved to deliver good and coherent
results. Figure 7.41 shows the Campbell diagram after the tracking (comparison of the cluster
properties with the reference properties of the vibration modes intended to be monitored). It can be
seen that, in this particular application, this simple tracking procedure was adequate to eliminate the
influence of the harmonics in the modes under analysis. For that reason, it was not necessary to use the
methods presented in section 5.6.3.
288
Chapter 7
Campbel Diagram
4.5
4 FA
4 3 SS*
3.5 3 FA/SS
3 2 SS
2 2 SS*
1.5
1 1 SS*
0.5 1 SS
1 FA
0
0 2 4 6 8 10 12 14 16 18
RPM
Figure 7.41 – Campbell diagram with the tracked vibration modes (p-LSCF algorithm)
Naturally, it was also necessary to define the input parameters for each modal identification algorithm.
Initial datasets were used to tune these parameters.
The SSI-COV algorithm requires two input parameters: the number of points of the correlation
function and the maximum order of the model. It was decided to adopt correlation functions of 128
points and a maximum order of 70.
Table 7.6 presents the main statistics related to the performance of this algorithm in the automated
identification of the 9 considered vibration modes. A success rate above 60 % is consistently obtained,
except for the 3 FA/3 SS, 3 SS* and 4 FA modes. Also, the 1 FA mode shows a considerably lower
success rate than its pair. This fact is probably due to the high values of damping of this mode which
hinders the identification of this mode by the algorithms. For the 3 FA/3 SS mode, the low success rate
is explained by the fact that this mode presents an unusual behaviour, which tends to change its main
direction of vibration according to the nacelle orientation. Thus, since a minimum MAC value for the
modal tracking was defined, only a small portion of the clusters identified in the frequency range of
this mode were considered. On the other hand, the low value obtained for the identification of the 3
SS* and 4 FA modes are mainly related to non-operating conditions (Regimes 1 and 2) and to the
Regime 5. In these three regimes, the success rate was considerably low.
From the analysis of Table 7.6, it is interesting to check that within the first pair of tower bending
modes, the 1 SS mode presents a much smaller dispersion of its natural frequency than the 1 FA mode.
Again, the high values of damping of the 1 FA mode may have influenced the quality of the detection
of this mode.
289
Structural Monitoring of Wind Turbines
Table 7.6 – Statistics related to the results obtained with the SSI-COV algorithm
Success Rate
Mode [Hz] [Hz] . .
[%]
1 SS 81.2 0.354 0.002 0.005
1 FA 67.4 0.355 0.007 0.020
1 SS* 63.4 1.325 0.051 0.039
2 SS* 78.3 1.791 0.060 0.034
2 FA 75.8 2.792 0.024 0.009
2 SS 83.7 2.828 0.030 0.011
3 FA/ 3SS 9.1 3.673 0.042 0.011
3 SS* 29.5 3.782 0.043 0.011
4 FA 38.3 4.236 0.072 0.017
The second time domain identification algorithm used was the SSI-DATA. For this algorithm, two
parameters were defined: the number of blocks of the Hankel matrix and the maximum order of the
model. After initial tuning, it was concluded that = 70 blocks, together with a maximum model order
of 70 led to the best results.
The main statistics about the automated modal identification with the SSI-DATA are introduced in
Table 7.7. The results are in line with the ones obtained with the SSI-COV. On the identification of the
4 FA mode, a considerably increase was obtained.
Table 7.7 – Statistics related to the results obtained with the SSI-Data algorithm
Success Rate
Mode [Hz] [Hz] . .
[%]
1 SS 84.7 0.355 0.002 0.006
1 FA 67.4 0.358 0.008 0.022
1 SS* 84.9 1.330 0.050 0.038
2 SS* 89.8 1.793 0.057 0.032
2 FA 74.0 2.789 0.024 0.009
2 SS 60.2 2.829 0.030 0.011
3 FA/ 3SS 14.9 3.690 0.039 0.011
3 SS* 44.9 3.789 0.047 0.012
4 FA 82.3 4.266 0.079 0.019
The last identification algorithm used in the dynamic monitoring system was the p-LSCF. Positive
time lags of the correlation with 1024 points were used to calculate the spectra. An exponential
window with a factor of 0.1 was also used. A maximum model order of 40 was considered. The
obtained results show that this algorithm provided much higher success rates for the modes not so
successfully identified with the other two algorithms.
290
Chapter 7
Table 7.8 – Statistics related to the results obtained with the p-LSCF algorithm
Success Rate
Mode [Hz] [Hz] . .
[%]
1 SS 80.8 0.354 0.002 0.006
1 FA 64.7 0.354 0.007 0.019
1 SS* 87.4 1.330 0.049 0.037
2 SS* 85.5 1.793 0.058 0.032
2 FA 80.1 2.796 0.017 0.006
2 SS 85.5 2.831 0.026 0.009
3 FA/ 3SS 43.2 3.700 0.030 0.008
3 SS* 67.8 3.809 0.058 0.015
4 FA 95.1 4.276 0.081 0.019
Once the modal parameters are identified, it is possible to analyse the variation of the parameters with
the operational and environmental factors. Figure 7.42 characterizes the evolution of the natural
frequencies of the 9 monitored vibration modes along the monitoring period. The variability of the
frequency values is evident.
4.5
3.5
3
Frequency [Hz]
2.5
1.5
0.5
Time
Figure 7.43 a) shows a zoom of the frequency variation of 1 FA and 1 SS vibration modes along the
recorded setups. It is visible that the variability of the 1 FA mode is much higher than of the 1 SS,
which is in agreement with the results presented in Tables 7.6 to 7.8. Figure 7.43 b) evidences the
variation of the natural frequencies of the 1 SS* and 2 SS* modes, mainly due to their dependence on
the rotor speed.
291
Structural Monitoring of Wind Turbines
0.4
1.8
0.38
1.7
Frequency [Hz]
Frequency [Hz]
1.6
0.36
1.5
0.34
1.4
1.3
0.32
1.2
0 0.5 1 1.5 2 0 0.5 1 1.5 2
Index 4 Index 4
x 10 x 10
a) b)
Figure 7.43 – Zoom of the natural frequencies along the recorded setups: a) 1 FA and 1 SS; b) 1 SS* and 2 SS*
In order to better understand the behaviour of the modal parameters of the 9 tracked modes, Figure
7.44 shows the variation of the natural frequencies for each operating regime. This graphic evidences
the suitability of defining different Regimes for modal tracking. For example, the variation of the
frequency values of the modes 1 SS* and 2 SS* is notorious. Apart from these modes, also the 2nd pair
of tower bending modes shows a different behaviour when the turbine starts operating, with a better
separation of the two modes.
4.5
3.5
3
Frequency [Hz]
2.5
1.5
0.5
0
Regime Regime Regime Regime Regime
1 2 3 4 5
Figure 7.44 – Variation of the natural frequency, for each operating regime, of the monitored vibration modes
Figure 7.45 a) presents a closer look of the first pair of tower bending modes. It is interesting to note
that the variability of the 1 FA mode is smaller than the 1 SS mode for non-operating conditions. Once
the turbine starts operating, the dispersion of the frequency values from this mode greatly increases.
On the other hand, the variability of the 1 SS mode is kept almost constant throughout the different
operating regimes. These facts reinforce the idea that the high values of damping of the 1 FA mode,
during operating conditions, may have influenced the accuracy of the detection of this mode. Figure
7.45 b) evidences the better separation of the 2nd pair of tower bending modes under operating
conditions.
292
Chapter 7
0.4 2.95
2.9
0.38
2.85
Frequency [Hz]
Frequency [Hz]
0.36
2.8
0.34 2.75
2.7
0.32
2.65
Reg. Reg. Reg. Reg. Reg. Reg. Reg. Reg. Reg. Reg.
1 2 3 4 5 1 2 3 4 5
a) b)
Figure 7.45 – Variation of the natural frequency, for each operating regime, of: a) 1 FA and 1 SS; 2 FA and 2 SS
The results obtained with the identification of the natural frequencies and damping ratios of 9
monitored vibration modes with the p-LSCF algorithm are shown in Tables 7.9 and 7.10 for each
regime.
Table 7.9 – Results obtained for each operating regime: natural frequencies
1 2 3 4 5
Mode
[Hz] [Hz] [Hz] [Hz] [Hz] [Hz] [Hz] [Hz] [Hz] [Hz]
1 SS 0.353 0.003 0.353 0.003 0.354 0.002 0.355 0.001 0.355 0.002
1 FA 0.352 0.001 0.352 0.001 0.355 0.004 0.357 0.006 0.349 0.006
1 SS* 1.258 0.011 1.277 0.011 1.400 0.020 1.362 0.033 1.294 0.018
2 SS* 1.625 0.011 1.812 0.017 1.729 0.025 1.793 0.045 1.844 0.013
2 FA 2.767 0.012 2.784 0.013 2.781 0.022 2.799 0.015 2.796 0.014
2 SS 2.792 0.020 2.790 0.021 2.819 0.022 2.842 0.016 2.841 0.012
3 FA/
- - 3.659 0.042 3.664 0.038 3.702 0.028 3.702 0.026
3SS
3 SS* 3.668 0.050 3.683 0.055 3.748 0.053 3.812 0.036 3.870 0.018
4 FA 4.107 0.033 4.167 0.046 4.253 0.068 4.301 0.040 4.349 0.048
293
Structural Monitoring of Wind Turbines
Table 7.10 – Results obtained for each operating regime: damping ratios
1 2 3 4 5
ߦ ߦ௦௧ௗ ߦ ߦ௦௧ௗ ߦ ߦ௦௧ௗ ߦ ߦ௦௧ௗ ߦ ߦ௦௧ௗ
Mode
[%] [Hz] [%] [Hz] [%] [Hz] [%] [Hz] [%] [Hz]
1 SS 0.974 0.631 0.733 0.438 0.707 0.590 0.630 0.394 1.096 0.548
1 FA 0.351 0.482 0.213 0.356 1.490 1.128 3.659 1.390 7.868 2.169
1 SS* 0.502 0.207 0.646 0.316 0.845 0.232 0.932 0.236 0.860 0.152
2 SS* 0.543 0.261 0.746 0.297 0.746 0.271 0.916 0.247 1.113 0.211
2 FA 0.175 0.100 0.195 0.120 0.624 0.231 0.884 0.320 1.291 0.284
2 SS 0.301 0.127 0.254 0.167 0.344 0.162 0.451 0.225 0.662 0.181
3 FA/
- - 0.778 0.275 1.013 0.350 1.395 0.427 1.820 0.407
3SS
3 SS* 0.572 0.230 0.551 0.228 0.660 0.224 0.719 0.149 0.888 0.232
4 FA 0.476 0.170 0.457 0.178 1.079 0.433 1.598 0.374 1.566 0.467
The evolution of the 1 SS* and 2 SS* modes with the rotor speed is illustrated in Figure 7.46. This
figure resembles the typical behaviour the of rotor whirling modes. It is visible that, when the turbine
is consistently rotating (for rotor speed higher than 8.7 rpm), the 1 SS* and 2 SS* modes have a similar
behaviour with, respectively, a backward and a forward whirling mode. Considering that, when in
operation, these modes are detected by tower motion in the SS direction, i. e., in the direction of the
rotor plane, these two modes are most likely related to edgewise vibration modes. From the three
configurations of first order edgewise modes shown in Figure 4.4, only asymmetrical modes create a
reaction force at the rotor and, thus, are capable of inducing a detectable tower motion. For these
reasons, it expected that 1 SS* and 2 SS* modes are in fact the first backward and forward edgewise
modes, respectively.
A regression model for each vibration mode is also presented in Figure 7.46 in dashed lines. These
regression models were adjusted to the frequency values obtained for the two modes during events
regarding Regime 4. Regime 5 was not considered for the definition of the models because this regime
implies the variation of the blades pitch angle which introduces changes in the behaviour. The
intercept of the regression models with the origin (rpm=0) gives an approximation of the frequency
values of these two modes under parked conditions and with a pitch angle around 0º. Values of 1.526
Hz and 1.558 Hz were obtained for, respectively, the backward and forward mode. However, this
configuration does not occur in the monitored wind turbine. When under parked conditions, the
blades are usually oriented with a pitch angle between 72º and 90º, changing the direction of vibration
of the edgewise modes from in-plane rotor motion (when in operation) to out-of-plane (when in non-
operation conditions). Consequently, the support conditions of the blades are also changed with this
transition. Due to these reasons, modes vibrating at the referred frequencies were not detected. It was
then investigated if, during events from Regime 1 and 2, there were modes with natural frequencies
next to these frequencies and with a tower mode shape similar to the ones from the 1 SS* and 2 SS*
modes but vibrating in the FA direction (as referred, during non-operating conditions, the edgewise
modes vibrate out of the plane rotor). The cluster referred to the found vibration modes during
Regime 1 and 2 are shown in Figure 7.46 in lighter colours. It is seen that, while for the 1 SS* mode, the
identified modes have natural frequencies close to each other during Regime 1 and 2 (݂ ≈ 1.26 Hz), the
same does not happen for the 2 SS* (݂ = 1.63 Hz and ݂ = 1.81 Hz for, respectively, Regime 1 and 2). In
294
Chapter 7
fact, the identified clusters for the 2 SS* mode present a considerably change in the natural frequency
between Regime 1 and 2. Unfortunately, this situation could only be deeper investigated if
instrumentation at the blades would be available. Even though, the referred clusters for Regime 1 and 2
were considered for monitoring purposes.
2
2 SS*
1.6
1.4
1.2 1 SS*
1
0 2 4 6 8 10 12 14 16
Rotor speed [rpm]
Figure 7.46 – Evolution of the frequency values of the 1 SS* and 2 SS* with the rotor speed
Likewise, although on a smaller scale, the natural frequencies of the second pair of tower bending
modes (2 FA and 2 SS) also shows an increasing trend with the operating regimes. On the other hand,
the 1 FA and 2 SS modes do not have a very distinct trend over the regimes, as shown in Figure 7.47
0.39 0.37
0.38
0.365
0.37
0.36
Frequency [Hz]
Frequency [Hz]
0.36
0.35 0.355
0.34
0.35
0.33
0.345
0.32
0.31 0.34
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
Rotor speed [RPM] Rotor speed [rpm]
0.39 0.37
0.38
0.365
0.37
0.36
Frequency [Hz]
Frequency [Hz]
0.36
0.35 0.355
0.34
0.35
0.33
0.345
0.32
0.31 0.34
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 1 2 3 4 5 6 7 8 9 11 13 15 17 19 21
Wind speed [m/s] Wind speed [m/s]
a) b)
Figure 7.47 – Median (and box plots) of the frequency values: a) 1 FA mode; b) 1 SS mode
295
Structural Monitoring of Wind Turbines
Globally, the damping ratio of the FA modes increases with the increase of the wind speed (which
usually implies a higher rotor speed). This is mostly due to the contribution of the aerodynamic
component of damping. As expected, this increase of the damping values is not notorious in the SS
modes due to the lower opposition to the wind flow in this direction. It is however noticed that the
damping of SS modes increases in Regime 5. This is probably due to the aerodynamic change
introduced by the increase of the blades pitch angle.
The analysis of the evolution of the damping of the two pairs of bending modes illustrates the complex
dynamics of a wind turbine. For non-operating conditions (Regimes 1 and 2), the damping values of
the SS modes are higher than for the FA modes. This situation is due to the deviation of the rotor from
the main wind direction and to the high pitch value of the blades. Under these conditions, the blades
present a larger opposition to the wind flow in the SS direction, leading to the appearance of
aerodynamic damping in this direction.
Once the turbine starts operating, the rotor is orientated to the main wind direction and the blades
pitch angle is set to its minimum value. These conditions, together with the higher wind speed, lead to
the increase of the aerodynamic damping in the FA modes and the almost vanishing of this
contribution in the SS modes. Since the aerodynamic damping component increases with the increase
of the wind speed, the damping values of the 1 FA and 2 FA modes are consistently growing over the
operating regimes. This effect is more noticeable for modes with large modal amplitude at the tower
top, where the wind force is higher. For this reason, the 1 FA mode presents higher values than the 2
FA (whose modal amplitude at the top is very small).
The top plots of Figure 7.48 illustrate the evolution of the damping ratio of the 1 FA and 1 SS vibration
modes with the rotor speed. The increase of the damping values of the 1 FA mode when the turbine
starts operating is evident. It is also noted that the damping increases when reaching the final part of
the rotor speed range. Notwithstanding, these figures only shows the evolution of damping until the
wind rated speed is achieved (around 13 m/s). For that reason, the lower graphic of Figure 7.48
presents the evolution of the damping of the 1 FA mode with the wind speed. The shape of the
adjusted line clearly shows a drop in the damping value around 15 – 17 m/s. This situation is
consequence of the increase of the pitch angle in Regime 5 which reduces the thrust force at the tower
top and, consequently, reduces the rotor blades opposition to the wind flow in the FA direction.
On the other hand, the adjusted line for the damping values of the 1 SS mode only shows small
variations over the different operating conditions, with a reduction when the turbine starts operating
and an increasing when reaching its rated speed.
296
Chapter 7
15 7
12.5 6
5
10
Damping [%]
Damping [%]
4
7.5
3
5
2
2.5 1
0 0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
Rotor speed [m/s] Rotor speed [RPM]
15 7
12.5 6
5
10
Damping [%]
Damping [%]
4
7.5
3
5
2
2.5 1
0 0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21
Wind speed [m/s] Wind speed [m/s]
a) b)
Figure 7.48 – Median (and box plots) of the damping ratios: a) 1 FA mode; b) 1 SS mode
The situation illustrated in Figures 7.36 and 7.37, about the possible change of mode shape orientation
of the 2 SS mode for high wind speed conditions, was further investigated during the modal tracking.
In that sense, an additional reference “mode” named 2 SS(/FA), based on the 2 SS mode but with an
orientation between the FA and the SS direction, was considered. As results, it was possible to confirm
that this “mode” was identified on setups in which the 2 SS was missing. Considering the number of
setups in which the 2 SS(/FA) “mode” was tracked, the success rate of the 2 SS increases from 67.1 % to
85.5 % (with the p-LSCF algorithm).
2.9 2.9
2 FA 2 FA
2 SS 2 SS
2 SS(/FA)
2.85 2.85
Frequency [Hz]
Frequency [Hz]
2.8 2.8
2.75 2.75
5 10 15 20 25 5 10 15 20 25
Wind speed [m/s] Wind speed [m/s]
a) b)
Figure 7.49 – a) Evolution of the 2 FA and 2 SS modes with the wind speed. b) Evolution of the 2 FA, 2 SS and the 2 SS/(FA)
297
Structural Monitoring of Wind Turbines
Acc. [m/s2]
0 0
-0.05
-0.1
-0.1
-0.2
-0.15
-0.2 -0.3
0 100 200 300 400 500 600 0 100 200 300 400 500 600
Time [s] Time [s]
a) b)
Figure 7.50 – Events in which a shutdown was detected: a) structural response following an exponential decay; b)
example of structural response not following the common exponential decay
298
Chapter 7
From the initially pre-selected events, 51 were chosen to be analysed. The majority of the events are
referred to situations in which the wind speed dropped to values below the cut-in. In addition, 7 events
were detected whose measured maximum wind speed was above the rated speed and the minimum
wind speed as also above the cut-in value. However, in none of these events, the maximum wind speed
is above the cut-out value. Figure 7.51 illustrates the minimum and maximum wind speed measured
by the SCADA system for the selected events.
25 Cut-out
Minimum
Maximum
20
Wind speed [m/s]
15
Rated
10
5
Cut-in
0
0 5 10 15 20 25 30 35 40 45 50
Index
Figure 7.51 – Measured maximum and minimum wind speed occurred in the selected 10 min. events containing a free
decay response
Figure 7.52 illustrates the diversity of the acceleration levels observed within the selected free decay
responses. Figure 7.52 a) shows the decay response from an event whose maximum and minimum
wind speed was, respectively, 19.4 m/s and 10.1 m/s. On the other hand, Figure 7.52 b) presents the
acceleration time series of a decay response from a period in which the wind speed varied between 4.2
m/s and 1.0 m/s, which means that the turbine stopped because the wind speed was too low for power
production. Comparing both figures, the difference in the level of acceleration is notorious.
1.5 0.08
+74.988 FA +74.988 FA
+48.392 FA
0.06
1 +48.392 FA
+21.772 FA +21.772 FA
0.04
0.5
0.02
Acc. [m/s2]
Acc. [m/s2]
0 0
-0.02
-0.5
-0.04
-1
-0.06
-1.5 -0.08
0 100 200 300 400 500 600 0 100 200 300 400 500 600
Time [s] Time [s]
a) b)
Figure 7.52 – Free decay response from: a) a high wind speed event; b) a low wind speed event
During the free decay events, the motion of the tower is mainly in the FA direction. Figure 7.53 shows
the evolution of the FA and SS acceleration at the tower top of the decay responses represented in
Figure 7.52 for the time segments enclosed by the vertical dashed lines. In this figure, corresponding to
299
Structural Monitoring of Wind Turbines
a top view of the tower motion, the hotter colours correspond to the beginning of the time segment
while the cold colours correspond to the final instants of the analysed period.
