You are on page 1of 16

This article was downloaded by: [Colorado College]

On: 24 February 2015, At: 18:17


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House,
37-41 Mortimer Street, London W1T 3JH, UK

Structure and Infrastructure Engineering:


Maintenance, Management, Life-Cycle Design and
Performance
Publication details, including instructions for authors and subscription information:
http://www.tandfonline.com/loi/nsie20

Control of wind/wave-induced vibrations of jacket-


type offshore wind turbines through tuned liquid
column gas dampers
a a a
Reza Dezvareh , Khosrow Bargi & Seyed Amin Mousavi
a
School of Civil Engineering, College of Engineering, University of Tehran, Tehran, Iran
Published online: 17 Feb 2015.

Click for updates

To cite this article: Reza Dezvareh, Khosrow Bargi & Seyed Amin Mousavi (2015): Control of wind/wave-induced vibrations
of jacket-type offshore wind turbines through tuned liquid column gas dampers, Structure and Infrastructure Engineering:
Maintenance, Management, Life-Cycle Design and Performance, DOI: 10.1080/15732479.2015.1011169

To link to this article: http://dx.doi.org/10.1080/15732479.2015.1011169

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the “Content”) contained
in the publications on our platform. However, Taylor & Francis, our agents, and our licensors make no
representations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose of the
Content. Any opinions and views expressed in this publication are the opinions and views of the authors, and
are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be relied upon and
should be independently verified with primary sources of information. Taylor and Francis shall not be liable for
any losses, actions, claims, proceedings, demands, costs, expenses, damages, and other liabilities whatsoever
or howsoever caused arising directly or indirectly in connection with, in relation to or arising out of the use of
the Content.

This article may be used for research, teaching, and private study purposes. Any substantial or systematic
reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any
form to anyone is expressly forbidden. Terms & Conditions of access and use can be found at http://
www.tandfonline.com/page/terms-and-conditions
Structure and Infrastructure Engineering, 2015
http://dx.doi.org/10.1080/15732479.2015.1011169

Control of wind/wave-induced vibrations of jacket-type offshore wind turbines through


tuned liquid column gas dampers
Reza Dezvareh*, Khosrow Bargi1 and Seyed Amin Mousavi2
School of Civil Engineering, College of Engineering, University of Tehran, Tehran, Iran
(Received 3 March 2014; final version received 30 September 2014; accepted 4 December 2014)

Wind/wave-induced vibrations of jacket-type offshore wind turbines (JOWTs) are suppressed by placing a passive vibration
absorber, called tuned liquid column gas damper (TLCGD), on the turbine nacelle. Adopting an ensemble of 75 wind/wave
combinations, three different JOWTs and a SimuLink-based nonlinear model in time domain, main parameters of the
TLCGD are optimised to reach the minimum standard deviation of nacelle displacement. Obtained results indicate that
contribution of the proposed TLCGD is more pronounced in the case of regular excitations such as those from sea waves and
less turbulent winds. Depending on the wind/wave combination, TLCGD can result in reductions up to 45% and 51% in
Downloaded by [Colorado College] at 18:17 24 February 2015

nacelle displacement standard deviation and maximum acceleration, respectively. As a result, TLCGD deemed to be well
suited to protect fatigue critical JOWTs as well as acceleration-sensitive devices of the nacelle.
Keywords: wind/wave excitation; offshore wind turbines; passive control; tuned liquid column gas dampers

1. Introduction and tension leg platforms (TLPs). The monopile support


As a renewable energy source, wind energy has gained structures are more suitable for depths of about 30 m
significant interest during the last decades. While wind while jacket support structures are commonly used for
energy harvest facilities have a long historical background, depths in the range of 30 – 70 m, and TLPs are mainly
modern wind turbines have been built in the early twentieth used for the depths of more than 70 m. As the support
century in some countries, such as Denmark, Germany, structure has a significant effect on the overall
USA, India and Italy (Vestergaard, Brandstrup, & Goddard, constructional costs of an offshore wind turbine (Seidel,
2004). Within the last 15 years or so, offshore wind turbines 2007), some efforts have been made to investigate jacket-
have gained more attention, due to the fact that they are type supports for higher depths. Beatrice wind farm
generally more stable and of higher speed in comparison (http://www.beatricewind.co.uk) is an example of the
with their land counterparts. With the installation of the aforementioned efforts in which jacket-type supports
wind turbine in an offshore location, however, new were installed in deep waters.
concerns would arise in terms of additional excitations The main focus of this study is also on the jacket-type
from sea waves and special constructional issues (Inter- offshore wind turbines (JOWTs). Commonly, compared to
national Electrotechnical Commission [IEC], 2009). conventional petroleum offshore jacket platforms, JOWTs
Compared to conventional civil engineering structures, have smaller natural frequencies as well as smaller
dynamic behaviour of wind turbines is not fully under- inherent damping. Both features make JOWTs quite
stood. An aeroelastic analysis is required to account for the sensitive to wind/wave-induced vibrations which need to
interaction between wind current and rotating blades of the be accounted for during the design phase of the project
turbine. The so-called blade element momentum (BEM) (Dong, Moan, & Gao, 2011). Moreover, it is well
theory is commonly adopted for the required aeroelastic recognised that welded joints in jacket-type supports are
analysis and many simulation tools, such as FAST vulnerable to fatigue demands. Consequently, a robust
(Jonkman & Buhl, 2005), have been developed based on vibration control strategy is required to address both
this theory. In addition, required simulations of offshore service and ultimate level performances of the JOWT
wind turbines are even more complicated compared to under dynamic excitations.
their land counterparts due to the water – structure Current control techniques can be categorised into two
interaction. general classes of passive and active, which have been
In general, support structures of offshore wind briefly discussed by Soong and Dargush (1997) and Soong
turbines can be categorised as monopiles, jacket-types and Spencer (2002), respectively. Vibration control is not