+74.988 m +74.988 m
1.5 0.05
1
SS acceleration [m/s2]
SS acceleration [m/s2]
0.025
0.5
0 0
-0.5
-0.025
-1
-1.5 -0.05
-1.5 -1 -0.5 0 0.5 1 1.5 -0.05 -0.025 0 0.025 0.05
FA acceleration [m/s2] FA acceleration [m/s2]
a) b)
Figure 7.53 – Evolution of the FA and SS acceleration (top view) from: a) a high wind speed event; b) a low wind speed
event (the hotter colors correspond to the beginning of the time segment while the cold colors are referred to the final
instants)
Initially, the two events introduced in Figure 7.52 were analysed by fitting an exponential function to
the free decay envelope, in the same manner as in section 5.8.4.2 for the rotor-stop tests performed for
the Izar Bonus 1.3MW/62 wind turbine. A low-pass filter with a frequency cut of 0.5 Hz was applied to
both signals. The exponential function was then fitted to the local maxima points of the filtered signals,
as illustrated in Figure 7.54. The results obtained for the frequency value and damping ratio of the 1
FA mode are presented in Table 7.11.
+74.988 m +74.988 m
1.5 0.08
1.203*e-0.0058*2*π*0.349*t 0.047*e-0.0060*2*π*0.352*t
0.06
1
0.04
0.5
0.02
Acc. [m/s 2 ]
Acc. [m/s 2 ]
0 0
-0.02
-0.5
-0.04
-1
-0.06
-1.5 -0.08
150 200 250 300 350 400 200 250 300 350 400
Time [s] Time [s]
a) b)
Figure 7.54 – Identification of the modal damping ratio of the 1 FA vibration mode through filtering and fitting of an
exponential function to the envelope of the free decay response from: a) a high wind speed event; b) a low wind speed
event
The two free decay responses were also analysed through the application of the SSI-COV algorithm
adapted to consider the measured free decays as input. This method was also used in section 5.8.4.2 to
study the structural response of the Izar Bonus 1.3MW/62 wind turbine to the rotor-stop test. The
300
Chapter 7
time segments defined to fit the exponential function in Figure 7.54 were used as inputs of the
algorithm. The obtained stabilization diagrams are shown in Figure 7.55. Besides the one referred to
the 1 FA mode, other vertical stable alignments are also evident. The results obtained for the 1 FA
mode are summarized in Table 7.11.
Stabilization diagram Stabilization diagram
20 20
15 15
Model Order
Model Order
10 10
5 5
0 0
0 1 2 3 4 5 0 1 2 3 4 5
Frequency [Hz] Frequency [Hz]
a) b)
Figure 7.55 – Stabilization diagram obtained with the application of the SSI-COV algorithm
The results obtained with the two procedures presented very similar values for frequency and damping
ratio. It is also visible that there is no apparent influence of the wind speed on the damping values of
the mode. It should be noted that in both free decays, due to change of orientation of the blades (from
low angles of pitch to values to close to 90º), the contribution of the aerodynamic damping for the FA
mode is not relevant.
Table 7.11 – Natural frequency ( ) and modal damping ratio ( ) of the 1 FA vibration mode identified with the filtering
and fitting procedure and the SSI-COV method
The described analysis was then extended to all the selected free decay events. For the procedure based
on filtering and fitting, a routine was developed to find the best exponential function to fit the decay
response of the wind turbine. The results obtained were then visually analysed in order to validate the
results. The obtained results are presented in Figure 7.56.
The segment of the acceleration time series used for the filtering and fitting procedure was also
considered as input for the SSI-COV algorithm. In order to automate the procedure based on this
method, a cluster analysis was performed after each run of the algorithm. With this processing
solution, the procedure becomes similar to the one described for the continuous modal identification
in section 7.3.6.1. The results obtained are also shown in Figure 7.56.
301
Structural Monitoring of Wind Turbines
1st FA mode
0.356
filt./ fitting
0.355
SSI-COV
0.354
0.352
0.351
0.35
0.349
0.348
0 5 10 15 20 25 30 35 40 45 50
Index
1st FA mode
1
filt./ fitting
0.9 SSI-COV
0.8
Damping ratio [%]
0.7
0.6
0.5
0.4
0.3
0 5 10 15 20 25 30 35 40 45 50
Index
Figure 7.56 – Results obtained for the 1st FA mode with the filtering and fitting procedure and SSI-COV algorithm for
frequency value (top plot) and damping ratio (bottom plot)
The results obtained show a good agreement between the two methods. Some variability in the results
is noticed though. It was thus decided to assess if the operational and environmental parameters could
explain this variance. Figure 7.57 illustrates the evolution of the modal properties identified with the
SSI-COV methodology with maximum absolute values (MAX) of the selected segments of the
acceleration decay responses. From these figures, it is visible that the majority of the decays is related
to events with values smaller than 0.4 m/s2 which difficulties the identification of a clear trend.
Notwithstanding, there is an apparent decrease of the frequency values of the 1 FA mode with the
acceleration. However, the number of events with MAX values higher than 0.4 is too low to confirm
this trend.
On the other hand, the identified values of modal damping do not present a distinct dependency on
the acceleration level. This conclusion is contrary to what usual occurs in common civil structure,
where the damping tends to increase with the acceleration level.
Theoretically, the study of the free decay events could allow the study of the structural damping of the
wind turbine. However, comparing the damping values identified with the free decay events with the
results from the automated modal tracking, it is visible that they are considerably higher than the ones
from the non-operating Regime 1 (in which the blades orientation is equal to the observed in the free
decay events after the impulse) and lower than the ones observed in Regime 3 (in which the wind
conditions are similar to the free decay events).
302
Chapter 7
0.355
0.9
0.354
Frequency [Hz]
0.8
Damping [%]
0.353
0.352 0.7
0.351
0.6
0.35
0.5
0.349
0.348 0.4
0 0.2 0.4 0.6 0.8 1 1.2 0 0.2 0.4 0.6 0.8 1 1.2 1.4
Acc. MAX [m/s ] 2
Acc. MAX [m/s2]
a) b)
Figure 7.57 – Evolution of the frequency value (a)) and damping ratio (b)) of the 1st FA mode with the maximum absolute
values of acceleration
Although the study of the frequency values indicates some possible relations with the vibration level, it
is not sufficient to explain the variation detected along the 51 decay events. This variation is mostly
related to the temperature dependency of this modal property. This relation is shown in Figure 7.59,
where a clear trend is visible. In this figure, some points (enclosed in a red circle) are visible with
frequency values lower than what the trend suggests. These points are referred to events with high
acceleration amplitudes, suggesting that, in fact, the vibration level may interfere in the frequency
values of this mode.
With regard to the damping, it cannot be identified the source of the observed scatter.
1st FA mode
0.356
0.355
0.354
Frequency [Hz]
0.353
0.352
0.351
0.35
Acc. RMS > 0.20
0.349
0.348
5 10 15 20 25 30
Temperature [ºC]
Figure 7.58 – Evolution of the frequency value of the 1st FA mode with temperature
7.3.6.4 Participation of the Vibration Modes/ Harmonics to the Measured Dynamic Response
The methodology presented in section 5.5 to decompose the measured acceleration time series into
several modal and stationary responses (due to harmonics) was implemented in the developed
dynamic monitoring system.
In the application of the modal response estimation, it is required the selection of a representative
model order. The state-space matrices computed for that model order will then be used to define the
303
Structural Monitoring of Wind Turbines
forward innovation model. While in the example of section 5.5, a manual selection was performed
based on the observation of the stabilization diagram, an automated strategy for choosing the best
model order is required as part of a monitoring system. The implemented strategy is thus based on the
selection of the order with the highest number of stable poles whose clusters were chosen during the
tracking procedure. When the turbine is operating (Regimes 4 and 5), cluster referred to the
harmonics were also considered in the selection of the model order. For a cluster to be considered as
representative of the stationary response due the harmonic excitation, its frequency value has to be
within a range of 10 % of the multiples of the mean rotor speed from the setup under analysis. In
cases in which there is more than one cluster meeting this condition, the cluster containing the higher
number of poles is chosen. If this cluster has more poles than a pre-defined minimum number, it is
defined as representative of the wind turbine response due to the harmonic excitation. In this
application, a minimum number of 6 poles was defined. The 1Ω, 3Ω and 6Ω harmonics were
considered in this analysis, although the first was only identified in 5.6 % of the operating setups.
Figure 7.59 characterizes the RMS values of the modal contributions associated with the modes
vibrating in the FA direction, alongside with the harmonics, at the +74.988 m level. It is visible that the
dynamic motion is dominated mainly by the 3Ω harmonic and the 1 FA mode. As expected, the
acceleration response in 1Ω harmonic frequency represents a very low level. It is also interesting to
note that the 1 FA mode (and also the 6Ω harmonic) present a different behaviour from the other
responses, do not continuously increasing with the increase of the rotor speed (or wind speed). It is
observable that, when the turbine starts operating, there is a steep increase on the values from the 1 FA
mode, which tend to disappear when the turbine reaches the 10 rpm. This increase in the modal
acceleration response of the 1 FA is probably due to the excitation of this mode by the 3Ω harmonic.
Considering Figure 7.35, it is possible to attest that the 3Ω harmonic is close to the 1 FA mode at rotor
speeds between 8.7 and 10 rpm which can lead to the excitation of this mode and, consequently, to the
increase of the vibration amplitudes. A similar situation seems to occur with 6Ω harmonic response.
For wind speed around 3 m/s (corresponding to the start of the turbine), an increase of the
acceleration levels is also visible in Figure 7.59 b). This situation is probably justified by the interaction
between this harmonic and the non-tracked FA mode with a frequency around 0.80 Hz (Figure 7.35).
Sensor: 1 (+74.988 m) +74.988 m
0.3 0.3
1Ω
0.25 3Ω
0.25
6Ω
0.2 0.2 1 FA
Acc. [m/s2]
Acc.[m/s 2]
2 FA
0.15 0.15 4 FA
0.1 0.1
0.05 0.05
0 0
0 2 4 6 8 10 12 14 16 17 0 5 10 15 20 25
Rotor speed [rpm] Wind speed [m/s]
a) b)
Figure 7.59 – RMS values from the modal and stationary responses in the FA direction at level +74.988 m with: a) rotor
speed; b) wind speed
304
Chapter 7
Once the recorded acceleration time series are decomposed into modal (and stationary) responses, it is
possible to use the procedure introduced in section 5.5.3 to quantify the relative contribution of each
responses to the measured signal. Likewise in the example from section 5.5.3, the acceleration
responses due to the harmonic excitation were also considered in this analysis. This analysis is helpful
to better understand the changes in the dynamic behaviour of the wind turbine structure over the
different conditions.
Figure 7.60 shows the evolution of the relative modal contribution of the FA modes and 1Ω, 3Ω and 6Ω
harmonics to the measured acceleration at the tower top sensor with the wind speed. The median
values (from each range of 0.5 m/s of wind speed) from the biggest contributors responses (1 FA
mode, 3Ω and 6Ω harmonics) are represented by the lines in light colour. For low wind speed
conditions (non-operating scenarios), the dynamic behaviour at the tower top is clearly dominated by
the 1 FA mode, with participations of almost 100 % in certain setups. Once the turbine starts
operating, the response is partially dominated by the 3Ω harmonic and by the 1 FA mode, whose
participation is still representative mainly due to the excitation of this mode by the 3Ω harmonic, as
previously explained. From the moment this effect vanishes, it is observed that the acceleration is
practically dominated by the 3Ω harmonic. From this point, the participation of this harmonic and the
1 FA mode tend to converge to a participation of 30 % each with the increase of the wind speed. The
participation of the 6Ω harmonic is only relevant when the turbine starts its operation, decreasing
from this point to participation levels below 5 % with the increase of the wind speed.
+74.988 m
1
1Ω
3Ω
0.75 6Ω
1 FA
Participation
2 FA
0.5 4 FA
0.25
0
0 5 10 15 20 25
Wind speed [m/s]
Figure 7.60 – Variation of the participation of the modal and stationary responses in the measured acceleration in the FA
direction at level +74.988 m with wind speed
The analysis presented to the sensor at the tower top was also performed at the +48.392 m level in the
FA direction. The results are presented in Figure 7.61. The acceleration, in the FA direction, is
dominated by the 2 FA mode with RMS values clearly higher than the ones presented in Figure 7.59
for the 1 FA mode. This result is probably justified by the excitation introduced by the blade rotation
(due to the tower shadow effect) at the level +48.392 m, corresponding to the maximum modal
amplitude of the 2 FA mode. Anyway, the dynamic displacements would be expected to be higher at
the tower top than at +48.392 m due to the lowest natural frequency of the 1 FA when compared to the
2 FA mode. This is confirmed later in section 7.3.7.
305
Structural Monitoring of Wind Turbines
+48.392 m +48.392 m
1 1
1Ω 1Ω
3Ω 3Ω
0.8 0.8 6Ω
6Ω
1 FA 1 FA
Acc.[m/s 2]
Acc.[m/s 2]
0.6 2 FA 0.6 2 FA
4 FA 4 FA
0.4 0.4
0.2 0.2
0 0
0 2 4 6 8 10 12 14 16 0 5 10 15 20 25
Rotor speed [rpm] Wind speed [m/s]
a) b)
Figure 7.61 – RMS values from the modal and stationary responses in the FA direction at level +48.392 m with: a) rotor
speed; b) wind speed
The quantification of the contribution of the modal and stationary responses to the measured
acceleration time series was also assessed. The results are shown in Figure 7.62. In this figure, besides
the modes considered in Figure 7.60, the stationary response due to the 9Ω harmonic was also
included. However, it is important to note that around a rotor speed of 14 rpm, this harmonic crosses
the potential untracked FA vibration mode, identified with a frequency of around 2.30 Hz (Figure
7.35). Since this potential mode was not considered during the tracking procedure, the automated
algorithm developed to identify clusters related to the harmonics cannot distinguish, for this region of
the Campbell diagram, a cluster representative of the 9Ω harmonic from one representative of this
mode. Thus, it is possible that clusters referred in Figure 7.62 as due to the 9Ω harmonic are actually
representative of this potential mode.
The results presented in Figure 7.62 clearly show that, under low wind speed conditions (non-
operating regimes), the acceleration is dominated by the 1 FA mode. Some sparse points from the 1 FA
with high modal contribution are also visible for high wind speeds but these are referred to setups with
no power production. When the rotor starts operating, the contribution of the 1 FA drops drastically,
keeping low for every wind speed condition. As expected, the acceleration is dominated by the 2 FA
mode while the remaining the modes and harmonics are kept low during all operating conditions.
Notwithstanding, it is observed that, for wind speeds around 7 and 13 m/s, the relative contribution of
the 2 FA decreases slightly. This region coincides with the increase of the contribution of the 9Ω
harmonic. This increase is due to the cross of this harmonic with the referred potential vibration mode
with frequency of 2.30 Hz, leading to a possible interaction or resonance of this mode. For wind speeds
higher than 13 m/s, the relative contribution of the 2 FA increases to levels similar to the ones obtained
to winds speed below 7 m/s and the contribution of the 9Ω harmonic drops, meaning that this effect
was reduced.
306
Chapter 7
+48.392 m
1
1Ω
3Ω
0.75 6Ω
9Ω
Participation
1 FA
0.5 2 FA
4 FA
0.25
0
0 5 10 15 20 25
Wind speed [m/s]
Figure 7.62 – Variation of the participation of the modal and stationary responses in the measured acceleration in the FA
direction at level +48.392 m with wind speed
307
Structural Monitoring of Wind Turbines
conditions. The operational and environmental factors recorded by the SCADA system and
considered in this analysis were:
o Wind speed (w. s.);
o Outdoor temperature (temp.);
o Rotor speed (r. s.);
o Blade pitch angle (pitch).
The operating regime 1 is characterized by values of blades pitch angle (around 90º) with small
variations, during non-production setups. Thus, the considered predictors influencing the most the
variability of the frequency values are the temperature and, in a small scale, the wind speed. These
expected results are confirmed by the obtained correlation coefficients (computed according to
equation (7.4)) and presented in Table 7.12. This coefficient permits to highlight the modes whose
natural frequencies are more (linearly) influenced by the factors under study. Figures 7.65 and 7.66
show two examples of the correlation between natural frequency and the wind speed and temperature
for two different modes, respectively.
Table 7.12 – Correlation coefficients between natural frequencies and operational/ environmental factors (Regime 1)
w. s. temp. r. s. pitch
1.3 2.85
1 SS* 2 FA
Linear fit 2.825 Linear fit
1.28
2.8
Frequency [Hz]
Frequency [Hz]
1.26
2.775
1.24
2.75
1.22
2.725
1.2 2.7
0 5 10 15 20 25 0 5 10 15 20 25 30 35
Wind speed [m/s] Temperature [ºC]
Figure 7.63 – Natural frequency of 1 SS* mode (Regime 1) Figure 7.64 – Natural frequency of 2 FA mode (Regime 1)
vs wind speed vs temperature
308
Chapter 7
Regime 2 is similar to Regime 1 to the extent that they are both referred to non-production setups, in
which the blades pitch angles is kept almost constant (although at lower values in Regime 2). Thus, it is
comprehensible that the predictors with the highest correlation with the natural frequencies are again
the wind speed and the temperature, as seen in the results of Table 7.13.
Table 7.13 – Correlation coefficients between natural frequencies and operational/ environmental factors (Regime 2)
w. s. temp. r. s. pitch
The correlation coefficients obtained between natural frequencies from Regime 3 and operational/
environmental factors are shown in Table 7.14. Operating regime 3 corresponds to transition situation
in which the turbine changes from parked/ idling configuration to operating conditions, or vice-versa.
For that reason, both the rotor speed and pitch angle show important changes during these setups.
However, apart from some modes (such as 1 SS*, 2 SS* and 4 FA modes), these changes do not seem to
have an important influence on the frequency values.
Table 7.14 – Correlation coefficients between natural frequencies and operational/ environmental factors (Regime 3)
w. s. temp. r. s. pitch
The number of identified setups regarded to Regime 4 is considerably higher than for the first 3
regimes. This fact permitted to obtain a better characterization of the variations of the natural
frequencies for this regime. The results obtain for the correlation analysis are presented in Table 7.15.
309
Structural Monitoring of Wind Turbines
Table 7.15 – Correlation coefficients between natural frequencies and operational/ environmental factors (Regime 4)
w. s. temp. r. s. pitch
Both the rotor speed and wind speed (which are themselves correlated) present interesting high values
of correlation with the frequencies. As expected, the pitch angle does not represent an interesting
source for explanation of the frequencies variability, since its value is practically constant during this
Regime. On the other hand, the temperature presents important linear relationship with the
frequencies, which are confirmed by visual observation of the results. The evolution of the frequency
values of the 2 FA mode with the temperature is illustrated in Figure 7.65.
The highest correlation values are obtained with the rotor speed. Due to the nature of the whirl modes,
this fact is especially important for the 1 SS* and 2 SS* modes. This phenomenon is illustrated for the 2
SS* mode in Figure 7.66.
Although slightly lower than with the rotor speed, the correlation coefficients obtained with the wind
speed are also relevant. However, this fact is mainly related to the relation between the wind speed and
the rotor speed than to physical relations between the wind and the natural frequencies.
3.95
2 FA 1.95
2 SS*
3.9 Linear fit
1.9 Linear fit
3.85 1.85
Frequency [Hz]
Frequency [Hz]
1.8
3.8
1.75
3.75
1.7
3.7
1.65
3.65 1.6
0 5 10 15 20 25 30 35 9 10 11 12 13 14 15
Temperature [”C] Rotor speed [m/s]
Figure 7.65 – Natural frequency of 3 SS* mode (Regime 4) Figure 7.66 – Natural frequency of 2 SS* mode (Regime 4)
vs temperature vs rotor speed
Finally, Regime 5 is related to production setups with the strongest wind speed conditions. It is
characterized by an increase of the blades pitch angle with the wind speed in order to control the rotor
speed. This Regime is then commanded by the pitch angle, which in turn is defined according to the
wind speed. The relation between the pitch angle and the wind speed was already introduced in Figure
7.14. Again, the 1 SS* and 2 SS* present a large variability of their frequency values due to operational
310
Chapter 7
factors. The variation introduced by the pitch angle in the frequency values of the 1 SS* mode is shown
in Figure 7.67. Also the temperature seems to be an important predictor of the natural frequency of the
1 SS mode. This relationship is illustrated in Figure 7.68.
Table 7.16 – Correlation coefficients between natural frequencies and operational/ environmental factors (Regime 5)
w. s. temp. r. s. pitch
1.34 0.37
1 SS* 1 SS
1.32 Linear fit 0.365 Linear fit
1.3 0.36
Frequency [Hz]
Frequency [Hz]
1.28 0.355
1.26 0.35
1.24 0.345
1.22 0.34
-5 0 5 10 15 20 25 0 5 10 15 20 25 30
Pitch angle [º] Temperature [ºC]
Figure 7.67 – Natural frequency of 1 SS* mode (Regime 5) Figure 7.68 – Natural frequency of 1 SS* mode (Regime 5)
vs blades pitch angle vs temperature
Once the influence of the predictors was analysed for each regime, two alternative regression models
were considered: SM1 and SM2.
SM1 only considers the predictors with important physical relations with the natural frequencies. For
that reason, even predictors with high correlation coefficients (such as wind speed in Regime 4 and 5)
were not considered, since the coefficient value is not related to a physical relation with this predictor
but to a correlation between other predictors, as previously referred. The predictors considered for this
model are presented in Table 7.17. On the other hand, the SM2 model considers all predictors, aiming
to define a general regression model with potential application without any previous analysis (Table
7.17).
311
Structural Monitoring of Wind Turbines
Table 7.17 – Considered predictors for regression models SM1 and SM2
SM 1 Regime SM 2 Regime
Predictors 1 2 3 4 5 Predictors 1 2 3 4 5
w. s. x x x w. s. x x x x x
temp. x x x x x temp. x x x x x
r. s. x r. s. x x x x x
pitch x pitch x x x x x
The suitability of both models to produce forecasts was quantified through the computation of the
coefficient of determination , presented in equation (7.8). For its calculation, the Period 2 of data
was used to assess the quality of the forecasts (as previously referred, this period of data was not
considered in the construction of the regression models). The results obtained for the two models are
very similar (Table 7.18). In that sense, the SM 2 was chosen since it considers all predictors, making
its use more general to be implemented in any wind turbine in any regime.
It is observed from Table 7.18 that the modes related to rotor motion (1 SS*, 2 SS* and 3 SS*) present
the highest values of . The good quality of the predictions of these modes is justified by the fact
that the variability of these modes is mostly driven by the rotor speed, which is a known predictor. On
the other hand, the 1st tower bending modes (1 SS and 1 FA) presents the lower coefficients of
determination of all modes, showing that their variability is less dependent on the considered
operational/ environmental effects. One possible cause to the low coefficients obtained for some
modes might be related to the potential influence of asymmetric soil conditions on the natural
frequency values, which could be considered by using the yaw angle as predictor in the regression
model. However, it was decided not to use this predictor in order to assess the accuracy of the
monitoring system to detect foundation asymmetries (see section 7.3.6.6). Notwithstanding, in
situations in which the installation of the monitoring system is performed at an early stage of the
turbine operation, the consideration of the yaw angle in the model would increase the accuracy of the
system to detect abnormal structural changes.
Modes SM 1 SM 2
1 SS 0.295 0.297
1 FA 0.377 0.391
1 SS* 0.965 0.967
2 SS* 0.950 0.952
2 FA 0.458 0.458
2 SS 0.706 0.719
3 SS* 0.860 0.864
4 FA 0.777 0.780
312
Chapter 7
Although steel structures do not usually present a significant thermal inertia, concrete structures do.
Considering the importance of the support conditions (i. e. concrete foundation) in the definition of
the dynamic properties of wind turbines, alternative dynamic models were also analysed in order to
assess the influence of temperature records from previous time instants on the structure (see section
7.2.1.1). In that sense, 5 dynamic models were considered. These models consider all the predictors
from the SM 2 model (including the temperature at the instant under consideration), along with the
temperature records referred to different time delays of 3, 6, 12, 18 and 24 hours. For each dynamic
model referred to a temperature measurement from a previous time instant, the temperatures from the
more recent time instants were also considered. This means that, for a dynamic model considering
temperature record with a delay of 12 hours, the temperature measurements from the current time
instant and from time delays of 3, 6 and 12 hours were considered. The results obtained are shown in
Table 7.19. In this table, the dynamic models are identified by the largest time delay considered. It can
be seen that, although some improvements are noticed for the 1 FA and 3 SS* modes, smaller values
were obtained for the 2nd pair of tower bending modes. Since the 2 FA and 2 SS modes are very
sensitive to structural changes in the foundation condition and, consequently, represent important
features to assess the existence of damage, the static model SM 2 was considered to be adequate to
eliminate the environmental and operational effects on the frequency values.
Table 7.19 – Coefficients of determination obtained with the dynamic models (the largest time delay considered for each
model is indicated within parenthesis)
DM 1 DM 2 DM 3 DM 4 DM 5
Modes
(3 hours) (6 hours) (12 hours) (18 hours) (24 hours)
1 SS 0.296 0.293 0.292 0.284 0.299
1 FA 0.396 0.399 0.392 0.388 0.391
1 SS* 0.966 0.967 0.966 0.964 0.965
2 SS* 0.949 0.949 0.951 0.947 0.947
2 FA 0.450 0.445 0.436 0.440 0.428
2 SS 0.716 0.713 0.721 0.707 0.710
3 SS* 0.866 0.866 0.864 0.871 0.875
4 FA 0.775 0.774 0.771 0.767 0.770
In order to illustrate the quality of the SM 2 model, Figure 7.69 presents the predicted values of natural
frequency of the 1 SS* and 4 FA modes for some data sets from Period 2, alongside with the identified
values of these modes with the monitoring system. The accuracy of the prediction obtained with this
model is clearly visible in this figure.
313
Structural Monitoring of Wind Turbines
1 SS* 4 FA
4.4
Frequency [Hz]
Frequency [Hz]
1.35 4.3
4.2
1.3
4.1
4
1.25
3.9
1.2 3.8
1.265 1.27 1.275 1.28 1.285 1.29 1.44 1.445 1.45 1.455 1.46 1.465 1.47 1.475
Index 4 Index 4
x 10 x 10
Figure 7.69 – Identified and predicted frequency value of the 1 SS* and 4 FA using the SM 2 mode during (data
corresponding to Period 2)
Using the regression model SM2, the effect of operational and environmental effects on the natural
frequencies of the modes can be then minimized. Figure 7.70 shows the time evolution of the
frequency values of the 1 SS* mode before and after the removal of these effects during period 2. These
effects (mainly the rotor speed) have a large influence on the variation of the frequency values of this
mode. Notwithstanding, the SM2 model is capable of mitigating the influence of these effects, as seen
in the bottom figure.
Before
1.5
1.45
1.4
Frequency [Hz]
1.35
1.3
1.25
1.2
0 1000 2000 3000 4000 5000 6000 7000 8000 9000
Index
After
1.5
1.45
1.4
Frequency [Hz]
1.35
1.3
1.25
1.2
0 1000 2000 3000 4000 5000 6000 7000 8000 9000
Index
Figure 7.70 – Time evolution of the frequency values of the 1 SS* mode before (top) and after (bottom) the removal of the
operational and environmental effects during period 2
314
Chapter 7
Figure 7.71 presents the frequency values of the 2 FA vibration mode along the time. Although this
mode does not present a variability as high as the 1 SS* mode, it is also possible to attest its reduction
after the correction of the frequency values with the regression model SM 2.
Before
2.95
2.9
2.8
2.75
2.7
2.65
0 2000 4000 6000 8000 10000
Index
After
2.95
2.9
Frequency [Hz]
2.85
2.8
2.75
2.7
2.65
0 2000 4000 6000 8000 10000
Index
Figure 7.71 – Time evolution of the frequency values of the 2 FA mode before (top) and after (bottom) the removal of the
operational and environmental effects during period 2
The reduction of variability of the frequency values of the vibration modes is also evident in the
histograms of Figure 7.72. As can be seen, the 1st tower bending mode in the SS direction presents the
lowest reduction in its variability from the three considered modes. This result is in accordance with
the results from Table 7.18. The reduction obtained with the correction of the frequency values of the 2
FA mode is notorious in the histogram, where the corrected frequencies are concentrated in a
narrower range. Lastly, the histogram of the 1 SS* mode illustrates the large reduction of variability
obtained with the regression model. It is also interesting to note the considerable different shape of the
histogram of this mode before the correction, illustrating the different nature of these rotor modes in
comparison with tower modes.
315
Structural Monitoring of Wind Turbines
1000 1500
1000
1000
500 500
500
0 0 0
0.34 0.35 0.36 0.37 1.25 1.3 1.35 1.4 1.45 2.7 2.75 2.8 2.85 2.9
Frequency [Hz] Frequency [Hz] Frequency [Hz]
Figure 7.72 – Histograms of the natural frequencies of the 1 SS, 1 SS* and 2 FA modes before and after the removal of
operational and environmental effects
316
Chapter 7
2 FA
2.86
2.84
2.82
Frequency [Hz]
2.8
2.78
2.76
2.74
0 40 80 120 160 200 240 280 320 360
Yaw angle [º]
2 SS
2.9
2.85
Frequency [Hz]
2.8
2.75
0 40 80 120 160 200 240 280 320 360
Yaw angle [º]
Figure 7.73 – Corrected frequency values of the 2 FA (top) and 2 SS (bottom) from Period 1 and 2 vs yaw angle
Although in a smaller scale, it is also possible to identify the same variation of the frequency value of
the 1 SS vibration mode with the yaw angle (Figure 7.74). This variation is similar to the one presented
by the 2 SS, which reinforces the idea of a possible asymmetry of the foundation stiffness. The 1 FA
mode do not evidences a noticeable variation, probably due to its larger variability which probably
hinders possible fluctuations due to asymmetric foundation conditions.
Nevertheless, it is not possible to say, at this moment, that these conclusions attest the existence of
damage in the foundation or, more likely, in the tower-foundation connection. It is possible that this
non-homogeneity of the foundation is already present since the beginning of operation of the turbine
and that is related to slightly different soil conditions under the large foundation of the support
structure. In that sense, the installation of a continuous dynamic monitoring at the initial stage of the
turbine operation would lead to conclusive results, permitting to identify if this non-homogeneity is a
consequence of structural changes occurred throughout its period of life or if it just represents the
stiffness of the foundation conditions.
317
Structural Monitoring of Wind Turbines
1 SS 0.12
Wind speed [m/s:]
0.36
3-4
0.1 4-5
0.358 5-6
6-7
0.08
7-8
Acc. [m/s 2]
Frequency [Hz]
0.356
0.06
0.354
0.04
0.352 0.02
0.35 0
0 40 80 120 160 200 240 280 320 360 0 40 80 120 160 200 240 280 320 360
Yaw angle [º] Yaw angle [º]
Figure 7.74 – Corrected frequency values of the 1 SS from Figure 7.75 – Median values of the RMS values of
Period 1 and 2 vs yaw angle acceleration of the top sensor in the FA direction
according to the yaw angle, considering setups when the
turbine was operating and several wind speed ranges
In the previously presented Figure 7.29, the variation of RMS values of acceleration from narrow
ranges of wind speeds with the yaw angle was analysed. This plot is again presented in Figure 7.75. It is
interesting to see that the direction with the highest acceleration RMS values (around 180º - 240º)
coincides with the softest direction. Still, the acceleration levels around 70º are very low, which may
seems a contradiction. However, the number of recorded operating events with the nacelle oriented to
the yaw sector between 0º and 70º is smaller than for the other sectors, which may not be
representative and, thus, hinder the effect of the softest foundation when the turbine is oriented to this
sector.
The yaw sector between 110º and 140º is also a sector with high acceleration RMS values. This is the
direction of the main wind incidence, as shown in Figure 7.15. This sector is located in a transition
location between the stiffest and softest directions of the foundation. For that reason, the high
acceleration levels detected might be also related to the wind action itself (higher levels of turbulence)
than just motivated by a possible loss of stiffness of the foundation.
As previously referred, damage in wind turbine support structures is usually related to foundation
problems. For offshore turbines, scour problems are likely to occur, decreasing the strength of the
foundation and potentially decreasing the expected fatigue life of the structure (Sørensen and Ibsen,
2013). Although in a small scale, some problems of excessive levels of vibrations were also detected due
to damage in the interface connection between tower and foundation (Pelayo, López-Aenlle et al.,
2011; Currie, Saafi et al., 2015).
Apart from foundation problems, rotor blades are also elements prone to damage and with a costly
replacement, as shown in section 3.1.3. In that sense, it would be interesting to assess the sensitivity of
the identified rotor vibration modes to small structural damages of the blades.
In order to evaluate the ability of the dynamic monitoring system to detect the presence of small
damage in the previously referred elements of a wind turbine, a strategy based on the introduction of
artificial damage was followed. For the assessment of damage at the foundation, two scenarios were
studied. The damage scenario D1 is related to scour problems in offshore wind turbines. In that sense,
the evolution of the first natural frequency of the support structure with the scour depth presented in
318
Chapter 7
(Prendergast, Gavin et al., 2015) was considered (see Figure 7.1). A small scour depth of around 0.075
times the base diameter of the monopile is considered. For the case of the Senvion MM82, with a tower
base diameter of 4.300 m, this damage corresponds to a scour depth of around 0.32 m, while for the
offshore Vestas V90 wind turbine represents a depth of 0.38 m. This damage is said to correspond to a
decrease of the first tower bending mode of around 0.2 %. However this study only considers the
evolution of the first natural frequency, while nothing is said about the second mode of the support
structure. To overcome this situation, results of the studies introduced in (Sørensen and Ibsen, 2013)
and (Zaaijer, Tempel et al., 2002) were considered. These studies present a ratio between the decrease
of the second tower bending mode relative to the first tower bending mode of 2.1 times to 2.6 times,
respectively, for monopile foundations. Following these results, a conservative value of 2.0 was used in
this damage scenario. Since it is not expected that rotor blades modes are influenced by this damage,
their frequency values is not changed.
The second case of foundation damage (scenario D2) reproduces the frequency variation detected in
(Pelayo, López-Aenlle et al., 2011). This study describes the dynamic tests performed in three wind
turbines, from which two of them presented anomalous vibration levels. It was also found that these
turbines presented cracks at the foundation. The test indicated that the frequency values of the
identified vibration modes from these two turbines were consistently lower than the ones from the
third healthy turbine. Initially, the frequency deviations of the first and second tower bending modes
presented by the turbine with the smallest amplitude of variations were considered in this scenario.
However, since this damage scenario was clearly identified, a smaller variation was studied. Thus, a
variation 5 times smaller than the variation reported in the study was used for this damage scenario.
Due to the difficulty in the comparison between the 4 FA mode and the vibration modes presented in
this study, this mode was also not considered. Again, the frequency value of the rotor blades modes
was not changed for this scenario.
The last damage scenario (D3) is related to blade damage. As referred in section 7.1.2, the majority of
published studies refer the small sensitivity of the natural frequencies of the blade modes to small,
common damages. In that sense, the model of the Senvion MM82 developed in the HAWC2 code
(introduced in section 7.3.2) was used to assess the sensitivity of the 1 SS* and 2 SS* modes to
structural damage at the blades. Considering that the eigenvalue analysis performed by the HAWC2
code is related to the turbine with the blades according to an operating configuration but under parked
conditions, the expected frequency value of these two modes, introduced in section 7.3.6.2, was used to
tune the stiffness of the blades. Due to the high uncertainties in modelling the structural elements
located inside the nacelle, as well as the rotor blades, the estimation of the frequency value of the 3 SS*
mode was not possible. The comparison between the frequency values obtained with the numerical
model and with the monitoring system is introduced in Table 7.20.
Table 7.20 – Comparison between the expected frequency values of the 1 SS* and 2 SS* modes during the monitored
period (non-operating conditions) with the results obtained with the HAWC2 model
319
Structural Monitoring of Wind Turbines
The presence of damage at just one (or two) blades leads to an anisotropic rotor. In that sense, the
identification of such damage would be more easily detected through the identification of the 1Ω
harmonic (as illustrated in Figure 4.18) or through the identification of phenomena such as the ones
described in (Ramírez, Tcherniak et al., 2015), where a clear distinction in the singular values spectrum
of the structure is identified between an isotropic and an anisotropic rotor.
Therefore, the damage scenario D3 is referred to a similar deterioration of three blades, which may be
caused by continuous wear of the blades due to operation. The stiffness of the three blades was
decreased in 15 % over a length of 2 m, corresponding to 5 % of the total length of the blade. The
damage was located at around 33 % in chord length from the blade root, which is said to be the
location more prone to damage (Ciang, Lee et al., 2008). The imposed variation of stiffness led to a
decrease of -0.65 % of the 1 SS* and 2 SS* rotor vibration modes. It was also noticed that the imposed
stiffness variation was not significantly reflected in the other modes. For that reason, only the rotor
modes 1 SS* and 2 SS* were used to detect this damage scenario.
The variation of the natural frequencies of the vibration modes with the considered damage scenarios
is resumed in Table 7.21. It is considered that the damage from each scenario occurred at the middle of
Period 2. The artificial damage is thus introduced in the detected frequencies from the second part of
Period 2 by reducing their values according to the values presented in Table 7.21.
Table 7.21 – Variation of the natural frequencies of the modes associated with the damage scenarios
Δ [%]
Modes D1 D2 D3
The methodology followed to assess the presence of damage in the wind turbine is based on the use of
control charts, already introduced in section 7.2.2. This multivariate technique allows assessing the
evolution of several vibration modes in the same chart. For the construction of these control charts,
the residual error was used. As referred, the residual error is the difference between the identified
frequency value of a vibration mode and its predicted value according to the defined SM 2 regression
model, as illustrated in Figure 7.69.
The control charts obtained for an undamaged situation and the three damage scenarios are presented
in Figure 7.76. In their construction, groups with 36 observations were considered (corresponding to a
quarter of a day). It is seen that the damages are clearly detected in all situations. However, a different
upper limit control of the chart from the one defined in equation (7.15) was used due to the large
number of false alarms obtained with this limit. This situation is most likely due to the fact that
320
Chapter 7
features considered do not respect all the assumptions adopted for the definition of the limits
(Magalhães, Cunha et al., 2012). To overcome this limitation, a different limit was imposed. This limit
is defined by considering that 95 % of the values from the training period are considered within the
safety region.
Undamaged scenario D1
300 300
250 250
200 200
T2
150
T2
150
100
100
50
50
0
0
Time
Time
D2 D3
300 300
250 250
200 200
T2
150
T2
150
100
100
50
50
0
0
Time
Time
Figure 7.76 – control charts associated with a scenario where no damage is present and with the three defined
damage scenarios
This is a very relevant result, since it demonstrates that quite small damages in the foundation (and,
although with less accuracy, in the blades) can be detected with the developed monitoring
methodology. In the case of scenario D1, it is relevant to note that the identified damage
(0.075 ) is considerably smaller than the recommended design value of 1.30 (Det
Norske Veritas (DNV), 2011).
It should be further referred that a longer monitoring period would even permit the detection of
smaller frequency variations due to a better and more complete regression model and due to the
possibility of using more observations to compute the control chart, leading to more robust results.
321
Structural Monitoring of Wind Turbines
1
0.5
0
-0.5
-1
-1.5
-2
-2.5
0 100 200 300 400 500 600
Time [s]
Prior to the integration process, the acceleration time series were multiplied by a time window with the
aim of softening the start and end of the time series as discussed in section 6.4.1.1. These initial and
last parts are then disregarded in the integrated displacement time series. After an initial analysis of
several setups, it was considered that the disregarded of the initial and last 30 seconds would guarantee
a safety margin to avoid transient errors.
For the double integration process, the filter introduced in section 6.4.1.1 (Figure 6.12) was used. The
constants of the filter ( and ) were set to only consider the dynamic component of the
displacements.
A good solution to tune these two parameters would be possible with the temporary use of
instrumentation to measure the stress condition of the tower at an arbitrary condition. Since it is
possible to estimate the stress condition at any position of the support structure with the presented
methodology, the results obtained with the temporary instrumentation would be compared with the
ones estimated with the accelerometers in order to tune the parameters. Since it was not possible to
install this instrumentation, a compromise between the amplification of possible low frequency errors
and the loss of relevant dynamic component information was studied with the help of numerical tests
performed with the HAWC2 code (Larsen and Hansen, 2007). As result, the constants f and f were
defined as, respectively, 0.10 Hz and 0.15 Hz. In that sense, the analysis described in this section
regarding the fatigue assessment of the Senvion MM82 wind turbine is focused on the validation of the
322
Chapter 7
developed procedure considering only the dynamic component of the fatigue, and not on the
estimation of the real fatigue condition of the structure.
Figure 7.78 shows the obtained displacements at the measurement levels in the FA direction. As
expected, the displacements are higher at the top, in opposition to what was obtained with the
acceleration time series (Figure 7.77). This result was expected since the displacements are mainly
driven by the 1st pair of tower bending modes. In this figure, the transient initial and last parts are
indicated by the vertical dashed lines and correspond to the referred 30 seconds in each part.
0.15
+74.988
+48.392
0.1 +21.772
Displacements [m]
0.05
-0.05
-0.1
Figure 7.78 – Dynamic displacement time series in FA direction (the vertical dashed lines indicate the disregarded initial
and last part)
Once the dynamic displacements along the turbine support structure are obtained, it is possible to
estimate the dynamic component of the stress condition at any level of the tower. As an example, the
bending stress at one point of the tower section at the foundation level was computed using the
procedure based on the SSI-DATA identification algorithm described in section 6.4.1.1. The result is
presented in Figure 7.79.
40
30
20
Bending stress [MPa]
10
-10
-20
-30
-40
0 100 200 300 400 500 600
Time [s]
Figure 7.79 – Estimated dynamic bending stress at the foundation level due to motion in the FA direction (the vertical
dashed lines indicate the disregarded initial and last part)
323
Structural Monitoring of Wind Turbines
Lastly, the load cycles are counted with the Rainflow algorithm. The histogram with the number of
bending stress cycles associated with different stress amplitudes is shown in Figure 7.80. The damage
due to fatigue can then be assessed with the S-N curve defined for the detail under analysis.
350
300
250
200
Nº cycles
150
100
50
0
0 10 20 30 40 50 60 70 80
Bending stress [MPa]
Considering that sensors installed at the tower structure are capable of detecting the motion in both
orthogonal directions, the stress due to bending in the SS direction can also be computed with this
methodology. With the stress condition computed in both FA and SS directions, it is thus possible to
estimate the bending stress condition at any point of the tower cross section. Initially, the bending
moment vectors in the FA and SS directions are reoriented according to the original SCADA
referential (N-S, E-W). This coordinate rotation, illustrated in Figure 7.81, depends on the yaw angle.