*Corresponding author. Email: rdezvareh@ut.ac.ir


q 2015 Taylor & Francis
2 R. Dezvareh et al.

a new technique for offshore structures and has been


considered in some earlier studies. For example, dynamic
behaviour of offshore jacket platforms under wave-
induced excitations was reduced through passive and
active dampers by Patil and Jangid (2005) and Mahadik
and Jangid (2003), respectively. Vibration mitigation for
offshore wind turbine is also not a new idea and many
studies have been devoted to this topic (Lackner & Rotea,
2011a, 2011b; Stewart & Lackner, 2011; Colwell & Basu,
2009). In the above-mentioned studies, contribution of
various vibration absorbers, such as tuned mass damper,
Figure 1. (a) Three samples of liquid column energy absorber
tuned liquid column damper (TLCD) and active mass and (b) the TLCGD motion, under base excitation of x and
damper, for monopile and floating offshore wind turbines associated parameters.
have been investigated. Behaviour of offshore wind
turbines with ball vibration absorbers under seismic
excitations as well as equivalent wind/wave excitations
gravitational restoring forces of the vibrating liquid as
has been studied by Zhang, Chen, and Li (2014), and
well as hydrodynamic loss mechanisms, such as friction,
Downloaded by [Colorado College] at 18:17 24 February 2015

encouraging results were also reported.


orifices and elbows.
In 2006, a new generation of TLCDs, called tuned
Another modification to earlier TLCDs is the liquid
liquid column gas damper (TLCGD), was introduced by
column vibration absorber (LCVA), in which vertical
Hochrainer and Ziegler (2006). TLCGD is similar to
columns have different cross-sections compared with the
TLCD in many ways and the main difference stems from
horizontal column. This minor modification could
its improved flexibility in frequency regulation. That is,
convey noticeable contribution in terms of efficiency as
frequency of TLCDs can be tuned only through adjusting
suggested by Chang and Hsu (1998) and Mousavi et al.
their geometrical parameters. However, in the case of
(2012). Note that, similar to LCVAs, TLCGDs can also
TLCGD, it is possible to bring such regulation by
have different cross-sections in vertical and horizontal
adjusting the gas pressure in the vertical columns.
columns.
Contribution of TLCGD towards petroleum offshore
jacket platforms was investigated by Mousavi, Zahrai, and
Bargi (2012, 2013) and different parameters of the
2.2. Governing equation of motion
TLCGD were optimised through comprehensive para-
metric studies. However, the carried out studies were The equation of liquid motion inside the TLCGD can be
limited to earthquake excitation and, to the authors’ derived in different ways. Consider a TLCGD like the one
knowledge, contribution of TLCGD towards offshore illustrated in Figure 1(b), where y represents displacement
wind turbines under wind/wave excitations is not fully of the liquid in the vertical columns and x denotes base
understood. Accordingly, this study aimed to investigate horizontal displacement of the TLCGD. Horizontal and
contribution of TLCGD in reducing wind/wave induced vertical column lengths (which are filled by liquid) are
vibrations of JOWTs. also represented by b and h, respectively. Moreover, Ab
and Ah stand for cross sectional area of the horizontal and
vertical columns, respectively. In Figure 1(b), ha denotes
length of the gas filled vertical columns with initial gas
2. Fundamentals of TLCGD pressure of P0.
2.1. Background According to earlier studies, equation of motion of
Similar to conventional TLCDs, TLCGDs consist of an the proposed TLCGD can be written as (Mousavi et al.,
U-shaped tube partially filled with liquid. However, in 2013):
contrast to TLCDs, vertical columns of the TLCGD are
sealed and its liquid is under a predetermined initial
y€ þ dL jy_ jy_ þ v2A y ¼ 2k1 x€ ; ð1Þ
pressure, as depicted in Figure 1(a). Accordingly,
frequency of the vibrating liquid can be regulated,
through the gas spring effect. In other words, the where
frequency of the TLCGD can be tuned not only by
geometrical parameters, but also through initial gas
pressure in the vertical columns. Similar to conventional 2gð1 þ nP0 =rgha Þ b
v2A ¼ ; k1 ¼ ; ð2Þ
TLCDs, parts of the vibration energy of the structure are 2h þ bl 2h þ bl
transmitted to the liquid and would be dissipated by
Structure and Infrastructure Engineering 3

 Meanwhile, the force F2 would be triggered due to the out


1  
dL ¼ K el 1 þ l 2 þ of phase acceleration of the liquid in the horizontal column
2ð2h þ blÞ with respect to the base acceleration of the TLCGD:

2f h f b 2
N b K or l þ N h K or þ K C þ K E þ
2
þ l
Dh Db
F 1 ¼ mf x€ ; F 2 ¼ rbAb ly€ ¼ rbAh y€ : ð4Þ
if l . 1 ! K C ¼ 0:5l ðl 2 1Þ; ð3Þ
K E ¼ ðl 2 1Þ2
In Equation (4), mf and r are mass and density of the
if l , 1 ! K C ¼ 0:5 ð1 2 lÞ; TLCGD liquid, respectively. Besides, x€ is the absolute
K E ¼ ð 1 2 lÞ 2 acceleration at the base of the TLCGD. For example, in the
if l ¼ 1 ! K C ¼ K E ¼ 0: case of earthquake excitations, ground acceleration should
be added to the relative acceleration of the structure to
obtain the true value of x€ . An algebraic summation of the
In the above equations, vA and dL are, respectively,
above-mentioned forces gives the overall controlling force
natural frequency and hydraulic head loss coefficient of
of the TLCGD as:
the TLCGD. Parameter l is the area ratio and defined
to be the cross-sectional ratio of the vertical column to
the horizontal one. Besides, Kel and Kor, respectively,  yÞ; b
k ¼
Downloaded by [Colorado College] at 18:17 24 February 2015

F ¼ mf ð€x þ k€ : ð5Þ
are hydraulic local head loss coefficients at the elbows 2h þ b=l
and orifices. Nb and Nh denote number of placed
orifices in the horizontal and vertical columns,
respectively. The diameter of the vertical column is In the presented relations, the mass of the TLCGD tubes
presented by Dh and Db stands for the diameter of the (without liquid) was ignored.
horizontal column. Moreover, KC and KE are the
contraction and expansion coefficients, respectively.
Parameter f is also the so-called Darcy’s coefficient of 3. JOWTs
friction. It is crucial to note that liquid and gas mixing 3.1. Considered wind turbine
is not allowed in the TLCGD and should be avoided. A 5 MW benchmark wind turbine is adopted in this study
To address this criterion, as proposed by Linder – with the main characteristics represented in Table 1 and
Silvester and Schneider (2005), the liquid speed in the Figure 2(a). Detailed discussion about the adopted turbine
vertical column needs to be less than 10 m/s. can be found, e.g., in Jonkman, Butterfield, Musial, and
The authors note that, while liquid –gas blending Scott (2009). As depicted in Figure 2(a), with the increase
would definitely change the governing equations, it might of the mean wind speed up to 11.4 m/s, thrust force as well
not necessarily disturb efficiency of the TLCGD. In any as generated power would also increase. However, when
case, behaviour of liquid –gas blended TLCGD is not fully the nominal power (5 MW) of the turbine is achieved,
understood and further discussion about this specific case pitch and stall regulations start such that no further thrust
is out of scope of this study. Finally, it should be pointed force would be transmitted to the rotor and subsequently
out that, similar to most liquid vibration absorbers, the no further power would be generated. Consequently, the
liquid inside the TLCGD is water and all presented generated power would be sustained to the nominal value
formulations have been derived based on this assumption. of 5 MW. Note that, during mean speeds exceeding 25 m/s,
Care should be exercised in adopting these formulations the turbine will cease functioning to avoid any potential
for TLCGDs filled with other liquids. damage.

2.3. Controlling force of the TLCGD


Table 1. The 5 MW turbine characteristics by NREL.
The vibration-reducing effect of TLCGD, or any other
vibration absorber, can be explained either through energy Nominal power 5 (MW)
balance or dynamic equilibrium. The latter point of view is
Rotor orientation Upwind
considered in this study as equations of motion of civil Blades length and number 3 blades, 61.5 m long
engineering structures are commonly written in the form Rotor diameter 126.3 m
of equilibrium of dynamic forces. Hub height 90 m
As denoted by Equation (4), TLCGD would impose Wind nominal speed 11.4 m/s
two types of lateral forces, F1 and F2, on the structure. The Rotor nominal speed 12.1 (RMP)
Generator efficiency 94.4 (%)
force F1 is simply attributed to the inertial force of the Rotor mass 110,000 kg
TLCGD (vertical and horizontal columns) and its liquid.
4 R. Dezvareh et al.

in which the sum of internal forces (inertia, damping and


elastic) equals the external forces, namely, the aero-
dynamic and hydrodynamic forces acting upon the turbine
and its support structure.

3.3.1. The wave load (hydrodynamic force)


According to the current state-of-the-practice, Morison’s
relation is well accepted in order to account for
hydrodynamic forces caused by water – structure inter-
action. The force exerted by waves upon a cylindrical
element is proportional to the ratio of the wavelength to
the cylinder diameter. When this ratio is . 5 (IEC, 2009),
incoming waves would not be affected by the cylindrical
element and the wave force can be written as a summation
of the drag and inertia forces, as follows:
Downloaded by [Colorado College] at 18:17 24 February 2015

F hydro ¼ F Morison ¼ F D þ F I
_ jv 2 u_ j
¼ 0:5rw C D Aðv 2 uÞ ð7Þ

þ rw BðC M v_ 2 ðC M 2 1ÞuÞ:

In Equation (7), Fhydro stands for the hydrodynamic force


Figure 2. (a) The 5 MW turbine characteristics versus speed in vector and FD and FI are the drag and the inertia force vectors,
steady state (Jonkman et al., 2009) and (b) considered JOWTs respectively. Besides, rw denotes water density. Matrices A
with different heights.
and B are the area and volume matrices of the jacket
platform. Scalar parameters CD and CM denote drag and
inertia coefficients, respectively. Moreover, v denotes the
3.2. Considered support structures
wave velocity vector and v̇ represents the wave acceleration
In this study, as shown in Figure 2(b), three jacket-type vector. Finally, u̇ and ü represent velocity and acceleration of
supports are considered for the aforementioned 5 MW the structure. In Equation (7), known as the Morison’s
turbine. All jackets are four-legged with different heights modified equation, effect of the relative movement between
in different water depths. Considered support structure the structure and water is also addressed. Also, the inertia
heights are 62 m, 88 m and 120 m installed in water depths force contains an added mass term proportional to the
of 50 m, 70 m and 100 m, respectively. As presented cylinder acceleration (Laya, Connor, & Sunder, 1984).
in Table 2, cross-sectional dimensions of the support
structure elements are considered to be the same in all
three cases. 3.3.2. The wind load (aerodynamic force)
Energy generation stems from interactions between
airfoils of the turbine blades and the wind current which
3.3. Equation of motion of the JOWT
leads to generation of aerodynamic forces (drag and lift).
The governing equation for JOWT under wind and wave In general, the aerodynamic forces made up from three
loadings can be written as follows: main components as follows:
(1) a rather monotonic component which is mainly
F Inertia þ F Damping þ F Elastic ¼ F Aero þ F Hydro ; ð6Þ produced by the mean wind velocity,
(2) periodic forces which are produced by wind
shears, blades rotation, non-axial winds and
Table 2. Adopted cross sectional dimensions of the support turbine tower shadow and
structures. (3) turbulent portion of aerodynamic forces with
irregular fluctuations.
Inclined brace Horizontal brace leg
BEM is used to calculate imposed aerodynamic forces
0.8 0.8 1.8 External diameter (m) on the considered turbines. According to the classic BEM,
20 20 30 Thickness (mm)
lift and drag forces per unit length of the airfoil can be
Structure and Infrastructure Engineering 5

estimated by (Hansen, 2008): torque, respectively. The number of turbine blades is


denoted by B, and r stands for the distance between the
L ¼ 0:5rair V 2rel cC l ð8Þ airfoil and the rotor centre (hub). Moreover, dr denotes the
radial length increment of each airfoil. Main parameters of
D ¼ 0:5rair V 2rel cC d ð9Þ BEM theory are schematically shown in Figure 3 in which
v, V0, a, u, a and a 0 are angular velocity of rotor, wind
where L and D denote the lift and drag forces imposed on speed, angle of attack, local pitch, axial induction factor
the airfoil, respectively. Also, rair represents the air and tangential induction factor, respectively.
density, c is the airfoil chord and Vrel stands for the sum of Related formulations of the above-mentioned par-
the wind velocity and rotational speed. ameters are not presented here due to the page limitation.
Parameters Cl and Cd denote the coefficients of lift and Readers are referred to Hansen (2008) for more detailed
drag. Projecting the calculated lift and drag forces, the discussion. The authors clarify that all aerodynamic forces
perpendicular and tangential forces on the rotor are are obtained from FAST code (Jonkman & Buhl, 2005)
obtained using: which is based on BEM theory. A fixed-bottom support
platform was adopted in the FAST model to derive the
pN ¼ L cos B þ D sin B ð10Þ
required thrust force. Further details about the FAST
pT ¼ L sin B 2 D cos B ð11Þ model of the 5 MW benchmark offshore wind turbine has
Downloaded by [Colorado College] at 18:17 24 February 2015

been reported by Jonkman et al. (2009). Moreover, only


where pN and pT represent the perpendicular and tangential the thrust load perpendicular to the rotor was considered in
aerodynamic force components on the rotor plane, this study as other wind-induced loads on other directions
respectively, and B denotes the angle between Vrel and the are negligible.
rotor plane. Therefore, the total perpendicular force on the
rotor plane (rotor thrust force) and the rotor torque are
obtained by: 4. Modelling
ð 4.1. Mathematical model of the JOWT– TLCGD system
T ¼ BpN dr ð12Þ Placing the TLCGD on the nacelle and adopting a discrete
lumped mass model, Equation (6) can be rewritten in the
ð matrix form of:
M ¼ rBpT dr ð13Þ
In Equations (12) and (13), T and M indicate the total
perpendicular force (the rotor thrust) and the total rotor Mu€ þ Cu_ þ Ku ¼ F Aero þ F Hydro þ F TLCGD ; ð14Þ

Figure 3. Aerodynamic forces per BEM theory.


6 R. Dezvareh et al.

where, M, C and K stand for mass, inherent damping and lumped at different levels. Mass, stiffness, area and
lateral stiffness matrices, respectively. Moreover, u denote volume matrices are obtained from physical character-
displacement vector and dot operator indicates differen- istics, while damping matrices are obtained from a
tiation with respect to time. Aerodynamic force vector is frequency-independent formulation in which damping
represented by FAero and FHydro is the hydrodynamic force ratio in all modes regulated to be 2%. Detailed discussion
vector. Besides, FTLCGD is the force vector corresponding about frequency-independent damping has been provided
to the TLCGD forces. Further details about aerodynamic by Clough and Penzien (1993).
and hydrodynamic forces are provided in Appendix A. In order to verify the accuracy of the used lumped mass
It should be pointed out that many terms in Equation models in SimuLink, obtained modal properties are
(14) are coupled with each other and conventional finite compared with those of SAP 2000 as presented in
element methods would fail to provide fast computational Tables 3 –5 and illustrated in Figure 4. It is interesting to
tools for these nonlinear-correlated equations. MATLAB note that, effective modal mass at the second mode is
(2008) proved to be a well-suited tool for such computations higher than that of the first mode in all considered JOWTs.
and very large number of analysis (in the order of 104 This is due to the specific stiffness/mass distribution along
analyses or more) can be carried out with reasonable speed the height of the JOWTs. The so-called equivalent static
and without parallel techniques. As a result, in the present procedure is not applicable as response of the structure is
study, Equation (14) is solved through a SimuLink based not well governed by its first mode. As a result, dynamic
Downloaded by [Colorado College] at 18:17 24 February 2015

model as briefly introduced in Appendix B. procedures, either spectral or time history, are required for
seismic design of JOWTs. Further discussion about this
interesting topic is out of scope of this study.
It can be seen that the lumped mass model is in good
4.2. Verification of the discrete model agreement with its 3D finite element counterpart. This is
As previously described, considered JOWTs include a mainly due to the fact that used stiffness matrices are
5 MW wind turbine with the previously presented obtained from flexibility matrices. Note that offshore wind
characteristics and four-legged support structures which turbines are flexible structures with noticeable rotational
are in different water depths of 50 m, 70 m and 100 m. All responses. Accordingly, stiffness matrix should be
JOWTs are modelled with 11 degrees-of-freedom (DOFs) estimated from the flexibility matrix to account for

Table 3. Modal information obtained from SAP and SimuLink models for JOWT-A (first five modes).

Type Mode 1 Mode 2 Mode 3 Mode 4 Mode 5


Period (s) – SAP 2.864 0.390 0.208 0.116 0.074
Period (s) – SimuLink 2.863 0.388 0.208 0.112 0.074
Modal mass ratio (%) – SAP 41 44 6.4 8.7 0.2
Modal mass ratio (%) – SimuLink 40.7 43.8 6.3 8.5 0.18

Table 4. Modal information obtained from SAP and SimuLink models for JOWT-B (first five modes).

Type Mode 1 Mode 2 Mode 3 Mode 4 Mode 5


Period (s) – SAP 2.941 0.529 0.224 0.154 0.084
Period (s) –SimuLink 2.941 0.525 0.223 0.150 0.083
Modal mass ratio (%) – SAP 37 46 5.9 8.7 1.7
Modal mass ratio (%) – Simulink 37.5 46.4 5.5 8.7 1.6

Table 5. Modal information obtained from SAP and Simulink models for JOWT-C (first five modes).

Type Mode 1 Mode 2 Mode 3 Mode 4 Mode 5


Period (s) – SAP 3.029 0.692 0.255 0.193 0.120
Period (s)- SimuLink 3.028 0.688 0.252 0.191 0.117
Modal mass ratio (%) – SAP 34 46 12 5.5 2.2
Modal mass ratio (%) – SimuLink 34.3 46.1 11.4 5.9 2.1
Structure and Infrastructure Engineering 7

occurs at an instant of time. As a result, damper


evaluation based merely on the reduction of maximum
response would be neither sound nor adequate. The
authors opted for standard deviation of the nacelle
response as the optimisation index. The formula can be
written as follows:

vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u N
u1 X 1X N
s¼t ðxi 2 x Þ2 where x ¼ xi ; ð15Þ
N i¼1 N i¼1

where s denotes standard deviation of the nacelle


response and N represents number of the time steps used
in the SimuLink model. Moreover, xi stands for the
nacelle response – displacement or acceleration – in the
ith time step, while x denotes its corresponding mean
value.
Downloaded by [Colorado College] at 18:17 24 February 2015

5.1. A parametric study on TLCGD parameters


Fluid mass and frequency of the TLCGD deemed to be
the main parameters in this study. Accordingly, mass
and frequency ratios of the TLCGD would be optimised
during this subsection assuming following constant
parameters:

b ¼ 16 m; h ¼ 3 m; Dh ¼ 1 m; Db ¼ 1:67 m;

dL ¼ 0:5:

Figure 4. SAP and lumped mass first three mode shapes of (a) It should be pointed out that the head loss coefficient
JOWT-A, (b) JOWT-B and (c) JOWT-C. of the TLCGD (dL) can also have noticeable effect on
the TLCGD efficiency. However, as reported by Mousavi
rotational DOFs as well. Obtained stiffness matrix is a et al. (2013), optimum head loss coefficients of 0.5– 1
condensed matrix with only translational DOFs (Chopra, result in good efficiency in most cases. As a result a
1995). It should be clarified that adopted discrete model constant value of 0.5 is adopted herein. Note that the
has only lateral translational DOF in each level and frequency ratio is defined to be the ratio of TLCGD
considered wind/wave excitations has the same direction, frequency to the JOWT fundamental (first mode)
as proposed in IEC61400-3 (2009). In addition, considered frequency. Besides, the mass ratio is the ratio of TLCGD
external loads assumed to be perpendicular to the rotor of fluid mass to total mass of the JOWT. Optimisation is
the wind turbine due to the yaw control mechanism of the achieved through a comprehensive parametric study in
nacelle and effect of the aerodynamic force on the tower of terms of frequency ratio (a) and mass ratio (m) with the
the wind turbine is neglected. following lower/upper bands and increments.