σP, E-W
N N
FA
MN-S MN-S
direction SS
direction
MSS P σP,N-S
R.cos(β) β
W E W E
ME-W R.sin(β) ME-W
MFA
S S
Figure 7.81 – Coordinate rotation from FA-SS referential to Figure 7.82 – Calculation of the bending stress at a generic
NS-EW referential point P
324
Chapter 7
After defining the moment vectors in the original SCADA referential, it is then possible to estimate the
stress condition at any point of the support structure cross section. This operation is illustrated in
Figure 7.82 for a generic point . The stress condition at the point is then defined according to:
. sin . cos
(7.18)
with:
A discretization of the cross section in 18 sectors was considered in order to assess the fatigue
condition at the foundation level. The detail 71 from Eurocode 3 (European Comitee for
Standardization (CEN), 2005) was used to study the welding-on flange connection between the tower
and the foundation (Sørensen and Sørensen, 2011). The S-N curve corresponding to this detail was
already introduced in Figure 6.7.
Alongside with the bending stress estimation, it is also of utmost importance to have confidence on the
obtained results. Since the procedure estimates the acceleration time series as a sum of vibration
modes/ harmonics, it is necessary to evaluate if the estimations do not underestimate the motion of the
structure (due to a badly defined state-space model) or, on the other hand, overestimate the
importance of some frequency content (due to numerical residues). In that sense, the residual error ∆
from the forward innovation model (equation (5.80)) can be used to assess the quality of the estimated
acceleration time series. However, in the context of stress estimation, in which a double integration is
performed to estimate the displacement time series, the use of residual error as the only quality control
parameter is not sufficient.
Figure 7.83 shows an example of an estimated acceleration time series obtained from the forward
innovation model, using all the 7 installed sensors. The residual error ∆ obtained is low (2 %),
indicating that the estimation is good. The good agreement between the two signals is visually verified.
325
Structural Monitoring of Wind Turbines
0.03 0.03
Measured
0.02 Estimated 0.02
0.01 0.01
Acc. [m/s2]
Acc. [m/s2]
0 0
-0.01 -0.01
-0.02 -0.02
-0.03 -0.03
0 100 200 300 400 500 600 325 330 335 340 345 350
Time [s] Time [s]
Figure 7.83 – Measured and estimated acceleration time series at the level +21.772 in the SS direction, considering all
sensors as input for the forward innovation model (the right hand side figure shows a zoom of the time segment
identified in the left hand side figure by the dashed lines)
However, looking at these results in the frequency domain, important errors in the lowest part of the
frequency range are observed. Figure 7.84 shows the power spectral density of the measured and
estimated acceleration time series illustrated in Figure 7.83 (the spectrum is averaged with segments of
1024 points in order to facilitate the explaination). From this figure, it is possible to attest that,
although a good estimation is obtained around important resonance peaks (namely the 1st and 2nd
tower bending modes, as well as the 3Ω and 6Ω harmonics), there is an erroneous increase of the
amplitude of several peaks in the low part of the spectrum. Since the response is clearly dominated by
the 2nd bending mode, these errors are almost imperceptible in time domain, as evidenced in Figure
7.83. This is the reason why the error ∆ is low.
-4 Measured
10
Estimated
-6
10
Amplitude
-8
10
-10
10
Figure 7.84 – Measured and estimated acceleration time series (in frequency domain) at the level +21.772 in the SS
direction, considering all sensors as input for the forward innovation model
The main problem related to the bad estimation of the lower part of the spectrum occurs during the
integration process to estimate the displacement time series. As referred in section 6.4.1.1, during this
operation, the amplitude of vibrations related to the low frequencies is increased. Consequently, the
displacement time series obtained from the measured and estimated acceleration will lead to
considerably different results. Figure 7.85 illustrates the displacement time series obtained from the
326
Chapter 7
measured and estimated acceleration signals represented in Figure 7.83. The erroneous majoration of
the displacements obtained from the estimated signals is evident. Under these circunstancies, the
estimation of the stress condition would naturally lead to erroneous results.
-3
x 10
1.5
Measured
1 Estimated
0.5
Disp. [m]
-0.5
-1
-1.5
0 100 200 300 400 500 600
Time [s]
Figure 7.85 – Displacement time series obtained with the double integration process based on the measured and
estimated acceleration signals
Due to possible occurrence of situations such as the one described, an additional quality index of the
estimation is required to assess the adjustment in the lowest part of the spectrum. An index ∆
based on the relative difference of the energy countained in the lowest part of the frequency range of
the estimated and measured acceleration signals was chosen. This index is defined as the ratio between
the area defined by the difference of the spectra of the estimated and measured acceleration and the
area defined by the spectrum of the measured acceleration signal (dark grey area in Figure 7.86). Only
the frequency range between and is considered to compute this index. In the present work, these
values were defined as 0.10 Hz (equal to ) and 0.35 Hz, in order to evaluate the adjustment of the
signal for the frequency range below the 1st tower bending mode.
f1 f2
-4 Measured
10
Estimated
-6
10
Amplitude
-8
10
-10
10
Figure 7.86 – Definition of the area between the spectra of the estimated and measured acceleration signals and the area
from the measured signal
The results obtained for the quality indices using data sets collected during one year of monitoring
were initially assessed. It was observed that the quality index ∆ is not very restrictive. Figure 7.87 a)
327
Structural Monitoring of Wind Turbines
shows the evolution of percentage of selected 10 min. time series events with the value adopted for ∆
considering the results obtained with the three tested modal identification algorithms. It is seen that
for the method based on the SSI-DATA algorithm almost all events are considered for an admissible
level of 0.15. For the other algorithms, it is visible that the estimations are apparently not so good.
100 100
Percentage of considered events [%]
60 60
40 40
SSI-DATA SSI-DATA
20 20 SSI-COV
SSI-COV
p-LSCF p-LSCF
0 0
0 0.1 0.2 0.3 0.4 0.5 0 0.5 1 1.5 2 2.5 3
Quality index Δe Quality index Δfreq
a) b)
Figure 7.87 – Percentage of considered events with the evolution of: a) ∆ index; b) ∆ index
The values computed for the second quality index ∆ , referred to the quality of the adjust on the
lowest frequency range of the spectrum, were also initially assessed. The percentage of events
considered with the evolution of the ∆ index is presented in Figure 7.87 b). It was concluded that
most results contain an increase of the energy in the frequency range considered for this index (0.1 –
0.35 Hz). It was thus noticed that the biggest problem associated with the bad estimation of this part of
the spectrum is related to the unrealistic increase of the computed stress levels of the structure. It is
interesting to note that the quality of the results provided by the alternative identification methods is
inverse to the quantified by the index ∆ . Nevertheless the difference is not very large.
Based on this preliminary analysis of the quality indices, a strategy was defined to accurately disregard
bad estimations of the fatigue damage. In that sense, it was decided to use the ∆ index to identify and
dismiss events in which the defined forward innovation model was clearly not capable to estimate the
acceleration time series as a sum of contributions of vibration modes/ harmonics. A value of ∆ = 0.20
was thus chosen, since it considers almost all events and, at the same time, imposes a limit that
guarantees the quality of the estimates. The maximum admissible value of the ∆ was defined to
admit a maximum error ∆ = 0.90. With this value, a rate of considered events of more than 90 %
was achieved when the combination of the best estimations obtained from the three algorithms is
computed.
At this point, the fatigue damage estimation from the recorded acceleration time series can be
computed. Figure 7.88 shows the evolution of accumulated damage calculated with the SSI-DATA
algorithm at the welding-on flanged connection between the tower and the foundation using the S-N
curve from the GL standard. From this figure, two periods of time with a rapid increase of damage are
noticed. The first period corresponds to events occurred during November 2013, while the second
occurs in February 2014. It is interesting to note that these two periods correspond to two long periods
of consecutive data sets with high mean wind speeds.
328
Chapter 7
It is also interesting to confirm that the most damaged sectors are the 100º-120º (and its symmetric
280º-300º) which is in accordance with the main wind direction of the wind (around 110º, see Figure
7.15).
-4
x 10
8
[0º-20º]/[180º-200º]
7 [20º-40º]/[200º-220º]
[40º-60º]/[220º-240º]
6
Accumulated Damage
[60º-80º]/[240º-260º]
5 [80º-100º]/[260º-280º]
[100º-120º]/[280º-300º]
4 [120º-140º]/[300º-320º]
[140º-160º]/[320º-340º]
3
[160º-180º]/[340º-360]º
2
0
Jul Aug Sep Oct Nov Dec Jan Feb Mar Apr May Jun Jul
2013 2013 2013 2013 2013 2013 2014 2014 2014 2014 2014 2014 2014
Time
Figure 7.88 – Time history of the accumulated damage at the foundation level (with S/N curve from GL standard) using
the methodology based on the SSI-DATA algorithm
A direct extrapolation of the accumulated damage to the expected life of the wind turbine (20 years)
can be performed according to equation (7.16). The results obtained with the three methods are
presented in Table 7.22. As expected, the values obtained with the S-N curve from the EC3 code are
slightly lower due to the consideration of the threshold referred in section 6.3.2 (Figure 6.7).
Table 7.22 – Direct extrapolated accumulated damage at the end of the design life
The results obtained present very low values of accumulated damage. Nevertheless, some important
aspects should be highlighted. Firstly, it should be noted that the estimated damage is only referred to
the dynamic component of the bending stress. Large loading cycles from the quasi-static frequency
regime, which might have important contributions to the fatigue damage of the structure are not
considered in these results. Two additional causes for this low accumulated damage should be also
referred. The location of the Senvion MM82 wind turbine is considered a low-turbulence site, which
means that high amplitude stress variations are not expected to occur frequently. This aspect has a
direct impact on the estimation of the fatigue life of the analysed structural detail, since it only depends
on the measured amplitude of the stress cycles and not on the mean stress level. In addition, technical
problems occurred at the beginning of October, preventing the operation of the dynamic monitoring
system during this month and almost the entire month of November. Thus, considering the detected
high rate of damage during the end of November, it is very likely the occurrence of important events
329
Structural Monitoring of Wind Turbines
during this period that were not measured, which may have masked the real value of accumulated
damage at the end of the monitoring period.
Considering the confidentiality usually associated with projects related to the measurement of real
fatigue damage values, it is not easy to frame the obtained results. Nevertheless, the work presented in
(Loraux and Brühwiler, 2015), related to the fatigue analysis of a wind turbine tower using strain
gauges, highlights the conservatism of the current codes, based on a preliminary assessment of
experimentally obtained accumulated fatigue damage. Although the results are not comparable, it is
interesting to note that, even when the static, quasi-static and dynamic responses are taken into
account, a considerably higher fatigue lifetime than the design value of 20 years is expected.
The direct extrapolated estimation of total accumulated damage due to the dynamic component of the
bending stress shows a considerably variation, with values between 0.0420 and 0.0213 (using the S-N
curve from the GL standard) and 0.0319 and 0.0195 (using the S-N curve from the EC3 standard). This
discrepancy is mainly related to the different acceleration time series considered in each algorithm for
fatigue estimation (due to the quality indices restrictions), leading to different results at the end. This
situation is partially shown by the different success rate presented at the last line of the table. The
methodology using the SSI-DATA presents a higher rate. These values are comprehensible since the
original methodology is based on the SSI-DATA, while the other two are adaptations of the method.
The methodology based on the p-LSCF presents the worst success rate probably because it requires the
manipulation of the state-space matrices to remove the unstable poles.
One consequence of using different acceleration events to extrapolate the accumulated damage at the
end of the expected fatigue life of the structure is that different wind loading conditions are
considered. Figure 7.89 a) shows the number of events for each mean wind speed in the sector 100º -
120º (corresponding to the main wind incidence direction) during one year using the data recorded by
the SCADA system. Alongside, Figure 7.89 b) presents the probability density function (PDF) of the
same data. In the same figures, the distribution of the events considered in the fatigue estimation with
each algorithm is also shown. It is seen that, while the events considered for the SSI-DATA algorithm
seems to follow the wind speed distribution, the other two methods diverge considerably. The SSI-
COV shows an increase in the occurrence of events in the low (lower than 7 m/s) and high wind
speeds (higher than 15 m/s), while there is a notorious lack of events in the range 7 - 15 m/s. On the
other hand, the events considered with the p-LSCF algorithm focus on the low wind speed region.
These facts contribute to the discrepancy in the results presented in Table 7.22.
330
Chapter 7
500 25 %
1 year 1 year
SSI-DATA SSI-DATA
300 15 %
200 10 %
100 5%
0 0
0 5 10 15 20 25 0 5 10 15 20 25
Wind speed [m/s] Wind speed [m/s]
a) b)
Figure 7.89 – Histogram a) and probability density function b) of mean wind speed (in the sector 100º-120º) during one
year according to the data recorded by the SCADA system and to the data considered for fatigue estimation with the
three algorithms
A simple way to attest the quality of the results obtained with the three algorithms is the computation
and comparison of damage estimated from the same acceleration time series events. In that sense, the
results obtained with the SSI-DATA were considered as reference, since they consider more data sets,
and only the events that were both considered for the other algorithms and the SSI-DATA were used
to compute the estimated damage. The accumulated damage obtained with SSI-COV and p-LSCF
algorithm is directly compared to the one computed with the SSI-DATA since they are related to the
same acceleration events. Table 7.23 shows the results obtained with this study, presenting the
deviation obtained relative to the reference method (SSI-DATA algorithm). As can be seen, the
agreement is now much better, especially for the estimation based on the SSI-COV where an error
lower than 8 % was achieved.
Table 7.23 – Deviation of accumulated damage of coincident events relative to the methodology based on SSI-DATA
algorithm
GL - + 7.5 % - 19.8 %
EC3 - + 7.8 % - 20.6 %
The advantage of using different algorithms to estimate the fatigue damage for the same period of time
is the ability to choose the best estimations obtained with each method for each acceleration time
series event according to the quality indices. The results obtained for the direct extrapolated damage
(according to the S-N curve from the GL and EC3 standards) using the best estimations from the three
algorithms are shown in Table 7.24. It is visible the increase of the success rate to 91.0 %.
331
Structural Monitoring of Wind Turbines
Table 7.24 – Estimation of accumulated damage at the end of design life considering a direct extrapolation and an
adjusted extrapolation according to the PDF of the wind speed using the best estimations from the three algorithms
GL 0.0371 0.0357
EC3 0.0283 0.0271
Success rate [%] 91.0
Another advantage of using different algorithms to estimate the fatigue damage for the same period of
time is the possibility of consider more loading scenarios and, consequently, approximate the
considered wind conditions to the actual distribution of wind conditions. This advantage is illustrated
in Figure 7.90, where the PDF of the mean wind speed in sector 100º - 120º obtained with the data
considered when the best estimations from the three algorithms were used is compared with the PDF
defined with the data recorded by the SCADA system during one year. Both PDFs seem to be similar
enough to allow an adjustment of the results obtained with the best estimations to follow the “real”
distribution of wind speed (as introduced in section 7.2.2.1). In this case, it was only possible to
quantify the damage accumulation associated with the mean wind speed of 10 min. period. The results
of accumulated damage at the end of the design life after adjusting the loading scenarios according to
the PDF of the mean wind speed are shown in Table 7.24.
12
1 year
10 Best Estimation
Percentage of occurrence [%]
0
0 5 10 15 20 25
Wind speed [m/s]
Figure 7.90 – Probability density function of mean wind speed (in the sector 100º-120º) during one year of data recorded
by the SCADA system (1 year) and the events included in the best estimation obtained with the three algorithms
Naturally, the estimation of damage due to the dynamic component of fatigue would improve if more
data from the monitoring system was available or, in case it is not possible, if a characterization of the
wind conditions over a longer period was known.
332
Chapter 7
number of sensors, both for modal tracking and fatigue assessment. At the most extreme solution, it
would be interesting to only use the top sensor as part of the monitoring system.
This section introduces the results obtained with the two modules of the monitoring system using
three different solutions of sensors layout. The first solution uses only one biaxial sensor at the top
level (+74.988 m) to perform both modal tracking (for damage detection) and fatigue analysis. This
represents the aforementioned most interesting solution from the economic point of view. The second
proposed solution uses only one biaxial sensor at the +48.392 m level of measurement (around 2/3 of
the tower height). Lastly, the third solution is based on the use of two biaxial sensor placed at +74.988
m and at +48.392 m levels.
333
Structural Monitoring of Wind Turbines
Campbel Diagram
4.5
4 FA
4 3 SS*
3.5 3 FA/SS
3 2 SS
Frequency [Hz]
2.5 2 FA
2 2 SS*
1.5
1 1 SS*
0.5 1 SS
1 FA
0
0 2 4 6 8 10 12 14 16 18
Rotor Speed [RPM]
Figure 7.91 – Campbell diagram with the tracked vibration modes with Layout 1 (SSI-COV algorithm)
The statistics related to the success in the identification of the 9 vibration modes are introduced in
Table 7.25. Comparing these results with the ones obtained when all sensors are used (Tables 7.6 to
7.8), it is concluded that the success rate of the first 4 modes have increased when only the top sensors
are considered. Although this might seem a contradiction, it is explained by the lower MAC condition
imposed to consider a cluster of poles. With this MAC condition, it is only possible to distinguish
between FA and SS modes which are, in fact, the only information that can be extracted from the mode
shapes with this sensors layout. The coefficient of variation of the frequency values of these modes
does not present an important increase relative to the solution using all sensors, which is a good
indicator for their use in damage assessment.
As regards the last 3 modes, the success in their identification was rather low and, for that reason, these
modes were not considered for damage detection.
Table 7.25 – Success rate and coefficient of variation obtained with the identification algorithms (with Layout 1)
The ability of the dynamic monitoring system to detect damage was then assessed using the first pair
of bending modes (1 FA and 1 SS) and the rotor blades modes 1 SS* and 2 SS*. Initially, the same
334
Chapter 7
damage scenarios shown in Table 7.21 (for a sensor layout with all sensors) were tested. It was verified
that the damage scenario D1, related to the scour problems in an offshore wind turbine, was not
properly detected. This scenario is especially relevant to the second pair of bending modes which are
not detected with this solution. For that reason, a new scenario D1* related to scour damage was
defined, reflecting the minimum detectable frequency variation of the first pair of modes. It was
concluded that, in order to be detectable, the scenario D1* would have to imply a frequency variation
of -0.50 % of the first pair of support structure bending modes (an increase of 2.5 times when
compared to the original D1 scenario). In terms of scour damage, the new damage scenario D1*
represents a depth of 0.162 times the base diameter of the monopile, i. e., a depth of 0.70 m and 0.82 m
if the Senvion MM82 tower base and the Vestas V90 monopile are used as reference, respectively.
Although this value is more than twice the value achieved when all sensors are used, it still represents a
very early state of damage, being considerably lower than the recommended design value of 1.30.
(Det Norske Veritas (DNV), 2011).
The second damage scenario D2, reproducing a frequency variation 5 times smaller than the one
described in a test of damaged onshore wind turbines (Pelayo, López-Aenlle et al., 2011), was also
analysed and it was concluded that, even with this layout solution, it would be detected. In fact, the
variation imposed by this damage scenario for the first pair of tower bending modes is similar to the
one observed in the scenario D1*.
The last damage scenario (D3) is related to blade damage. It was concluded that, for the same imposed
stiffness variation described in section 7.3.6.6, the damage is still detected with this layout solution.
Table 7.26 – Variation of the natural frequencies of the tracked modes associated with the damage scenarios
Δ [%]
Modes D1* D2 D3
Figure 7.92 introduces the control charts obtained for the referred damage scenarios with Layout 1. As
can be seen, the damage is easily detected which attest the feasibility of this solution. Although some
accuracy was lost in the damage scenario D1, it is seen that it provides a good solution for detection of
damage at an early stage.
335
Structural Monitoring of Wind Turbines
250
200
200
150
T2
T2
150
100
100
50
50
0 0
Time Time
D3
250
200
150
T2
100
50
Time
Figure 7.92 – control charts associated with a scenario where no damage is present and with the three defined
damage scenarios (Layout 1)
Fatigue Results
The estimation of fatigue damage using a single level of measurement is an interesting application for
the methodology introduced with the virtual sensors in section 6.4.2. With this methodology, it is
possible to estimate the response of the wind turbine support structure at any point of the structure,
using only a (“well positioned”) single reference measurement point. In the case of the Layout 1, this
reference point is the top measurement level. However, as was previously referred, this measurement
position cannot be considered as “well positioned” since it coincides with a level of very low modal
amplitude of an important vibration mode. Thus, it is not possible to accurately estimate the mode
shape of this mode. In that sense, it was decided to despise all vibration modes and harmonics with a
frequency value higher than 2 Hz. With this imposition, not only the second pair of bending modes is
disregarded, but also some integer multiples of the harmonics are avoided since their operational
deflection shapes also present a very low modal amplitude at the top of the tower.
The importance of the second pair of tower bending modes in the context of the dynamic behaviour of
the wind turbine structure was already described in section 7.3.6.4. It is thus expected that, using this
sensor layout, the estimation of the acceleration at the middle and lower part of the tower will be
underestimated. Figure 7.93 illustrates this problem, using the acceleration time series collected by the
S8 and S9 sensors during a production period (rotor speed = 12.4 rpm). The acceleration signals were
then processed by the methodology based on the p-LSCF algorithm. The top figures compare the
336
Chapter 7
original signal recorded by the sensor in the FA direction with the estimated signal as a sum of modes/
harmonics at the measured level in time and frequency domain. On the other hand, the plots at the
middle and bottom of Figure 7.93 are referred to the estimations of the acceleration at unmeasured
levels. Taking advantage of the non-used sensors (at levels +21.722 m and +48.392 m), the estimated
acceleration time series at unmeasured locations are compared with the real measured time series.