0:1 Spaced
a ¼ 0:5 ! 1:4
5. Numerical analysis
TLCGD is a rather new vibration absorber and its major
parameters, namely mass ratio and frequency ratio, are 2% Spaced
m ¼ 1%; 2% ! 10%:
not yet optimised for offshore wind turbines under wind/
wave excitations. To assess TLCGD performance in the
time domain, a criterion beyond merely the maximum An ensemble of 75 different wind/wave combinations
response has to be considered, as the maximum response is considered to cover a broad range of possible loading
8 R. Dezvareh et al.

states:
4 Spaced
T pwave ¼ 8 s ! 16 s:
m 4 Spaced m
uwind ¼8 ! 24
s s Note that uwind is the mean wind speed at the nacelle
elevation, Hswave is the significant wave height and
2 Spaced Tpwave is the wave peak period. Obtained results are
H swave ¼ 2 m ! 10 m
summarised in Figure 5 in which as the mass ratio
Downloaded by [Colorado College] at 18:17 24 February 2015

Figure 5. Mean S.D. Displacement reduction versus a for mass ratios of (a) 1%, (b) 2%, (c) 4%, (d) 6%, (e) 8% and (f) 10%.
Structure and Infrastructure Engineering 9

increases, the optimum frequency ratio (the peaks of the 5.2. Contribution of the TLCGD under wind and
presented curves) decreases. Moreover, Figure 5 indi- wave excitations
cates that the higher the mass ratio, the more sensitive 5.2.1. Wind load
the TLCGD’s performance to its frequency. In other
Winds with mean speeds varying from 8 to 24 m/s are
words, as the liquid mass increases, the frequency span
adopted in this section. Note that considered wind speeds
in which a favourable performance was achieved (the
are larger than cut-in and smaller than cut-out speeds.
positive side of the vertical axis in Figure 5), would be
As a result, maximum possible aerodynamic forces are
narrowed. Both of the above-mentioned features are also
expected to occur during the carried out simulations.
the case for TLCGD under seismic excitations (Mousavi
Obtained results in terms displacement reduction at the
et al., 2013).
Downloaded by [Colorado College] at 18:17 24 February 2015

Figure 6. Mean S.D. Displacement reduction percent under different wind speeds.
10 R. Dezvareh et al.

nacelle are illustrated in Figure 6. Note that aforemen- In other words, the waves with height higher than 10 m are
tioned sea states (Hs of 2, 4, 6, 8 and 10 m and Tp of 8, 12 less frequent during annual periods and are considered as
and 16 s) are also considered in this section and each point an ultimate level wave.
on the depicted curves of Figure 6 represents averaged Figure 7 indicates that as the wave height increases the
result of different sea states. efficiency of the placed TLCGD would also increase. This is
due to the fact that when compared with aerodynamic forces,
hydrodynamic forces are rather regular with narrower
5.2.2. Wave load frequency band. As a result, an increase in the hydrodynamic
To investigate wave effects on the TLCGD performance, force leads to more regularity in the structural response and
waves with significant heights of 2, 4, 6, 8 and 10 m and consequently improves damper performance. Note that each
period of 8 s are considered. This range is in accordance point in the represented curves of Figure 7 denotes averaged
with the annual statistics of waves in different seas. result of different wind mean speeds.
Downloaded by [Colorado College] at 18:17 24 February 2015

Figure 7. Mean S.D. Displacement reduction percent versus different wave heights.
Structure and Infrastructure Engineering 11
Downloaded by [Colorado College] at 18:17 24 February 2015

Figure 8. (a) Periodic and (b) irregular aerodynamic force for the mean speed of 12 m/s.

5.2.3. Effect of wind turbulence intensity contribution of the TLCGD is more pronounced in the
In the earlier sections, aerodynamic forces were case of periodic aerodynamic forces which lead to more
accompanied by their inherent turbulences and it is regular structural vibrations compared with those of
observed that higher irregularities in structural vibration turbulent aerodynamic forces. This is in agreement with
results in lower efficiency of the placed TLCGD. Note that obtained results in Figure 7 in which TLCGD performed
all turbulence intensities were considered based on more effectively in the cases of higher wave heights
IEC61400-3 (2009) in which turbulence intensity varies (higher hydrodynamic forces). This behaviour deemed to
with mean wind velocity. be due to the irregularity compensation made by higher
To achieve a better understanding about sensitivity waves which are more regular.
of the TLCGD to wind turbulence, another parametric From Tables 6 –8, it can be seen that reduction in
study is carried out considering both periodic and displacement standard deviation is more encouraging
turbulent aerodynamic forces. The authors clarify that when compared with that of maximum displacement.
periodic aerodynamic forces came from wind shear, Therefore, TLCGD seems to be more effective in the
rotor spinning, non-axial winds and tower shadow. case of fatigue dominated offshore wind turbines.
Simulated in the FAST code, Figure 8 illustrates periodic Moreover, regardless of wind turbulence and wave
and irregular turbulent aerodynamic forces for a sample height, maximum nacelle acceleration was substantially
wind with the mean speed of 12 m/s. In this section, reduced in the case of wind turbine with TLCGD. The
considered TLCGD has a mass ratio of 6% and aforementioned reduction roughly ranges from 10% to
frequency ratio of 0.8. 50% depending on the loading state. As a result,
The computed aerodynamic forces are imposed on the TLCGD is a well-suited controller in order to improve
SimuLink model, and Figures 9 and 10, respectively, serviceability performance of the nacelle mechanical/
illustrate nacelle displacement and acceleration. Besides, electrical devices.
Tables 6 –8 represent the contribution of the TLCGD in From the obtained results, it is turned out that
terms of reduction percentage of the nacelle response. efficiency of a pre-regulated TLCGD has minor sensitivity
Note that only JOWT-C is considered in this section, and to the height and flexibility of its JOWT. The same
adopted wave heights are 2, 6 and 10 m. Again, conclusion has been made earlier by Mousavi et al. (2013)
12 R. Dezvareh et al.
Downloaded by [Colorado College] at 18:17 24 February 2015

Figure 9. Nacelle displacement under mean wind speed of 12 m/s. (a) Irregular force, Hs ¼ 2 m, (b) periodic force, Hs ¼ 2 m,
(c) irregular force, Hs ¼ 6 m, (d) periodic force, Hs ¼ 6 m, (e) irregular force, Hs ¼ 10 m, and (f) periodic force, Hs ¼ 10 m.