+74.988 (FA) - measured 2 +74.988 (FA) - measured
0.2 10
Original Original
0.15 1
Estimated 10 Estimated
0.1
0.05 0
Acc. [m/s2]
Amplitude
10
0
-1
-0.05 10
-0.1 -2
10
-0.15
-3
-0.2 10
0 100 200 300 400 500 600 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5
Time [s] Frequency [Hz]
0.05 0
Acc. [m/s2]
Amplitude
10
0
-1
-0.05 10
-0.1 -2
10
-0.15 Original
-3 Estimated
-0.2 10
0 100 200 300 400 500 600 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5
Time [s] Frequency [Hz]
Amplitude
-2
10
-0.05
Original
-4 Estimated
-0.1 10
0 100 200 300 400 500 600 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5
Time [s] Frequency [Hz]
Figure 7.93 – Original and estimated acceleration time series at the measured level (+74.988 m) and unmeasured levels
(+48.392 m and +21.772 m) in the FA direction
The observation of the plots in Figure 7.93 clearly shows that, as expected, the disregard of the energy
contribution from the spectrum for frequency values higher than 2 Hz leads to an important
underestimation of the acceleration at the lowest 2/3 of the tower. Nevertheless, it is seen that, for the
337
Structural Monitoring of Wind Turbines
considered frequency range, the agreement between the estimated and the real acceleration signal is
quite good. Even the contribution of the 3Ω harmonic (represented by the large peak around 0.62 Hz),
which requires the estimation of its operational deflection shape, shows a good agreement.
The processing for fatigue damage estimation was then applied to the estimated acceleration signals. In
order to illustrate the quality of the results obtained with this layout configuration, the dynamic
component of the bending stress from a 10 min. period acceleration event at the foundation level was
computed. The results were compared with the stress estimation obtained with the methodology based
on the SSI-DATA considering the three levels of measurement, since it is considered the most accurate
of the implemented procedures. The bending stress computed with the two procedures is shown in
Figure 7.94. It is visible that the bending stress is underestimated mostly due to the non-consideration
of the energy of the spectrum for frequency values higher than 2 Hz. Computing the damage
associated with the two estimations (using the S-N curve from the GL standard), it is concluded that
the value obtained with the procedure based on the p-LSCF with Layout 1 is 3.2 times lower than the
value achieved with the SSI-DATA methodology using all sensors.
2 9
10
SSI-DATA (all sensors
1.5 SSI-DATA (all sensors
p-LSCF (Layout 1) 8
10 p-LSCF (Layout 1)
1
Bending stress [MPa]
7
0.5 10
Amplitude
6
0 10
-0.5 5
10
-1
4
10
-1.5
3
-2 10
0 100 200 300 400 500 600 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5
Time [s] Frequency [Hz]
Figure 7.94 – Estimated dynamic bending stress at the foundation level due to motion in the FA direction (the vertical
dashed lines indicate the disregarded initial and last part)
The same quality indices introduced in section 7.3.7 (∆ and ∆ ) were used to despise erroneous
estimations of fatigue damage. Naturally, these indices are only referred to the results associated with
the considered S8 and S9 sensors. Figure 7.95 illustrates the evolution of percentage of acceleration
events with the increase of the indices. Comparing these results with the ones presented in Figure 7.87,
it is visible with the index ∆ that the errors associated with the quality of estimation of the
acceleration time series are higher. Qualitatively, it is seen that the procedure based on the SSI-DATA
algorithm shows the best results, followed by the SSI-COV and, lastly, the p-LSCF, just like when all
sensors were used. In relation to the index ∆ , the p-LSCF algorithm shows, again, the best results.
However, with this sensor layout solution, the difference to the other methods is higher than when all
sensors are used.
338
Chapter 7
100 100
Percentage of considered events [%]
60 60
40 40
SSI-DATA SSI-DATA
20 SSI-COV 20 SSI-COV
p-LSCF p-LSCF
0 0
0 0.1 0.2 0.3 0.4 0.5 0 1 2 3 4
Quality index Δe Quality index Δfreq
a) b)
Figure 7.95 – Percentage of considered events with Layout 1 with the evolution of: a) ∆ index; b) ∆ index
As referred in section 7.3.7, the comparison of the direct extrapolated damage obtained with each
methodology does not lead to a correct estimation of the quality of the results since it compares
estimations of damage from different events. It is visible in the probability density function of Figure
7.96 a) that the considered events for each methodology differ considerably from each other.
20 12
1 year 1 year
SSI-DATA 10 Layout 1
Percentage of occurrence [%]
SSI-COV
15
p-LSCF
8
10 6
4
5
2
0 0
0 5 10 15 20 25 0 5 10 15 20 25
Wind speed [m/s] Wind speed [m/s]
a) b)
Figure 7.96 – Probability density function of wind speed (in the sector 100º-120º) recorded by the SCADA system and
from the data considered for fatigue estimation with Layout 1 with: a) the thee algorithms; b) best estimations from the
thee algorithms
Thus, the estimated damage obtained with each methodology was compared with the damage
estimated with the SSI-DATA using all sensors (considered the reference methodology) for the same
events. The deviations obtained with the three methodologies are presented in Table 7.27.
339
Structural Monitoring of Wind Turbines
Table 7.27 – Deviation of accumulated damage of coincident events between results with Layout 1 and the methodology
based on SSI-DATA algorithm using all sensors
From the results, it is evident that the damage estimations with Layout 1 were not very good. The
damage obtained with the three methodologies is clearly underestimated. Also the success rate
achieved is low, with only one methodology (based on the SSI-DATA) showing a rate higher than 50
%.
Lastly, the fatigue damage using the best estimations from the three algorithms was computed. With
this estimation, the distribution of considered wind speed events is close to the distribution obtained
with the SCADA data (Figure 7.96 b)). The results are introduced in Table 7.28, where they are
compared to the ones obtained with the best estimation using all sensors. Again, the values are
considerably underestimated.
Table 7.28 – Estimation of accumulated damage at the end of design life considering an adjusted extrapolation according
to the PDF of the wind speed using the best estimations from the three algorithms
340
Chapter 7
of the 3 SS* mode. In fact, some clusters referred to this harmonic are misidentified as being from this
mode. As for the other vibration modes, no further erroneous modal identifications were detected.
Campbel Diagram
4.5
4 FA
4 3 SS*
3.5 3 FA/SS
3 2 SS
Frequency [Hz]
2.5 2 FA
2 2 SS*
1.5
1 1 SS*
0.5 1 SS
1 FA
0
0 2 4 6 8 10 12 14 16 18
Rotor Speed [RPM]
Figure 7.97 – Campbell diagram with the tracked vibration modes with Layout 2 (p-LSCF algorithm)
The success rate in the identification of the 9 vibration modes is summarized in Table 7.29. It is
concluded that the results are good for the 1 SS, 1 SS*, 2 SS*, 2 FA and 2 SS modes, with high rates of
success in the tracking process. On the other hand, the identification of the 1 FA presents a very low
success rate, especially for the p-LSCF.
Table 7.29 – Success rate and coefficient of variation obtained with the identification algorithms (with Layout 2)
After the modal tracking, the environmental and operational effects were removed from the frequency
values, using a multivariate analysis (similar to the procedure described in section 7.3.6.5). Thus, using
the frequency values as damage indicators, the three damage scenarios presented in section 7.3.6.6
(considering the dynamic monitoring system with three levels of measurement) were tested (see Table
7.21).
341
Structural Monitoring of Wind Turbines
The control charts obtained are shown in Figure 7.98. It is evident that the three damage scenarios are
well identified. Notwithstanding, the detection of damage referred to scenario D1 is not as clear as
when all sensors are used.
Undamaged scenario D1
300 300
250 250
200 200
T2
T2
150 150
100 100
50 50
0 0
Time Time
D2 D3
300 250
250
200
200
150
T2
T2
150
100
100
50
50
0 0
Time Time
Figure 7.98 – ܶ ଶ control charts associated with a scenario where no damage is present and with the three defined
damage scenarios (Layout 2)
Fatigue Results
Considering the position of the selected measurement level, it is expected that the problems noticed
during the estimation of the acceleration time series with Layout 1 would disappear with this solution.
In that sense, the frequency cut of 2 Hz defined with the Layout 1 was not imposed.
An example of the estimation of the acceleration time series obtained during production regimes (with
a mean rotor speed of 11.7 RPM) in introduced in Figure 7.99. In this figure, the estimation of the
acceleration at the measured level (+48.392 m) is presented alongside with the estimations at two
unmeasured levels. Again, the +74.988 m and +21.772 m levels were selected in order to be possible to
compare the estimations with the real acceleration time series. In this example, the methodology based
on the SSI-COV algorithm was used. The very good agreement of the estimations at the unmeasured
spots is notorious.
342
Chapter 7
0.05 0.05
Acc. [m/s2]
Acc. [m/s2]
0 0
-0.05 -0.05
-0.1 -0.1
-0.15 -0.15
0 100 200 300 400 500 600 105 110 115 120 125 130
Time [m] Time [m]
0.05 0.05
Acc. [m/s2]
Acc. [m/s2]
0 0
-0.05 -0.05
-0.1 -0.1
-0.15 -0.15
-0.2 -0.2
0 100 200 300 400 500 600 105 110 115 120 125 130
Time [m] Time [m]
0.02 0.02
Acc. [m/s2]
Acc. [m/s2]
0 0
-0.02 -0.02
-0.04 -0.04
-0.06 -0.06
-0.08 -0.08
0 100 200 300 400 500 600 105 110 115 120 125 130
Time [m] Time [m]
Figure 7.99 – Original and estimated acceleration time series at the measured level (+48.392 m) and unmeasured levels
(+74.988 m and +21.772 m) in the FA direction
Naturally, the same quality indices were used to discharge erroneous estimations. It is noted that,
qualitatively, the results are similar to the ones obtained with Layout 1. Nevertheless, there is an
important improvement in the index ∆ for the methodology based on the SSI-COV algorithm.
343
Structural Monitoring of Wind Turbines
100 100
Percentage of considered events [%]
60 60
40 40
SSI-DATA SSI-DATA
20 SSI-COV 20 SSI-COV
p-LSCF p-LSCF
0 0
0 0.1 0.2 0.3 0.4 0.5 0 1 2 3 4
Quality index Δe Quality index Δfreq
a) b)
Figure 7.100 – Percentage of considered events with Layout 2 with the evolution of: a) ∆ index; b) ∆ index
The considered events for each methodology are again not enough to independently reproduce the
density function of the wind speed, as seen in Figure 7.101 a). Thus, the same strategy previously
followed to compare only the coincident events between the reference results (methodology based on
SSI-DATA, using all sensors) and the various methods using Layout 2 was followed. The results are
summarized in Table 7.30.
20 12
1 year 1 year
SSI-DATA 10 Layout 2
Percentage of Occurrence [%]
SSI-COV
15
p-LSCF
8
10 6
4
5
2
0 0
0 5 10 15 20 25 0 5 10 15 20 25
Wind speed [m/s] Wind speed [m/s]
a) b)
Figure 7.101 – Probability density function of wind speed (in the sector 100º-120º) recorded by the SCADA system and
from the data considered for fatigue estimation with Layout 2 with: a) the thee algorithms; b) best estimations from the
thee algorithms
The results obtained confirm the good premise indicated by the quality of the estimations. Both the
methodologies based on the SSI-DATA and SSI-COV show damage estimations very close to the ones
obtained with the reference results. The success rate achieved with the SSI-COV algorithms should
also be highlighted, which is considerably higher than the with SSI-DATA. This is an indication of the
quality of the adaptation of the post-processing tools presented in section 5.5 (originally developed to
the SSI-DATA) to the SSI-COV algorithm. Again, the p-LSCF presented the worst results, both for the
success rate and damage estimation.
344
Chapter 7
Table 7.30 – Deviation of accumulated damage of coincident events between results from Layout 2 the methodology
based on SSI-DATA algorithm using all sensors
Lastly, the best estimations from the three algorithms were selected to estimate damage. With this
consideration, the success rate increased to values similar to the ones obtained when the three levels of
measurement are considered. At the same time, the distribution of the wind speed becomes similar to
the one recorded by the SCADA system during one year (Figure 7.101 b)). Under these circumstances,
the estimated damage is again close to the estimation using all sensors, with an error lower than 15 %.
Considering the sensitivity of the S-N curve to small variations of the load cycles and the reduced
number of recording channels used with this solution, this result has to be considered a very good
approximation.
Table 7.31 – Estimation of accumulated damage at the end of design life considering an adjusted extrapolation according
to the PDF of the wind speed using the best estimations from the three algorithms
345
Structural Monitoring of Wind Turbines
The Campbell diagram obtained for the analysed period of data with this layout solution is shown in
Figure 7.102. The results are in good agreement with the ones obtained when all sensors were used.
However, it is noted a small influence of the 15Ω harmonic on the identification of the 3 SS* vibration
mode, although in a smaller extent than in Layout 2.
Campbel Diagram
4.5
4 FA
4 3 SS*
3.5 3 FA/SS
3 2 SS
Frequency [Hz]
2.5 2 FA
2 2 SS*
1.5
1 1 SS*
0.5 1 SS
1 FA
0
0 2 4 6 8 10 12 14 16 18
Rotor speed [RPM]
Figure 7.102 – Campbell diagram with the tracked vibration modes with Layout 3 (p-LSCF algorithm)
The statistics about the identification of the 9 vibration modes are summarized in Table 7.32. A
significant improvement is clearly noticed in the success rate of the identification of almost all
vibration modes. In fact, for some modes, the results are better than the ones obtained when all sensors
are used. Nevertheless, the identification of the 1 FA mode by the p-LSCF algorithm is slightly lower
than when all sensors are considered.
Table 7.32 – Success rate and coefficient of variation obtained with the identification algorithms (with Layout 3)
Considering the quality of the results achieved, the same damage scenarios defined in section 7.3.6.6
were also tested for this layout solution (see Table 7.21).
346
Chapter 7
The control charts obtained for the three damage scenarios are exposed in Figure 7.103. It is visible
that the damage events are clearly identified, with results in line with the ones presented for a layout
solution based on all sensors.
Undamaged scenario D1
300 300
250 250
200 200
T2
T2
150 150
100 100
50 50
0 0
0 0 50 100 150
Time Time
D2 D3
250 200
200
150
150
T2
T2
100
100
50
50
0 0
0
Time Time
Figure 7.103 – control charts associated with a scenario where no damage is present and with the three defined
damage scenarios (Layout 3)
Fatigue Results
Given the distribution of sensors in Layout 3 and the good results achieved with the modal tracking, it
is expected that the estimation of the acceleration time series at unmeasured locations would be also
good. Just like in the Layout 2, the frequency cut imposed for Layout 1 was not implemented in this
solution.
The good estimation of the acceleration time series at unmeasured locations can be attested with
Figure 7.104, referred to an event during a production period (with a mean rotor speed of 14.1 RPM).
In this figure, the estimated accelerations at measured (+74.988 m and +48.392 m) and unmeasured
(+21.772) positions in the FA direction are compared with the recorded time series by the sensors at
the same positions. For this example, the methodology based on the SSI-DATA algorithm was used. It
is visible that the estimated acceleration time series at the unmeasured position is very good.
347
Structural Monitoring of Wind Turbines
0.05 0.05
Acc. [m/s2]
Acc. [m/s2]
0 0
-0.05 -0.05
-0.1 -0.1
0 100 200 300 400 500 600 100 105 110 115 120 125
Time [s] Time [s]
0.1 0.1
Acc. [m/s2]
Acc. [m/s2]
0 0
-0.1 -0.1
-0.2 -0.2
0 100 200 300 400 500 600 100 105 110 115 120 125
Time [s] Time [s]
0.05 0.05
Acc. [m/s2]
Acc. [m/s2]
0 0
-0.05 -0.05
-0.1 -0.1
0 100 200 300 400 500 600 100 105 110 115 120 125
Time [s] Time [s]
Figure 7.104 – Original and estimated acceleration time series at the measured levels (+74.988 m and +48.392 m) and
unmeasured levels (+21.772 m)
Figure 7.105 evidences the evolution of the number of events with the increase of the two considered
quality indices ( and ) referred to the 4 used sensors. The results obtained with this layout
solution are similar to the ones achieved with all sensors.
348
Chapter 7
100 100
Percentage of considered events [%]
60 60
40 40
SSI-DATA SSI-DATA
20 20
SSI-COV SSI-COV
p-LSCF p-LSCF
0 0
0 0.1 0.2 0.3 0.4 0.5 0 0.5 1 1.5 2 2.5 3
Quality index Δe Quality index Δfreq
a) b)
Figure 7.105 – Percentage of considered events with Layout 3 with the evolution of: a) ∆ index; b) ∆ index
The deviation associated with the direct comparison of the damage estimated with each methodology
(using Layout 3) and the damage estimated with the reference methodology (SSI-DATA algorithm,
using all sensors) for coincident events was again used to assess the quality of the results. The
deviations are presented in Table 7.33. It is seen that results obtained with the SSI-DATA algorithm
represent a very good agreement with the reference results, also showing a good success rate. Both the
other two algorithms show lower success rates. Notwithstanding, they both show a good agreement of
the damage estimation with the reference results, especially the methodology based on p-LSCF
algorithm.
30 0.12
1 year 1 year
SSI-DATA Layout 3
25 0.1
Percentage of occurrence [%]
SSI-COV
p-LSCF
20 0.08
15 0.06
10 0.04
5 0.02
0 0
0 5 10 15 20 25 0 5 10 15 20 25
Wind speed [m/s] Wind speed [m/s]
a) b)
Figure 7.106 – Probability density function of wind speed (in the sector 100º-120º) recorded by the SCADA system and
from the data considered for fatigue estimation with Layout 3 with: a) the thee algorithms; b) best estimations from the
thee algorithms
349
Structural Monitoring of Wind Turbines
Table 7.33 – Deviation of accumulated damage of coincident events between results from Layout 3 the methodology
based on SSI-DATA algorithm using all sensors
Using the best estimations provided by the three algorithms to increase the range of damage scenarios
included in the analysis, it is possible to extrapolate the damage expected at the end of the design
fatigue life according to the measured probability density function (PDF) of the mean wind speed.
Figure 7.106 compares the PDF of the mean wind speed measured for the sector 100º-120º for the
acceleration events considered when the best estimations from the three algorithms are used with the
PDF obtained with the data recorded by the SCADA system during one year. The results obtained with
the best estimation were then adjusted to follow the PDF of the mean wind speed obtained from the
SCADA system. The results obtained are summarized in Table 7.34. It is visible that the estimation
obtained with this layout solution is very similar to the estimated using all sensors (deviations smaller
than 7 %), while a good success rate is also achieved.
Table 7.34 – Estimation of accumulated damage at the end of design life considering an adjusted extrapolation according
to the PDF of the wind speed using the best estimations from the three algorithms
7.3.9 CONCLUSIONS
This section presented a complete analysis of the installation and implementation of a vibration-based
structural health monitoring system in a 2.0 MW wind turbine. It starts with the presentation of the
main characteristics of the wind turbine, together with a description of its operating conditions.
Alongside, the modal results obtained with two developed finite element models are presented, serving
for an initial assessment of the wind turbine vibration modes.
Then, the dynamic behaviour of the wind turbine structure under operational and parked conditions
was analysed. The application of the implemented algorithms (SSI-COV, SSI-DATA and p-LSCF) for
modal identification was successfully employed, allowing to identify 9 vibration modes, among tower
and rotor blades modes. The continuous assessment of the wind turbine dynamic properties permitted
to completely characterize its modal parameters throughout the different operating regimes. The
developed methodology allowed to remove the influence of the harmonics in the identification of the
most important vibration modes.
The analysis of the collected data during more than one year permitted to identify 51 data sets
corresponding to free decay events. These acceleration events were analysed using two different
350
Chapter 7
methodologies: one based on the adjustment of an exponential window and another based on the SSI-
COV algorithm. The results obtained with both methods presented very coherent outcomes.
Nevertheless, the values obtained for the damping of the 1 FA mode showed some variability which
could not be fully explained.
An analysis about the relative contribution of the vibration modes and rotor harmonics to the
measured acceleration was also performed following the procedure described in section 5.5. It was
noticed that the participation of the modes/ harmonics on the measured signal is clearly dependent on
the operating conditions and on the tower section under analysis.
After this, the identified natural frequency values of the tracked vibration modes were used to assess
the influence of operational and environmental effects. In that sense, static and dynamic models were
tested. Among these, the static model SM2 provided the best results, demonstrating the ability of the
implemented methodology to minimize the influence of these external effects on the measured natural
frequencies.
Using the outcomes obtained with the application of the regression model to the identified natural
frequencies, the ability of the developed methodology to detect abnormal structural changes was
tested. It was demonstrated the accuracy of the system to identify small variations of the tower bending
modes along the yaw angle. This represents an important achievement since it allows to identify
possible asymmetric conditions of the foundation stiffness. In addition, the methodology also proved
being able to detect frequency variations associated with small damages on the foundation of onshore
and offshore wind turbines and, although with a smaller precision, on the blades.
The fatigue condition of the wind turbine tower was also assessed following the procedure introduced
in section 6.4.1.1. Due to the impossibility to install sensors to quantify the quasi-static motion of the
tower, it was only possible to estimate the condition of the structure due to the dynamic component of
fatigue. In that sense, the considerations made in this context only intended to validate the proposed
methodology and did not aim to analyse the real fatigue condition of the structure. It was
demonstrated that, for this scenario of study, the methodology based on the SSI-DATA and SSI-COV
algorithms delivered the best results. A strategy based on the use of the best estimation from the three
algorithms was also tested with good results.
Finally, an optimization of the dynamic monitoring system considering different sensors layout was
proposed and tested. The first solution, based on a biaxial sensor at the tower top, revealed some
deficiencies on the identification of some modes. Even though, it has proved successful in the
identification of small damages (although without the precision of the solution based on all sensors).
On the other hand, this layout performed very poorly for the fatigue assessment of the tower. Both
layout solutions 2 and 3, based on one and two biaxial sensors, respectively, evidenced the good ability
for detection of damage at an early stage and for fatigue assessment of wind turbine towers.
In short, this section proved the suitability of the developed vibration-based monitoring system to be
implemented in wind turbines. The usefulness of the damage detection module was clearly
demonstrated, while the component regarding the fatigue assessment of the tower is still at a
development stage, requiring additional testing. Nevertheless, the results already obtained permit to
anticipate the potential of this solution.
351
Structural Monitoring of Wind Turbines
352
Chapter 8
8
CONCLUSIONS AND FUTURE
RESEARCH
8.1 CONCLUSIONS
This work explored the potential of using operational modal analysis in the monitoring of wind
turbines. In this scope, several processing routines were developed and implemented to extract the
maximum information that a monitoring system based on the dynamic response of wind turbines can
deliver. The developed methodology was implemented into a vibration-based monitoring system,
capable of operating continuously and in an automated way, that transforms the collected data into
useful information for the wind farm owners.
The thesis starts with a global perspective of the wind energy industry. An overview of the main
technical evolutions of wind turbines was presented, focusing on two crucial aspects that defined the
success of some models: reliability and cost effectiveness. Then, a description of the main structural
elements of wind turbines is made, with a special emphasis on their reliability and cost. A brief
description of the main theories describing the performance of wind turbines is also introduced,
aiming to facilitate the comprehension of the response of the turbine under different operating
conditions.
Then, the focus of the thesis is orientated to the analysis and characterization of the dynamic
behaviour of wind turbines. This characterization was helpful to study the applicability of the OMA
techniques, used in the context of the work (SSI-DATA, SSI-COV and p-LSCF), to wind turbines.
Several methodologies developed and implemented in the context of the dynamic monitoring system
are also described:
o Automated procedure for operational modal analysis;
o Computation of the relative contribution of the vibration modes and harmonics to the
measured acceleration response;
o Statistical tools for detection of damage;
o Fatigue assessment of the support structure using a reduced number of sensors.
353
Conclusions and Future Research
The processing tools developed in this thesis were validated using three case studies. The first case
study comprises the analysis of vibration data collected during two distinct periods on an Izar Bonus
1.3MW/62 wind turbine. With this study, some important achievements should be referred:
o The ability of the SSI-COV and p-LSCF algorithms to accurately identify the main tower and
rotor blades vibration modes of the Izar Bonus 1.3MW/62 wind turbine using both
accelerometers at the tower and fibre Bragg grating sensors at the blades was demonstrated;
o Analysis of the data collected before and after the Rotor Blade Extension was used to
demonstrate the precision of the implemented modal identification algorithms to detect small
variations of frequency, being able to detect the frequency shift associated with an increase of 1
% of the blades mass.
The second case study is an offshore Vestas V90-3.0MW wind turbine. For this study, a short period of
data collected with a dynamic monitoring system was considered. The main conclusions regarding this
analysis were:
o The potential of the SSI-COV and p-LSCF algorithms to identify the most important vibration
modes of an offshore wind turbine was demonstrated for operating and non-operating
conditions;
o The developed methodology for automated identification of the modal properties of the wind
turbine was successfully employed, giving a good indication about the possibility of its
application to generic onshore and offshore wind turbines.
Lastly, the third case is the most complete study of this thesis. It comprises the analysis of the
acceleration data collected during a period of one year at a Senvion MM82 wind turbine. The major
achievements obtained with this analysis are:
o The methodology developed for continuous assessment of the modal properties was
successfully tested in the Senvion MM82 wind turbine. With this analysis, the suitability of this
methodology for identification of the properties of the main vibration modes under several
operating conditions was demonstrated;
o The analysis of the relative contribution of the vibration modes and harmonics for the
acceleration response of the wind turbine showed the influence of the varying operating
conditions on the dynamic behaviour of the turbine. Depending on the tower section under
analysis and on the conditions of operation, the acceleration response may be dominated by
the structural response due to the harmonic excitation 3Ω or by the 1 FA or 2 FA vibration
modes;
o The use of a regression model demonstrated to be a suitable solution to minimize the
operational and environmental effects. A model composed by individual static regression
models for each operating regime proved to deliver the best results;
o The developed monitoring methodology showed its ability to detect small variations of the
natural frequencies, associated with damages at a very early stage on foundation of onshore
and offshore turbines. A lower precision was achieved in the detection of blades damage;
354
Chapter 8
o The proposed procedure for assessment of the wind turbine tower due to the dynamic
component of fatigue was validated. Additional testing is required to extract meaningful
results from this methodology;
o An optimization of the dynamic monitoring system considering a reduction of the number of
sensors was tested. It was concluded that significant results can be obtained for detection of
damage using only one biaxial sensor at the tower top. However, in order to achieve results for
both damage detection and fatigue assessment with a precision similar to the ones obtained
when all sensors are considered, a solution based on one biaxial sensor installed on a
convenient position or a solution based on two biaxial sensors are required. This last solution
may, in fact, only require the installation of one additional sensor if the sensor incorporated in
turbine control system can be used.
355
Conclusions and Future Research
356
Appendix A
APPENDIX A
MODEL RESPONSE ESTIMATION -
VALIDATION
A.1. INTRODUCTION
In order to validate the concepts introduced in section 5.5, related to the decomposition of the
recorded acceleration signal into several modal (and harmonic) responses, a simple example is
presented in this section. In that sense, a simple small-scale model of a two-floor frame structure is
used. The analysis is conducted using only one direction, as shown in Figure A.1 a). The stiffness ,
mass and damping matrices of the structural system are introduced in Figure A.1 b).
f2(t) y2(t)
14.655 7.328
/
7.328 7.328
0.0046 0.0023
. /
0.0023 0.0023
a) b)
Figure A.1 – a) Illustration of the two-floor frame structure; b) Structural matrices of the model
With the previously introduced matrices, two vibration modes with well separated natural frequency
values are obtained. The natural frequency and modal damping ratio values of the system as presented
in Figure A.2, as well as illustrations of the mode shapes.
357
Model Response Estimation - Validation
6.336 16.587
0.618 % 1.618 %
Table A.1 – Input parameters of the different algorithms used for modal identification and modal decomposition of
the signal
Modal Nº of points of
100 2. 1 257 512
Identification the Half-spectra
358
Appendix A
The results obtained with the application of the modal response post-processing tool to the output
acceleration time series from the first floor are presented in Figure A.3. In this figure, the simulated
acceleration is shown at the top, while the modal responses corresponding to the 1st and 2nd mode are
illustrated in the middle and bottom plots. It is possible to attest that the estimated modal responses
obtained with the different algorithms are almost coincident.
Level 1
0.3
0.2
0.1
Acc. [m/s2]
-0.1
-0.2
-0.3
0 100 200 300 400 500 600
Time [s]
Level 1 - Mode 1
0.15
SSI-DATA
0.1 SSI-COV
p-LSCF
0.05
Acc. [m/s2]
-0.05
-0.1
-0.15
0 100 200 300 400 500 600
Time [s]
Level 1 - Mode 2
0.2 SSI-DATA
SSI-COV
p-LSCF
0.1
Acc. [m/s2]
-0.1
-0.2
Figure A.3 – Output acceleration time series at level 1 (top) and estimated responses for the 1st and 2nd vibration modes
obtained with the different algorithms during the whole period of simulation
359
Model Response Estimation - Validation
Aiming to highlight the good initial impression given by Figure A.3, Figure A.4 presents the same
results from the previous figure for a shorter time length (10 seconds). Again, the top plot refers to the
numerically obtained acceleration, while the other two shows the modal responses from to the 1st and
2nd modes obtained with the different algorithms. In this figure, the very good agreement between the
modal responses is evident.
Level 1
0.3
0.2
0.1
Acc. [m/s2]
-0.1
-0.2
-0.3
380 381 382 383 384 385 386 387 388 389 390
Time [s]
Level 1 - Mode 1
0.15
SSI-DATA
0.1 SSI-COV
p-LSCF
0.05
Acc. [m/s2]
-0.05
-0.1
-0.15
380 381 382 383 384 385 386 387 388 389 390
Time [s]
Level 1 - Mode 2
0.15
SSI-DATA
0.1 SSI-COV
p-LSCF
0.05
Acc. [m/s2]
-0.05
-0.1
-0.15
380 381 382 383 384 385 386 387 388 389 390
Time [s]
Figure A.4 – Output acceleration time series at level 1 (top) and estimated responses for the 1st and 2nd vibration modes
obtained with the different algorithms during a period of 10 seconds
360
Appendix A
The analysis previously presented for the results achieved for the first floor of the structural system was
also performed with the acceleration output of the second floor.
Level 2
0.3
0.2
0.1
Acc. [m/s2]
-0.1
-0.2
-0.3
0 100 200 300 400 500 600
Time [s]
Level 2 - Mode 1
0.2 SSI-DATA
SSI-COV
p-LSCF
0.1
Acc. [m/s2]
-0.1
-0.2
-0.05
-0.1
-0.15
0 100 200 300 400 500 600
Time [s]
Figure A.5 – Output acceleration time series at level 2 (top) and estimated responses for the 1st and 2nd vibration modes
obtained with the different algorithms during the whole period of simulation
361
Model Response Estimation - Validation
A detailed representation of the results is introduced in Figure A.6. In this figure, the good agreement
obtained for the modal responses is clearly noticed.
Level 2
0.3
0.2
0.1
Acc. [m/s2]
-0.1
-0.2
-0.3
380 381 382 383 384 385 386 387 388 389 390
Time [s]
Level 2 - Mode 1
0.2 SSI-DATA
SSI-COV
p-LSCF
0.1
Acc. [m/s2]
-0.1
-0.2
380 381 382 383 384 385 386 387 388 389 390
Time [s]
Level 2 - Mode 2
0.15
SSI-DATA
0.1 SSI-COV
p-LSCF
0.05
Acc. [m/s2]
-0.05
-0.1
-0.15
380 381 382 383 384 385 386 387 388 389 390
Time [s]
Figure A.6 – Output acceleration time series at level 2 (top) and estimated responses for the 1st and 2nd vibration modes
obtained with the different algorithms during a period of 10 seconds
362
Appendix B
APPENDIX B
IZAR BONUS 1.3MW/62 WIND
TURBINE – MODE SHAPES
B.1. INTRODUCTION
In this annex the mode shapes of the identified vibration modes of the Izar Bonus 1.3MW/62 wind
turbine are presented.
݂ଵ = 0.45 Hz
70 70
30
60 60
20
50 50
10
SS direction
40 40
0
30 30
-10
20 20
-20
10 10
-30
0 0
-20 0 20 -20 0 20 -30 -20 -10 0 10 20 30
FA direction SS direction FA direction
363
Izar Bonus 1.3MW/62 Wind Turbine – Mode Shapes
݂ଶ = 0.46 Hz
70 70
30
60 60
20
50 50
10
SS direction
40 40
0
30 30
-10
20 20
-20
10 10
-30
0 0
-20 0 20 -20 0 20 -30 -20 -10 0 10 20 30
FA direction SS direction FA direction
݂ଷ = 0.53 Hz
70 70
30
60 60
20
50 50
10
SS direction
40 40
0
30 30
-10
20 20
-20
10 10
-30
0 0
-20 0 20 -20 0 20 -30 -20 -10 0 10 20 30
FA direction SS direction FA direction
݂ସ = 1.13 Hz
70 70
30
60 60
20
50 50
10
40
SS direction
40
0
30 30
-10
20 20
-20
10 10
-30
0 0
-20 0 20 -20 0 20 -30 -20 -10 0 10 20 30
FA direction SS direction FA direction
364
Appendix B
݂ହ = 1.14 Hz
70 70
30
60 60
20
50 50
10
40 40
SS direction
0
30 30
-10
20 20
-20
10 10
-30
0 0
-20 0 20 -20 0 20 -30 -20 -10 0 10 20 30
FA direction SS direction FA direction
݂ = 1.32 Hz
70 70
30
60 60
20
50 50
10
SS direction
40 40
0
30 30
-10
20 20
-20
10 10
-30
0 0
-20 0 20 -20 0 20 -30 -20 -10 0 10 20 30
FA direction SS direction FA direction
݂ = 1.47 Hz
70 70
30
60 60
20
50 50
10
40 40
SS direction
30 30
-10
20 20
-20
10 10
-30
0 0
-20 0 20 -20 0 20 -30 -20 -10 0 10 20 30
FA direction SS direction FA direction
365
Izar Bonus 1.3MW/62 Wind Turbine – Mode Shapes
଼݂ = 2.02 Hz
70 70
30
60 60
20
50 50
10
SS direction
40 40
0
30 30
-10
20 20
-20
10 10
-30
0 0
-20 0 20 -20 0 20 -30 -20 -10 0 10 20 30
FA direction SS direction FA direction
݂ଽ = 2.06 Hz
70 70
30
60 60
20
50 50
10
40 40
SS direction
30 30
-10
20 20
-20
10 10
-30
0 0
-20 0 20 -20 0 20 -30 -20 -10 0 10 20 30
FA direction SS direction FA direction
݂ଵ = 3.33 Hz
70 70
30
60 60
20
50 50
10
40 40
SS direction
0
30 30
-10
20 20
-20
10 10
-30
0 0
-20 0 20 -20 0 20 -30 -20 -10 0 10 20 30
FA direction SS direction FA direction
366
Appendix B
݂ଵଵ = 3.45 Hz
70 70
30
60 60
20
50 50
10
40 40
SS direction
0
30 30
-10
20 20
-20
10 10
-30
0 0
-20 0 20 -20 0 20 -30 -20 -10 0 10 20 30
FA direction SS direction FA direction
݂ଵଶ = 3.74 Hz
70 70
30
60 60
20
50 50
10
40
SS direction
40
0
30 30
-10
20 20
-20
10 10
-30
0 0
-20 0 20 -20 0 20 -30 -20 -10 0 10 20 30
FA direction SS direction FA direction
݂ଵଷ = 4.03 Hz
70 70
30
60 60
20
50 50
10
SS direction
40 40
0
30 30
-10
20 20
-20
10 10
-30
0 0
-20 0 20 -20 0 20 -30 -20 -10 0 10 20 30
FA direction SS direction FA direction
367
Izar Bonus 1.3MW/62 Wind Turbine – Mode Shapes
݂ଵସ = 4.86 Hz
70 70
30
60 60
20
50 50
10
40 40
SS direction
0
30 30
-10
20 20
-20
10 10
-30
0 0
-20 0 20 -20 0 20 -30 -20 -10 0 10 20 30
FA direction SS direction FA direction
Figure B.1 – Mode shapes of the vibration modes identified with the measurement system based on accelerometers
݂ଷ = 0.61 Hz ݂ଷ = 0.59 Hz
Mode 3 Mode 3
Blade A Blade B Blade C Blade A Blade B Blade C
1 1
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0 0
-0.2 -0.2
-0.4 -0.4
-0.6 -0.6
-0.8 -0.8
-1 -1
E A C K E A C K E A C K E A C K E A C K E A C K
Flapwise Flapwise Flapwise Flapwise Flapwise Flapwise
Edgewise Edgewise Edgewise Edgewise Edgewise Edgewise
݂ସ = 1.13 Hz ݂ସ = 1.07 Hz
368
Appendix B
Mode 4 Mode 4
Blade A Blade B Blade C Blade A Blade B Blade C
1 1
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0 0
-0.2 -0.2
-0.4 -0.4
-0.6 -0.6
-0.8 -0.8
-1 -1
E A C K E A C K E A C K E A C K E A C K E A C K
Flapwise Flapwise Flapwise Flapwise Flapwise Flapwise
Edgewise Edgewise Edgewise Edgewise Edgewise Edgewise
݂ହ = 1.15 Hz ݂ହ = 1.10 Hz
Mode 5 Mode 5
Blade A Blade B Blade C Blade A Blade B Blade C
1 1
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0 0
-0.2 -0.2
-0.4 -0.4
-0.6 -0.6
-0.8 -0.8
-1 -1
E A C K E A C K E A C K E A C K E A C K E A C K
Flapwise Flapwise Flapwise Flapwise Flapwise Flapwise
Edgewise Edgewise Edgewise Edgewise Edgewise Edgewise
݂ = 1.33 Hz ݂ = 1.29 Hz
Mode 6 Mode 6
Blade A Blade B Blade C Blade A Blade B Blade C
1 1
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0 0
-0.2 -0.2
-0.4 -0.4
-0.6 -0.6
-0.8 -0.8
-1 -1
E A C K E A C K E A C K E A C K E A C K E A C K
Flapwise Flapwise Flapwise Flapwise Flapwise Flapwise
Edgewise Edgewise Edgewise Edgewise Edgewise Edgewise
݂ = 1.45 Hz ݂ = 1.43 Hz
369
Izar Bonus 1.3MW/62 Wind Turbine – Mode Shapes
Mode 7 Mode 7
Blade A Blade B Blade C Blade A Blade B Blade C
1 1
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0 0
-0.2 -0.2
-0.4 -0.4
-0.6 -0.6
-0.8 -0.8
-1 -1
E A C K E A C K E A C K E A C K E A C K E A C K
Flapwise Flapwise Flapwise Flapwise Flapwise Flapwise
Edgewise Edgewise Edgewise Edgewise Edgewise Edgewise
଼݂ = 2.04 Hz ଼݂ = 1.95 Hz
Mode 8 Mode 8
Blade A Blade B Blade C Blade A Blade B Blade C
1 1
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0 0
-0.2 -0.2
-0.4 -0.4
-0.6 -0.6
-0.8 -0.8
-1 -1
E A C K E A C K E A C K E A C K E A C K E A C K
Flapwise Flapwise Flapwise Flapwise Flapwise Flapwise
Edgewise Edgewise Edgewise Edgewise Edgewise Edgewise
Figure B.2 – Comparison of the amplitude of the modal curvature of the first 6 identified rotor related modes before and after
the RBE
370
References
REFERENCES
Ackermann, T. and L. Söder (2002). An overview of wind energy-status 2002. Renewable and
Sustainable Energy Reviews 6(1–2): 67-127.
Adams, D., J. White, M. Rumsey and C. Farrar (2011). Structural health monitoring of wind turbines:
method and application to a HAWT. Wind Energy 14.
Aenlle, M. L., L. Hermanns, P. Fernández and A. Fraile (2013). Stress Estimation in a Scale Model of a
Symmetric Two Story Building. 5th International Operational Modal Analysis Conference (IOMAC).
Guimarães, Portugal.
Aenlle, M. L., A. Skafte, P. Fernández and R. Brincker (2013). Strain Estimation in a Glass Beam using
Operational Modal Analysis. XXXI International Modal Analysis Conference (IMAC). Florida, USA.
Ameri, N., C. Grappasonni, G. Coppotelli and D. J. Ewins (2013). Ground vibration tests of a helicopter
structure using OMA techniques. Mechanical Systems and Signal Processing(35): 35-51.
Andersen, L. (2010). Assessment of lumped-parameter models for rigid footings. Computers &
Structures 88(23–24): 1333-1347.
Andersen, P., R. Brincker, C. Ventura and R. Cantieni (2007). Estimating Modal Parameters of Civil
Engineering Structures subject to Ambient and Harmonic Excitation. Experimental Vibration Analysis
for Civil Engineering Structures (EVACES). Porto, Portugal.
Anon (1890). Mr. Brush's Windmill Dynamo. Scientific American Vol. LXIII(No. 25): cover and p.
389.
Appleyard, D. (2013). Wind Insurance Claim Analysis Reveals Need for Monitoring. 2013, from
http://www.renewableenergyworld.com/rea/news/article/2013/08/wind-insurance-claim-analysis-
reveals-monitoring-need.
371
References
Arántegui, R. L., T. Corsatea and K. Suomalainen (2013). 2012 JRC Wind Status Report, European
Commission. Joint Research Institute (JRC). Institute for Energy and Transport.
Arrigan, J., V. Pakrashi, B. Basu and S. Nagarajaiah (2011). Control of flapwise vibrations in wind
turbine blades using semi-active tuned mass dampers. Structural Control and Health Monitoring 18:
840-851.
Bak, C., P. Fuglsang, J. Johansen and I. Antoniou (2000). Wind tunnel tests of the NACA 63-415 and a
modified NACA 63-415 airfoil. Technical Report Risø–R–1193. Roskilde, Denmark, Risø National
Laboratory.
BARD Engineering GmbH. BARD - Wind Offshore Projects. 2013, from http://www.bard-
offshore.de/en/projects.html.
Becker, E. and P. Poste (2006). Keeping the blades turning: Condition monitoring of wind turbine gears.
Refocus 7(2): 26-32.
Belwind Offshore Energy (2009). Belwind - Wind farm in the North Sea (Press information).
Betz, A. (1920). Das Maximum der theoretisch möglichen Ausnutzung des Windes durch Windmotoren
(The maximum of the theoretically possible exploitation of the wind by wind motors). Zeitschrift f. d.
gesamte Turbinenwesen 17.
Betz, A. (1926). Windenergie und Ihre Ausnutzung durch Windmühlen,. Göttingen, Germany,
Vandenhoeck.