Figure 10. Nacelle acceleration under mean wind speed of 12 m/s. (a) Irregular force, Hs ¼ 2 m, (b) periodic force, Hs ¼ 2 m,
(c) irregular force, Hs ¼ 6 m, (d) periodic force, Hs ¼ 6 m, (e) irregular force, Hs ¼ 10 m, (f) periodic force, Hs ¼ 10 m.
Structure and Infrastructure Engineering 13

Table 6. Nacelle displacement and acceleration of JOWT-C, under mean wind speed of 12 m/s and Hs ¼ 2 m.

Periodic aerodynamic force Irregular aerodynamic force


With TLCGD Without TLCGD Reduction (%) With TLCGD Without TLCGD Reduction (%)
S.D. Disp. 0.013 0.017 22.87 0.060 0.062 4.01
Maximum Disp. 0.362 0.378 4.29 0.487 0.472 2 3.19
S.D. Acc. 0.044 0.064 31.66 0.074 0.109 32.65
Maximum Acc. 0.126 0.183 31.02 0.348 0.381 8.78

Table 7. Nacelle displacement and acceleration of JOWT-C, under mean wind speed of 12 m/s and Hs ¼ 6 m.

Periodic aerodynamic force Irregular aerodynamic force


With TLCGD Without TLCGD Reduction (%) With TLCGD Without TLCGD Reduction (%)
S.D. Disp. 0.033 0.045 28.22 0.067 0.075 10.59
Maximum Disp. 0.411 0.469 12.27 0.507 0.541 6.38
S.D. Acc. 0.108 0.170 36.89 0.118 0.186 36.21
Downloaded by [Colorado College] at 18:17 24 February 2015

Maximum Acc. 0.260 0.536 51.52 0.380 0.609 37.56

Table 8. Nacelle displacement and acceleration of JOWT-C, under mean wind speed of 12 m/s and Hs ¼ 10m.

Periodic Aerodynamic Force Irregular Aerodynamic Force


With TLCGD Without TLCGD Reduction % With TLCGD Without TLCGD Reduction %
St.Dev. Disp. 0.058 0.106 45.35 0.081 0.122 33.43
Maximum Disp. 0.480 0.592 18.90 0.555 0.672 17.44
St.Dev. Acc. 0.192 0.406 52.80 0.199 0.419 52.43
Maximum Acc. 0.518 1.019 49.13 0.579 1.020 43.26

for offshore jacket platforms under seismic excitations. this study, is the ratio of TLCGD mass to total mass of the
As a result, the authors believe that reported results of this JOWT. In general, as the mass ratio increases, the
study can be generalised for other JOWTs as well. Based optimum frequency ratio decreases which is in agreement
on this study, to achieve a reliable and robust TLCGD, with earlier studies. Moreover, TLCGDs with higher mass
a mass ratio of 4– 6% with frequency ratio of 0.8– 0.85 ratios are more sensitive to their frequency ratios such that
can be proposed. they can impose destructive effects in the cases of ill-
regulated frequencies. It should be pointed out that
observed optimum frequency ratios are smaller than those
commonly used in other structures. This can be attributed
6. Conclusions to the turbulent nature of the considered excitations as well
This study attempts to reduce vibrations of JOWTs under as pitch/stall regulations of the turbine which lead to
wave and wind excitation through a rather new vibration further distraction in the wind thrust.
absorber called TLCGD. Considering water – structure The comparative study revealed that contribution of
interaction and nonlinear damping mechanism of the TLCGD would be more pronounced in the case of regular
TLCGD, a SimuLink model is developed based on the excitations such as sea waves and winds with low
correlated equations of motions of the JOWT –TLCGD turbulence intensity. Accordingly, TLCGD is more
system. Considering different wind/wave combinations, suitable in offshore area with higher wave heights and
frequency and fluid mass of the TLCGD are optimised lower wind turbulence. Finally the authors point out that
through a parametric study for three JOWTs with different TLCGD is well suited for fatigue critical JOWTs as it
heights. leads to more reduction in standard deviation of
The obtained results indicate that TLCGD would displacements compared with the maximum displace-
perform better in the case of higher mass ratios. However, ments. For example, obtained results indicated that
no noticeable improvement would be achieved for mass TLCGD could reduce standard deviation of nacelle
ratios more than 4%. Note that the defined mass ratio, in displacement up to 45%, while the corresponding
14 R. Dezvareh et al.