Bir, G. (2008). Multiblade Coordinate Transformation and Its Application to Wind Turbine Analysis.
ASME Wind Energy Symposium. Nevada, USA.
Blanco, M. I. (2009). The economics of wind energy. Renewable and Sustainable Energy Reviews 13(6–
7): 1372-1382.
Bloomberg New Energy Finance (2015). Sustainable Energy in America: 2015 Factbook. London, UK.
Blough, J. R. (2006). Adaptive Resampling - Transforming From the Time to the Angle Domain. XXIV
International Modal Analysis Conference (IMAC). Missouri, USA.
372
References
Bogert, B. P., M. J. R. Healy and J. W. Tukey (1963). The Quefrency Alanysis of Time Series for Echoes:
Cepstrum, Pseudo-Autocovariance, Cross-Cepstrum, and Saphe Cracking. Time Series Analysis(15):
209-243.
Boller, C., F.-K. Chang and Y. Fujino (2009). Encyclopedia of Structural Health Monitoring, John Wiley
& Sons.
Braam, H., T. S. Obdam and T. W. Verbruggen (2012). Low Cost Load Monitoring for Offshore Wind
Farms. International Conference on Noise and Vibration Engineering (ISMA). Leuven, Belgium.
Brincker, R., P. Andersen and N. Møller (2000). Indicator for Separation of Structural and Harmonic
Modes in Output-only Modal Testing. XVIII International Modal Analysis Conference (IMAC). San
Antonio, USA.
Brownjohn, J. M. W., F. Magalhães, E. Caetano and Á. Cunha (2010). Ambient vibration re-testing and
operational modal analysis of the Humber Bridge. Engineering Structures(32): 2003-2018.
Burton, T., D. Sharpe, N. Jenkins and E. Bossanyi (2001). Wind Energy Handbook, John Wiley & Sons.
Butterfield, C. P., W. Musial and J. Jonkman (2007). Overview of Offshore Wind Technology, National
Renewable Energy Laboratory (NREL).
Butterfield, S., W. Musial and J. Jonkman (2005). Engineering Challenges for Floating Offshore Wind
Turbines, National Renewable Energy Laboratory (NREL).
Byrne, B., G. Houlsby, C. Martin and P. Fish (2002). Suction Caisson Foundations for Offshore Wind
Turbines. Wiind Engineering 26(3): 145-155.
Caetano, E., F. Magalhães, Á. Cunha, O. Flamand and G. Grillaud (2007). Comparison of Sochastic
Identification Methods Applied to the Natural Response of Millau Viaduct. Experimental Vibration
Analysis for Civil Engineering Structures (EVACES). Porto, Portugal.
Cara, F. J., J. Juan, E. Alarcón, E. Reynders and G. De Roeck (2013). Modal contribution and state space
order selection in operational modal analysis. Mechanical Systems and Signal Processing 38(2): 276-
298.
373
References
Carbon Trust. (2013). Successful installation of first suction bucket foundation at Dogger Bank. 2013,
from http://www.carbontrust.com/news/2013/02/successful-installation-of-first-suction-bucket-
foundation-at-dogger-bank.
Carcangiu, C. E., D. Tcherniak, S. Chauhan, J. Basurko and M. Rossetti (2012). Numerical and
Experimental Modal Characterization of a 3MW Wind Turbine. EWEA 2012 Annual Event.
Copenhagen, Denmark.
Carden, E. P. and P. Fanning (2009). Vibration Based Condition Monitoring: A Review. Structural
Health Monitoring(4): 355-377.
Carne, T. G. and G. H. James Iii (2010). The inception of OMA in the development of modal testing
technology for wind turbines. Mechanical Systems and Signal Processing 24(5): 1213-1226.
Carne, T. G., J. P. Lauffer and A. J. Gomez (1988). Modal Testing the EOLE. New Mexico, USA, Sandia
National Laboratories.
Carne, T. G., D. W. Lobitz, A. R. Nord and R. A. Watson (1982). Finite Element Analysis and Modal
Testing of a Rotating Wind Turbine. New Mexico, USA, Sandia National Laboratories.
Carne, T. G. and A. R. Nord (1983). Modal Testing of a Rotating Wind Turbine. New Mexico, USA,
Sandia National Laboratories.
Cauberghe, B. (2004). Applied frequency-domain system identification in the field of experimental and
operational modal analysis. PhD thesis, V.U.Brussel.
Chauhan, S., D. Tcherniak, J. Basurko, O. Salgado, I. Urresti, C. E. Carcangiu and M. Rossetti (2011).
Operational Modal Analysis of Operating Wind Turbines: Application to Measured Data. IXXX
International Modal Analysis Conference (IMAC). Florida, USA.
Chou, J.-S. and W.-T. Tu (2011). Failure analysis and risk management of a collapsed large wind
turbine tower. Engineering Failure Analysis 18(1): 295-313.
Ciang, C. C., J.-R. Lee and H.-J. Bang (2008). Structural health monitoring for a wind turbine system: a
review of damage detection methods. Measurement Science and Technology 19: 1-20.
Clough, R. W. and J. Penzien (1995). Dynamics of Structures. Berkeley, USA, Computers & Structures,
Inc.
Collins, J. L., R. K. Shaltens, R. H. Poor and R. S. Barton (1982). Experience and Assessment of the
DOE/NASA MOD-1 2000 kW Wind Turbine Generator at Boorne, North Carolina.
374
References
Colwell, S. and B. Basu (2009). Tuned liquid column dampers in offshore wind turbines for structural
control. Engineering Structures 31(2): 358-368.
Connel, J. R. (1980). Turbulence spectrum observed by a fast-rotating wind turbine blade, Battelle
Pacific Northwest Laboratory.
Corrigan, R. D. and C. B. F. Ernsworth (1986). Design and Initial Testing of a One Bladed 30-Meter
Diameter Rotor on the NASA-DOE Mod-0 Wind Turbine. DOE/NASA/20320-70 NASA TM-88810.
Cunha, Á. and E. Caetano (2006). Experimental Modal Analysis of Civil Engineering Structures. Sound
and Vibration(40): 12-20.
Cunha, Á., E. Caetano and R. Delgado (2001). Dynamic Tests on a Large Cable-Stayed Bridge. An
Efficient Approach. Journal of Bridge Engineering 6(1): 54-62.
Currie, M., M. Saafi, C. Tachtatzis and F. Quail (2015). Structural integrity monitoring of onshore wind
turbine concrete foundations. Renewable Energy 83: 1131-1138.
Damgaard, M., L. B. Ibsen, L. V. Andersen and J. K. F. Andersen (2013a). Cross-wind modal properties
of offshore wind turbines identified by full scale testing. Journal of Wind Engineering and Industrial
Aerodynamics 116: 94-108.
Darrieus, G. J. (1931). Turbine Having Its Rotating Shaft Transverse to the Flow of The Current. U. S. P.
Offcie.
Department of Energy & Climate Change (2010). Value breakdown for the offshore wind sector.
Det Norske Veritas (DNV) (2002). Guidelines for Design of Wind Turbines.
Det Norske Veritas (DNV) (2011). DNV-OS-J101 - Design of Offshore Wind Turbine Structures.
Devriendt, C., F. Magalhães, W. Weijtjens, G. De Sitter, Á. Cunha and P. Guillaume (2014). Structural
health monitoring of offshore wind turbines using automated operational modal analysis. Structural
Health Monitoring 13(6): 644-659.
375
References
Devriendt, C., W. Weijtjens, M. El-Kafafy and G. D. Sitter (2014) Monitoring resonant frequencies and
damping values of an offshore wind turbine in parked conditions. IET Renewable Power Generation 8,
433-441.
Dong, W., T. Moan and Z. Gao (2012). Fatigue reliability analysis of the jacket support structure for
offshore wind turbine considering the effect of corrosion and inspection. Reliability Engineering &
System Safety 106: 11-27.
European Comitee for Standardization (CEN) (2005). EN 1993-1-9 Eurocode 3: Design of steel
structures - Part 1-9: Fatigue.
European Commission (2011). UpWind - Design limits and solutions for very large wind turbines, The
Sixth Framework Programme for Research and Development of the European Commission (FP6).
European Commission (2015c). The European Union leading in renewables. Brussels, Belgium.
European Wind Energy Association (EWEA) (2007). Powering Change - EWEA 2006 Annual Report.
Brussels, Belgium.
European Wind Energy Association (EWEA) (2008). Delivering Eenery and Climate Solutions - EWEA
2007 Annual Report. Brussels, Belgium.
European Wind Energy Association (EWEA) (2009). Winning with European Wind - EWEA 2008
annual report. Brussels, Belgium.
European Wind Energy Association (EWEA) (2010). Wind in Power - 2009 European statistics.
Brussels, Belgium.
European Wind Energy Association (EWEA) (2011a). The European offshore wind industry key trends
and statistics 2010. Brussels, Belgium.
376
References
European Wind Energy Association (EWEA) (2011b). Wind in Power - 2010 European statistics.
Brussels, Belgium.
European Wind Energy Association (EWEA) (2012a). The European offshore wind industry key 2011
trends and statistics. Brussels, Belgium.
European Wind Energy Association (EWEA) (2012b). Wind in Power - 2011 European Statistics.
Brussels, Belgium.
European Wind Energy Association (EWEA) (2013a). The European offshore wind industry - key
trends and statistics 2012. Brussels, Belgium.
European Wind Energy Association (EWEA) (2013b). Wind in Power - 2012 European Statistics.
Brussels, Belgium.
European Wind Energy Association (EWEA) (2014a). The European offshore wind industry - key
trends and statistics 2013. Brussels, Belgium.
European Wind Energy Association (EWEA) (2014b). Wind in power - 2013 European statistics.
Brussels, Belgium.
European Wind Energy Association (EWEA) (2015a). The European offshore wind industry - key trend
and statistics 2014.
European Wind Energy Association (EWEA) (2015b). Wind in power - 2014 European statistics.
Brussels, Belgium.
Ewins, D. J. (2000). Modal Testing: Theory, Practice and Application. Baldock, England, Research
Studies Press Ltd.
Fast, M. and T. Palmé (2010). Application of artificial neural networks to the condition monitoring and
diagnosis of a combined heat and power plant. Energy 35(2): 1114-1120.
Faulstich, S. (2010). Component Reliabilty Ranking with Respect to WT Concept and External
Environmental Conditions Project UpWind.
Felber, A. J. (1993). Development of a Hybrid Bridge Evaluation system. PhD thesis, The University of
British Columbia.
Freris, L. L. (1990). Wind Energy Conversion System. New York, USA, Prentice-Hall.
377
References
Fritzen, C.-P. (2005). Vibration-Based Structural Health Monitoring – Concepts and Applications. Key
Engineering Materials 293 - 294(3): 3 - 20.
García Márquez, F. P., A. M. Tobias, J. M. Pinar Pérez and M. Papaelias (2012). Condition monitoring
of wind turbines: Techniques and methods. Renewable Energy 46(0): 169-178.
Gasch, R. and J. Twele (2012). Wind Power Plants - Fundamentals, Design, Construction and
Operation, Springer.
Gellermann, T. (2012). Extension of the scope of Condition Monitoring Systems for multi-MW and
offshore wind turbines1. VDI-Conference Schwingungen von Windenergieanlagen 2012. Bremen,
Germany.
Gentile, C. and A. Saisi (2007). Ambient vibration testing of historic masonry towers for structural
identification and damage assessment. Construction and Building Materials(21): 1311-1321.
Germanischer Lloyd (GL) (2003). Guideline for the Certification of Condition Monitoring Systems for
Wind Turbines. Rules and Guidelines - IV Industrial Services. Hamburg, Germany.
Germanischer Lloyd (GL) (2005). Overall Damping for Piled Offshore Support Structures. Note on
Engineering Details.
Germanischer Lloyd (GL) (2010). Guideline for the Certification of Wind Turbines. Hamburg,
Germany.
Germanischer Lloyd (GL) (2012). Guideline for the Certification of Offshore Wind Turbines. Hamburg,
Germany.
Glauert, H. (1926). The Analysis of Experimental Results in the Windmill Brake and Vortex Ring States
of an Airscrew. Aeronautical Research Committee Reports and Memoranda.
Global Wind Energy Council (GWEC) (2007). Global Wind 2006 Report. Brussels, Belgium.
Global Wind Energy Council (GWEC) (2008). Global Wind 2007 Report. Brussels, Belgium.
Global Wind Energy Council (GWEC) (2009). Global Wind 2008 Report. Brussels, Belgium.
Global Wind Energy Council (GWEC) (2010). Global Wind 2009 Report. Brussels, Belgium.
Global Wind Energy Council (GWEC) (2011). Global Wind Report - Annual Market Update 2010.
Brussel, Belgium.
378
References
Global Wind Energy Council (GWEC) (2012). Global Wind Report - Annual market update 2011.
Brussel, Belgium.
Global Wind Energy Council (GWEC) (2013). Global Wind Report - Annual market update 2012.
Brussel, Belgium.
Global Wind Energy Council (GWEC) (2014). Global Wind Report - Annual Market Update 2013.
Brussels, Belgium.
Global Wind Energy Council (GWEC) (2015). Global Wind Report - Annual Market Update 2014.
Brussels, Belgium.
Gordon, L. H., J. S. Andrews and D. K. Zimmerman (1983). Mod-2 Wind Turbine Development.
DOE/NASN20305-9 NASATM-83460.
Goursat, M., M. Döhler, L. Mevel and P. Andersen (2010). Crystal Clear SSI for Operational Modal
Analysis of Aerospace Vehicles. XXVIII International Modal Analysis Conference (IMAC). Florida,
USA.
Griffith, D. T. and T. G. Carne (2010). Experimental Modal Analysis of 9-meter Research-sized Wind
Turbine Blades. XXVIII International Modal Analysis Conference (IMAC). Florida, USA.
Griffith, D. T., R. L. Mayes and P. S. Hunter (2010). Excitation Methods for a 60 kW Vertical Axis Wind
Turbine. XXVIII International Modal Analysis Conference (IMAC), Florida, USA.
Griffith, D. T., G. Smith, M. Casias, S. Reese and T. W. Simmermacher (2006). Modal Testing of the
TX-100 Wind Turbine Blade. New Mexico, USA, Sandia National Laboratories.
Griffith, D. T., N. C. Yoder, B. Resor, J. White and J. Paquette1 (2013). Structural health and
prognostics management for the enhancement of offshore wind turbine operations and maintenance
strategies. Wind Energy.
379
References
Guo, P. and D. Infield (2012). Wind Turbine Tower Vibration Modeling and Monitoring by the
Nonlinear State Estimation Technique (NSET). Energies 5: 5279-5293.
Hall, M., B. Buckham and C. Crawford (2013). Evaluating the importance of mooring line model fidelity
in floating offshore wind turbine simulations. Wind Energy: n/a-n/a.
Hameed, Z., Y. S. Hong, Y. M. Cho, S. H. Ahn and C. K. Song (2009). Condition monitoring and fault
detection of wind turbines and related algorithms: A review. Renewable and Sustainable Energy Reviews
13(1): 1-39.
Hansen, M. H. (2003). Improved Modal Dynamics of Wind Turbines to Avoid Stall-induced Vibrations.
Wind Energy 6: 179-195.
Hansen, M. H. (2007). Aeroelastic Instability Problems for Wind Turbines. Wind Energy 10(6): 551-
577.
Hansen, M. H., K. Thomsen and P. Fuglsang (2006). Two Methods for Estimating Aeroelastic Damping
of Operational Wind Turbine Modes from Experiments. Wind Energy(9): 179-191.
Harman, K., R. Walker and M. Wilkinson (2008). Availability trends observed at operational wind
farms. European Wind Energy Conference (EWEC). Brussels, Belgium.
HAWC2 web site. (2013). HAWC2 model - NREL 5-MW Reference Wind Turbine. from
http://www.hawc2.dk.
Henderson, A. R. (2003). Support Structures for Floating Offshore. Workshop on Deep Water Offshore
Wind Energy Systems.
380
References
Hermans, L. and H. V. d. Auweraer (1999). Modal Testing and Analysis of Structures Under
Operational Conditions: Industrial Applications. Mechanical Systems and Signal Processing 13(2): 193-
216.
Heylen, W., S. Lammens and P. Sas (2007). Modal Analysis Theory and Testing. Belgium, Katholieke
Universiteit Leuven.
Hjelm, H. P., R. Brincker, J. Graugaard-Jensen and K. Munch (2005). Determination of Stress Histories
in Structures by Natural Input Modal Analysis. XXIII International Modal Analysis Conference
(IMAC). Florida, USA.
Hoell, S. and P. Omenzetter (2014). Detection and localization of trailing edge disbond in a large wind
turbine blade. International Conference on Offshore Renewable Energy. Glasgow, UK.
Hu, W.-H. (2011). Operational Modal Analysis and Continuous Dynamic Monitoring of Footbridges.
PhD thesis, Faculdade de Engenharia da Universidade do Porto.
Hu, W.-H., E. Caetano and Á. Cunha Structural health monitoring of a stress-ribbon footbridge.
Engineering Structures 57.
Hu, W.-H., C. Moutinho, E. Caetano, F. Magalhães and Á. Cunha (2012). Continuous dynamic
monitoring of a lively footbridge for serviceability assessment and damage detection. Mechanical Systems
and Signal Processing 33: 38-55.
Hu, W.-H., S. Thöns, R. G. Rohrmann, S. Said and W. Rücker (2015a). Vibration-based structural
health monitoring of a wind turbine system Part II: Environmental/operational effects on dynamic
properties. Engineering Structures 89: 273-290.
Hu, W.-H., S. Thöns, R. G. Rohrmann, S. Said and W. Rücker (2015b). Vibration-based structural
health monitoring of a wind turbine system. Part I: Resonance phenomenon. Engineering Structures 89:
260-272.
Ibsen, L. B. and M. Liingaard (2006). Prototype bucket foundation for wind turbines - natural frequency
estimation. Aalborg, Denmark, Aalborg University - Department of Civil Engineering.
Ibsen, L. B., M. Liingaard and S. A. Nielsen (2005). Bucket Foundation, a status. Copenhagen Offshore
Wind, Copenhagen, Denmark.
381
References
Institute of Mechanical Engineering and Industrial Management (INEGI) and Portuguese Renewable
Energy Association (APREN) (2015). Wind Farms in Portugal - December 2014.
International Electrotechnical Commission (IEC) (2009). IEC 61400-3 Wind turbines – Part 3: Design
requirements for offshore wind turbines.
International Electrotechnical Commission (IEC) (2014). IEC 61400-1 Wind turbines –Part 1: Design
requirements.
International Energy Agency (IEA) (1984). International Energy Agency - Co-operation in the
Development of Large-Scale Wind Energy Systems (Annual Report of 1983), IEA LS WECS Executive
Commitee.
International Energy Agency (IEA) (1989). IEA Large-Scale Wind Energy - Annual Report 1988, IEA
LS WECS Execuiive Committee.
Jacobsen, N.-J., P. Andersen and R. Brincker (2006). Using Enhanced Frequency Domain
Decomposition as a Robust Technique to Harmonic Excitation in Operational Modal Analysis.
International Conference on Noise & Vibration Engineering 2006 (ISMA). Lueven, Belgium.
Jacobsen, N.-J., P. Andersen and R. Brincker (2008). Applications of Frequency Domain Curve-fitting in
the EFDD Technique. XXVI International Modal Analysis Conference (IMAC). Florida, USA.
James, G. H. (1994). Extraction of modal parameters from an operating HAWT using the Natural
Excitation Technique (NExT). 13th Energy-Sources Technology Conference and Exhibition (ETCE) on
Wind Energy. Los Angeles, USA.
James, G. H., T. G. Carne and J. P. Lauffer (1993). The Natural Excitation Technique (NExT) for Modal
Parameter Extraction From Operating Wind Turbines. New Mexico, USA, Sandia National
Laboratories.
Johnson, R. A. and D. W. Wichern (2002). Applied multivariate statistical analysis. New Jersey, USA,
Prentice Hall.
Jonkman, J., S. Butterfield, W. Musial and G. Scott (2009). Definition of a 5-MW Reference Wind
Turbine for Offshore System Development National Renewable Energy Laboratory.
Juang, J.-N. (1994). Applied System Identification. New Jersey, USA, Prentice Hall Englewood Cliffs.
382
References
Kármán, T. v. (1948). Progress in the statistical theory of turbulence. Proceedings of the National
Academy of Science of the United States of America 34: 530-539.
Kuhn, M. J. (2001). Dynamics and Design Optimisation of Offshore Wind Energy Conversion Systems.
PhD thesis, Delft University of Technology.
Kvittem, M. I. and T. Moan (2015). Time domain analysis procedures for fatigue assessment of a semi-
submersible wind turbine. Marine Structures 40: 38-59.
LeBlanc, C. and N. J. Tarp-Johansen (2011). Monopile in sand, stiffness and damping. Oral
Presentation, EWEA Conference. Brussels, Belgium.
Li, Z.-q., S.-j. Chen, H. Ma and T. Feng (2013). Design defect of wind turbine operating in typhoon
activity zone. Engineering Failure Analysis(27): 165-172.
Liu, W., B. Tang and Y. Jiang (2010a). Status and problems of wind turbine structural health monitoring
techniques in China. Renewable Energy 35(7): 1414-1418.
Liu, W., B. Tang and Y. Jiang (2010b). Status and problems of wind turbine structural health
monitoring techniques in China. Renewable Energy(35): 1414-1418.
Liu, W. Y., B. P. Tang, J. G. Han, X. N. Lu, N. N. Hu and Z. Z. He (2015). The structure healthy
condition monitoring and fault diagnosis methods in wind turbines: A review. Renewable and
Sustainable Energy Reviews 44: 466-472.
Loraux, C. and E. Brühwiler (2015). Realistic examination of the fatigue life of a wind turbine tower
using data from long term monitoring. IABSE Conference – Structural Engineering: Providing
Solutions to Global Challenges. Geneva, Switzerland.
LORC. (2012a). The Jacket - A Path to Deeper Waters. 2013, from http://www.lorc.dk/offshore-
wind/foundations/jackets#post-piling.
Lundsager, P., S. Frandsen and C. J. Christensen (1980). Analysis of Data from the Gedser Wind turbine
1977-1979. Riso-M-2242. Roskilde, Denmark, Riso National Laboratory.
383
References
Madsen, P. H., S. Frandsen, W. E. Holley and J. C. Hansen (1984). Dynamics and Fatigue Damage of
Wind Turbine Rotors during Steady Operation. Riso-R-512. Roskilde, Denmark, Riso National
Laboratory.
Magalhães, F. (2010). Operational Modal Analysis for Testing and Monitoring of Bridges and Special
Structures. PhD thesis, Faculdade de Engenharia da Universidade do Porto.
Magalhães, F. and Á. Cunha (2011). Explaining operational modal analysis with data from an arch
bridge. Mechanical Systems and Signal Processing 25: 1431-1450.
Magalhães, F., A. Cunha and E. Caetano (2012). Vibration based structural health monitoring of an
arch bridge: From automated OMA to damage detection. Mechanical Systems and Signal Processing
28(0): 212-228.
Magalhães, F., Á. Cunha and E. Caetano (2009). Online automatic identification of the modal
parameters of a long span arch bridge. Mechanical Systems and Signal Processing 23(2): 316-329.
Magalhães, F., Á. Cunha, E. Caetano and R. Brincker (2010). Damping estimation using free decays and
ambient vibration tests. Mechanical Systems and Signal Processing 24(5): 1274-1290.
Main(e) International Consulting LLC (2012). Floating Offshore Wind Foundations: Industry Consortia
and Projects in the United States, Europe and Japan - An Overview.
Manwell, J. F., J. G. McGowan and A. L. Rogers (2010). Wind Energy Explained - Theory, Design and
Application, John Wiley & Sons.
Manzato, S., C. Devriendt, W. Weijtjens, E. Di Lorenzo, B. Peeters and P. Guillaume (2014). Removing
the Influence of Rotor Harmonics for Improved Monitoring of Offshore Wind Turbines. Dynamics of
Civil Structures, Volume 4. F. N. Catbas, Springer International Publishing: 299-312.
Manzato, S., J. R. White, B. LeBlanc, B. Peeters and K. Janssens (2013). Advanced identification
techniques for operational wind turbine data. XXXI International Modal Analysis Conference (IMAC),
California, USA.
Marinone, T., D. Cloutier, B. LeBlanc, T. Carne and P. Andersen (2014). Artificial and Natural
Excitation Testing of SWiFT Vestas V27 Wind Turbines. XXXII International Modal Analysis
Conference (IMAC). Florida, USA.
Marulo, F., G. Petrone, E. D. Lorenzo and A. Cutolo (2015). Operational modal analysis for SHM of a
wind turbine blade. 7th International Conference on Structural Health Monitoring. Torino, Italy.
384
References
McMillan, D. and G. W. Ault (2007). Quantification of Condition Monitoring Benefit for Offshore Wind
Turbines. Wind Engineering 31(4): 267-285.
Molenaar, D. (2003). Experimental Modal Analysis of a 750 kW Wind Turbine for Structural Model
Validation. 41st AIAA Aerospace Sciences Meeting and Exhibit. Nevada, USA.
Montgomery, D. C. (2009). Statistical quality control : a modern introduction. New York, John Wiley &
Sons.
Morthorst, P. E. and H. Jacobsen (2004). Wind Energy - The Facts, European Commission (EC)
European Wind Energy Association (EWEA). Volume 2 - Costs & Prices.
Moser, P. and B. Moaveni (2011). Environmental effects on the identified natural frequencies of the
Dowling Hall Footbridge. Mechanical Systems and Signal Processing 25(7): 2336-2357.
Murtagh, P. J., B. Basu and B. M. Broderick (2005). Along-wind response of a wind turbine tower with
blade coupling subjected to rotationally sampled wind loading. Engineering Structures 27(8): 1209-
1219.
Musial, W., S. Butterfield and A. Boone (2004). Feasibility of Floating Platform Systems for Wind
Turbines. 23rd ASME Wind Energy Symposium. Nevada, USA, Nationa Renewable Energy Laboratory
(NREL).
Musial, W. and B. Ram (2010). Large-Scale Offshore Wind Power in the United States - Assessment of
Opportunities and Barriers, National Renweable Energy Laboratory (NREL).
Naess, A. and T. Moan (2012). Stochastic Dynamics of Marine Structures. USA, Cambridge University
Press.
Nielsen, J. J. and J. D. Sørensen (2011). On risk-based operation and maintenance of offshore wind
turbine components. Reliability Engineering and System Safety(96): 218-229.
Oliveira, G., F. Magalhães, E. Caetano and Á. Cunha (2013). Modal Identification and FE Model
Correlation of a Wind Turbine Tower. 5th International Operational Modal Analysis Conference
(IOMAC). Guimarães, Portugal.
385
References
Oliveira, G., F. Magalhães and Á. Cunha (2014). Rotor blade extension project – Determination of the
blades natural frequencies. Porto, Portugal.
Oliveira, G., W. Weijtjens, G. D. Sitter, F. Magalhães, Á. Cunha, E. Caetano and C. Devriendt (2014).
Dynamic Monitoring of wind turbines: case studies on- and off-shore. 1st International Conference on
Renewable Energies Offshore. Lisboa, Portugal.
Osgood, R. (2001). Dynamic Characterization Testing of Wind Turbines. Colorado, USA, National
Renewable Energy Laboratory (NREL).
Osgood, R., G. Bir, H. Mutha, B. Peeters, M. Luczak and G. Sablon (2010). Full-scale modal wind
turbine tests: comparing shaker excitation with wind excitation. XXVIII International Modal Analysis
Conference (IMAC). Florida, USA.
Overschee, P. V. and B. D. Moor (1996). Subspace Identification for Linear Systems - Theory,
Implementation, Applications, Kluwer Academinc Publishers.
Ozbek, M., F. Meng and D. J. Rixen (2013). Challenges in testing and monitoring the in-operation
vibration characteristics of wind turbines. Mechanical Systems and Signal Processing.
Ozbek, M. and D. J. Rixen (2012). Operational modal analysis of a 2.5 MW wind turbine using optical
measurement techniques and strain gauges. Wind Energy(16).
Ozbek, M., D. J. Rixen, O. Erne and G. Sanowb (2010). Feasibility of monitoring large wind turbines
using photogrammetry. Energy 35: 4802-4811.
Pappa, R. S., G. H. J. III and D. C. Zimmerman (1997). Autonomous modal identification of the Space
Shuttle tail rudder. Virginia, USA, National Aeronautics and Space Administration (NASA).
Peeters, B. (2000). System Identification and Damage Detection in Civil Engineering. PhD thesis,
Katholieke Universiteit Leuven.
Peeters, B. and H. V. d. Auweraer (2005). PolyMax: a Revolution in Operational Modal Analysis. 1st
International Operational Modal Analysi Conference (IOMAC). Copenhagen, Denmark.
Peeters, B., R. Cornelius, K. Janssens and H. V. d. Auweraer (2007). Removing Disturbing Harmonics in
Operational Modal Analysis. 2nd International Operational Modal Analysis Conference (IOMAC).
Copenhagen, Denmark.
Peeters, B., H. Van der Auweraer, P. Guillaume and J. Leuridan (2004). The PolyMAX Frequency-
Domain Method: A New Standard for Modal Parameter Estimation? Shock and Vibration 11(3-4): 395-
409.
386
References
Pelayo, F., M. López-Aenlle, A. Fernández-Canteli and R. Cantieni (2011). Operational modal analysis
of two wind turbines with foundation problems. 4th International Operational Modal Analysis
Conference. Istanbul, Turkey.
Pierson, W. and L. Moskowitz (1964). A proposed spectral form of fully developed wind seas based on
the similarity theory of S. A. Kitaigorodskii. Journal of Geophysical Research 69(4): 5181-5190.
Pollino, M. and A. Huckelbridge (2012). In-situ Measurements of Fatigue Demands on a Wind Turbine
Support Structure. IEEE Energy Tech 2012. Ohio, USA.
Portuguese Renewable Energy Association (APREN). (2015). Produção de Electricidade por Fonte em
Portugal em 2015. Retrieved 05/01/2016, from http://www.apren.pt/pt/dados-tecnicos-3/dados-
nacionais-2/producao-2/a-producao-de-electricidade-em-portugal-2/fontes-de-producao-de-
eletricidade/.
Prendergast, L. J., K. Gavin and P. Doherty (2015). An investigation into the effect of scour on the
natural frequency of an offshore wind turbine. Ocean Engineering 101: 1-11.
Putter, S. and H. Manor (1978). Natural Frequencies of Radial Rotating Beams. Journal of Sound and
Vibration 52(2): 175-185.
Quilter, J. (2013). World's largest blade begins journey to Scotland. 2013, from
http://www.windpowermonthly.com/article/1191655/picture-gallery---worlds-largest-blade-begins-
journey-scotland?DCMP=EMC-CONWindpowerWeekly&bulletin=windpower-weekly.
Ramírez, O., D. Tcherniak and G. Larsen (2015). Comparative study of OMA applied to experimental
and simulated data from an operating Vestas V27 wind turbine. 6th International Operational Modal
Analysis Conference (IOMAC). Gijón, Spain.
Ramler, J. R. and R. M. Donovan (1979). Wind turbines for Electric Utilities: Development Status and
Economisc. N. T.-. DOE/NASA/1028-79/23, AIAA-79-0965.
Randall, R. B. and N. Sawalhi (2011). A New Method for Separating Discrete Components from a Signal.
Sound & Vibration 45(5): 6 - 9.
Rankine, W. J. M. (1865). On the Mechanical Principles of the Action of Propellers, Transactions of the
Institution of Naval Architects.
Reynders, E. (2009). System Identification and Modal analysis in Structural Mechanics. PhD Thesis, K.
U. Leuven.
387
References
Reynders, E., J. Houbrechts and G. De Roeck (2012). Fully automated (operational) modal analysis.
Mechanical Systems and Signal Processing 29(0): 228-250.
Rücker, W. (2007). Offshore wind energy plants – Problems and possible solutions. Experimental
Vibration Analysis for Civil Engineering Structures (EVACES). Porto, Portugal.
Rytter, A. (1993). Vibrational Based Inspection of Civil Engineering Structures. PhD thesis, Aalborg
University.
Salavessa, R. (2014). Improving instead of repowering is a technical, ecological and economic solution.
Prototype - Rotor Blade Extension. European Wind Energy Association (EWEA) 2014 Annual Event.
Barcelona, Spain.
Salzman, D. J. C. and J. v. d. Tempel (2005). Aerodynamic Damping in the Design of Support Structures
for Offshore Wind Turbines. Copenhagen Offshore Wind 2005. Copenhagen, Denmark.
Savonius, S. J. (1931). The S-Rotor and Its Applications. Mechanical Engineering 53(5): 333-338.
Schwerin, B. (2010). Featured Invention: NASA Modeling Innovations Advance Wind-Energy Industry.
Ask Magazine: 22 - 24.
Scionti, M., J. Lanslots, I. Goethals, A. Vecchio, H. V. d. Auweraer, B. Peeters and B. D. Moor (2003).
Tools to improve detection of structural changes from in-flight flutter data. 8th International Conference
on Recent Advances in Structural Dynamics (ISVR). Southampton, UK.
Scionti, M. and J. P. Lanslots (2005). Stabilisation diagrams: Pole identification using fuzzy clustering
techniques. Advances in Engineering Software 36(11–12): 768-779.
Sectorov, W. R. (1973). The first aerodynamic three-phase electric power plant in Balakawa, National
Aeronautics and Space Administration (NASA).
Şeker, S., E. Ayaz and E. Türkcan (2003). Elman's recurrent neural network applications to condition
monitoring in nuclear power plant and rotating machinery. Engineering Applications of Artificial
Intelligence 16(7–8): 647-656.
388
References
Shi, W., J. Shan and X. Lu (2012). Modal identification of Shanghai World Financial Center both from
free and ambient vibration response. Engineering Structures(36): 14-26.
Skjoldan, P. F. (2011). Aeroelastic modal dynamics of wind turbines including anisotropic effects. Risø-
PhD-Report-66. Roskilde, Denmark, Riso - DTU.
Sohn, H. (2007). Effects of environmental and operational variability on structural health monitoring.
Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences
365(1851): 539-560.
Sørensen, J. D. and J. N. Sørensen (2011). Wind Energy Systems: Optimising Design and Construction
for Safe and Reliable Operation. Cambridge, UK
Philadelphia, PA, Woodhead Publishing.
Sørensen, S. P. H. and L. B. Ibsen (2013). Assessment of foundation design for offshore monopiles
unprotected against scour. Ocean Engineering 63: 17-25.
Spera, D. (1994). Wind Turbine Technology - Fundamental Concepts of Wind Turbine Engineering, The
American Society of Mechanical Engineers.
Sutherland, H. J. (2000). A summary of the fatigue properties of wind turbine materials. Wind Energy
3(1): 1-34.
Tcherniak, D., S. Chauhan and M. H. Hansen (2010a). Applicability Limits of Operational Modal
Analysis to Operational Wind Turbines. XXVIII International Modal Analysis Conference (IMAC).
Florida, USA.
389
References
Tcherniak, D., S. Chauhan, M. Rossetti, I. Font, J. Basurko and O. Salgado (2010b). Output-only Modal
Analysis on Operating Wind Turbines: Application to Simulated Data. European Wind Energy
Conference. Warsaw, Poland.
Tegen, S., M. Hand, B. Maples, E. Lantz, P. Schwabe and A. Smith (2012). 2010 Cost of Wind Energy
Review, National Renweable Energy Laboratory (NREL).
Tempel, J. V. D. (2006). Design of Support Structures for Offshore Wind Turbines. PhD Thesis,
Technische Universiteit Delft.
Thiele, H. M. (1984). Growian Rotor Blades: Production Development, Construction and Test. NASA
TM-77479.
Thies, P. R., L. Johanning, V. Harnois, H. C. M. Smith and D. N. Parish (2014). Mooring line fatigue
damage evaluation for floating marine energy converters: Field measurements and prediction.
Renewable Energy 63(0): 133-144.
Thomsen, J. H., T. Forsberg and R. Bittmer (2007). Offshore Wind Turbine Foundations - The Cowi
Experience. 26th International Conference on Offshore Mechanics and Artic Engineering (OMAE).
California, USA.
Thomsen, K., J. T. Petersen, E. Nim, S. Øye and B. Petersen (2000). A Method for Determination of
Damping for Edgewise Blade Vibrations. Wind Energy(3): 233-246.
Thresher, R. W. and D. M. Dodge (1998). Trends in the evolution of wind turbine generator
configurations and systems. Wind ENERGY 1(S1): 70 - 86.
Tong, W. (2010). Wind Power Generation and Wind Turbine Design. Great Britain, WIT Press.
Tricklebank, A. H., P. H. Halberstadt, B. J. Magee and A. Bromage (2007). Concrete Towers for
Onshore and Offshore Wind Farms.
U.S Energy Information Administration (2015). Levelized Cost and Levelized Avoided Cost of New
Generation Resources in the Annual Energy Outlook 2015. Washington, DC.
Ulriksen, M. D., D. Tcherniak, P. H. Kirkegaard and L. Damkilde (2015). EWSHM 2014: Operational
modal analysis and wavelet transformation for damage identification in wind turbine blades. Structural
Health Monitoring.
van Kuik, G. A. M. (2007). The Lanchester–Betz–Joukowsky limit. Wind Energy 10(3): 289-291.
Vargo, D. J. (1974). Wind Energy Developments on the 20th Centurty. Missouri, USA.
390
References
Ventura, C. E., W. D. L. Finn, J.-F. Lord and N. Fujita (2003). Dynamic characteristics of a base isolated
building from ambient vibration measurements and low level earthquake shaking. Soil Dynamics and
Earthquake Engineering(23): 313-322.
Verboven, P., P. Guillaume, B. Cauberghe, E. Parloo and S. Vanlanduit (2003). Stabilization Charts
and Uncertainty Bounds for Frequency-Domain Linear Least Squares Estimators. XXI International
Modal Analysis Conference (IMAC). Florida, USA.
Verboven, P., E. Parloo, P. Guillaume and M. Van Overmeire (2002). Autonomous Structural Heath
Monitoring - Part I: Modal Parameter Estimation and Tracking. Mechanical Systems and Signal
Processing 16(4): 637-657.
Versteijlen, W. G., V. Metrikine, J. S. Hoving, E. Smid and W. E. d. Vries (2011). Estimation of the
Vibration Decrement of an Offshore Wind Turbine Support Structure Caused by its Interaction with
Soil. EWEA Offshore Conference. Amsterdam, The Netherlands.
Vestergaard, J., L. Brandstrup and R. D. G. III (2004). A Brief History of the Wind Turbine Industries in
Denmark and the United States. Academy of International Business (Southeast USA Chapter).
Knoxville, USA: 322-327.
Voaden, G. H. (1943). The Smith-Putnam Wind Turbine. Turbine Topics 1(3): 1-6.
Vugts, J. H. (2000). Considerations on the dynamics of support structures for an OW EC. PhD thesis,
Delft University of Technology.
Wang, Y., M. Liang and J. Xiang (2014). Damage detection method for wind turbine blades based on
dynamics analysis and mode shape difference curvature information. Mechanical Systems and Signal
Processing 48(1–2): 351-367.
Weijtjens, W. (2014). Advanced methods for estimating and monitoring the modal parameters of
operational systems subjected to non-white excitation. PhD thesis, Vrije Universiteit Brussel.
Weijtjens, W., A. Iliopoulos, J. Helsen and C. Devriendt (2015). Monitoring the consumed fatigue life of
wind turbines on a monopile foundation. EWEA Offshore 2015. Copenhagen, Denmark.
Weijtjens, W., T. Verbelen, G. De Sitter and C. Devriendt (2015). Foundation structural health
monitoring of an offshore wind turbine-a full-scale case study. Structural Health Monitoring.
391
References
White, J. R., D. E. Adams and M. A. Rumsey (2010). Modal Analysis of CX-100 Rotor Blade and Micon
65/13 Wind Turbine. XXVIII International Modal Analysis Conference (IMAC). Florida, USA.
Wilkinson, M. and B. Hendriks (2011). Report on Wind Turbine Reliability Profiles, Reliawind.
Wilson, R. E. and P. B. S. Lissaman (1974). Applied Aerodynamics of Wind Power Machine. Oregon
State University.
Wind Directions (January/ February 2007). European Wind Energy Asscociation (EWEA).
Wind Power Monthly. (2003). Wind and insurance industries head for peace treaty. 2013, from
http://www.windpowermonthly.com/article/955632/wind-insurance-industries-head-peace-treaty.
Worden, K., C. R. Farrar, G. Manson and G. Park (2007). The fundamental axioms of structural health
monitoring. Proceedings of the Royal Society A: Mathematical, Physical and Engineering Science
463(2082): 1639-1664.
World Energy Council (WEC) (2013). World Energy Perspective - Cost of Energy Technologies. London,
UK.
Wymore, M. L., J. E. Van Dam, H. Ceylan and D. Qiao (2015). A survey of health monitoring systems
for wind turbines. Renewable and Sustainable Energy Reviews 52: 976-990.
Yang, B. and D. Sun (2013). Testing, inspecting and monitoring technologies for wind turbine blades: A
survey. Renewable and Sustainable Energy Reviews(22): 515-526.
Yang, S. and M. S. Allen (2012). Output-only Modal Analysis using Continuous-Scan Laser Doppler
Vibrometry and applicationto a 20 kW wind turbine. Mechanical Systems and Signal Processing 31:
228-245.
392
References
Yang, W., P. J. Tavner, C. J. Crabtree, Y. Feng and Y. Qiu (2014). Wind turbine condition monitoring:
technical and commercial challenges. Wind Energy(17): 673-693.
Ye, X. W., Y. Q. Ni, K. Y. Wong and J. M. Ko (2012). Statistical analysis of stress spectra for fatigue life
assessment of steel bridges with structural health monitoring data. Engineering Structures 45(0): 166-
176.
Yoo, H. H. and S. H. Shin (1998). Vibration Analysis of Rotating Cantilever Beam. Journal of Sound
and Vibration 212(5): 807-828.
Yoon, S. Y., Z. Lin and P. E. Allaire (2013). Control of Surge in Centrifugal Compressors by Active
Magnetic Bearings. London, England, Springer.
Young, W. C. and R. G. Buynas (2002). Roark's Formulas for Stress and Strain. The United States of
America, McGraw-Hill.
Zaaijer, M., J. v. d. Tempel, P. Fish, T. Topper and A. Vos (2002). Tripod support structure - pre-design
and natural frequency assessment for the 6 MW DOWEC. Delft, the Netherlands, TUDelft.
Zhang, L., R. Brincker and P. Andersen (2005). An Overview of Operational Modal Analysis: Major
Development and Issues. International Operational Modal Analysis. Copenhagen, Denmark.
393