reduction in maximum displacement was 18%. Observed Lackner, M.A., & Rotea, M. (2011a). Passive structural control
reductions are highly sensitive to the imposed wind/wave of offshore wind turbines. Wind Energy, 14, 373– 388.
doi: 10.1002/we.426
combination. Moreover, maximum nacelle accelerations
Lackner, M.A., & Rotea, M. (2011b). Structural control of
can be dramatically reduced (up to 50%) by the TLCGD. floating wind turbines. Mechatronics, 21, 704 – 719.
This feature would improve serviceability of mechanical/ doi: 10.1016/j.mechatronics.2010.11.007
electrical devices of the wind turbine which are commonly Laya, E.J., Connor, J.J., & Sunder, S.S. (1984). Hydrodynamic
lumped at the nacelle. Finally, it should be pointed out that forces on flexible offshore structures. Journal of Engineering
Mechanics, 110, 433– 448. doi: 10.1061/(ASCE)0733-9399
TLCGD is an evolving technology and more numerical,
(1984)110:3(433)
experimental and case studies in this field are welcome. Linder-Silvester, T., & Schneider, W. (2005). The moving
contact line with weak viscosity effects- an application and
evaluation of Shikhmurzaev’s model. Acta Mechanica, 176,
254– 258.
Disclosure statement
Mahadik, A.S., & Jangid, R.S. (2003). Active control of offshore
No potential conflict of interest was reported by the authors. jacket platforms. International Shipbuilding Progress, 50,
277– 295.
MATLAB (2008). User guide, Simulink version 7.6.0. Natick,
MA: MathWorks Inc..
Notes Mousavi, S.A., Bargi, K., & Zahrai, S.M. (2013). Optimum
1. Email: kbargi@ut.ac.ir mailto:rdezvareh@ut.ac.ir parameters of tuned liquid column-gas damper for mitigation
Downloaded by [Colorado College] at 18:17 24 February 2015

2. Email: s.a.mousavi@ut.ac.ir of seismic-induced vibrations of offshore jacket platforms.


Structural Control and Health Monitoring, 20, 422– 444. doi:
10.1002/stc.505
Mousavi, S.A., Zahrai, S.M., & Bargi, K. (2012). Optimum
References geometry of tuned liquid column-gas damper for control of
Chang, C.C., & Hsu, C.T. (1998). Control performance of liquid offshore jacket platform vibrations under seismic excitation.
column vibration absorbers. Engineering Structures, 20, Earthquake Engineering and Engineering Vibration, 11,
580– 586. doi: 10.1016/S0141-0296(97)00062-X 579– 592. doi: 10.1007/s11803-012-0143-z
Chopra, A.K. (1995). Dynamics of structures. Theory and Patil, K.C., & Jangid, R.S. (2005). Passive control of offshore
applications to earthquake engineering. Englewood Cliffs, jacket platforms. Ocean Engineering, 32, 1933 – 1949.
NJ: Prentice Hall. doi: 10.1016/j.oceaneng.2005.01.002
Clough, R.W., & Penzien, J. (1993). Dynamic of structures. Seidel, M. (2007). Jacket substructures for the REpower 5 M
New York, NY: McGraw-Hill. wind turbine. In Conference Proceedings European Offshore
Colwell, S., & Basu, B. (2009). Tuned liquid column dampers in Wind 2007. Brussels: European Wind Energy Association
offshore wind turbines for structural control. Engineering (EWEA).
Structures, 31, 358 – 368. doi: 10.1016/j.engstruct.2008. Soong, T.T., & Dargush, F. (1997). Passive energy dissipation
09.001 systems in structural engineering. Chichester: Wiley.
Dean, R.G., & Dalrymple, R.A. (1992). Water wave mechanics Soong, T.T., & Spencer, B.F. (2002). Supplemental energy
for engineering and scientists. Farrer Road, Singapore: dissipation: State-of-the-art and state-of-the-practice.
World Scientific Publishing Co.. Engineering Structures, 24, 243–259. doi: 10.1016/S0141-
Dong, W., Moan, T., & Gao, Z. (2011). Long-term fatigue 0296(01)00092-X
analysis of multi-planar tubular joints for jacket-type Stewart, G., & Lackner, M.A. (2011). The effect of actuator
offshore wind turbine in time domain. Engineering dynamics on active structural control of offshore wind
Structures, 33, 2002– 2014. doi: 10.1016/j.engstruct.2011. turbines. Engineering Structures, 33, 1807– 1816. doi: 10.
02.037 1016/j.engstruct.2011.02.020
Hansen, M.O.L. (2008). Aerodynamics of wind turbines. London: Vestergaard, J., Brandstrup, L., & Goddard, R.D. (2004). A brief
Etherscan. history of the wind turbine industries in Denmark and the
Hochrainer, M.J., & Ziegler, F. (2006). Control of tall building United States. In Academy of International Business
vibrations by sealed tuned liquid column dampers. Structural (Southeast USA Chapter) Conference Proceedings, Stock-
Control and Health Monitoring, 13, 980– 1002. doi: 10.1002/ holm (pp. 322 – 327). East Lansing, MI: Academy of
stc.90 International Business.
International Electrotechnical Commission (2009). Wind tur- Zhang, Z.L., Chen, J.B., & Li, J. (2014). Theoretical study and
bines-Part 3: Design requirements for offshore wind turbines. experimental verification of vibration control of offshore
Geneva: IEC (IEC International Standard 61400-3). wind turbines by a ball vibration absorber. Structure and
Jonkman, J.M., & Buhl, M.L. Jr (2005). FAST user’s guide Infrastructure Engineering, 10, 1087– 1100. doi: 10.1080/
(NREL/EL-500-38230). Golden, CO: National Renewable 15732479.2013.792098
Energy Laboratory.
Jonkman, J.M., Butterfield, S., Musial, W., & Scott, G. (2009). Appendix A
Definition of a 5-MW reference wind turbine for offshore
system development (NREL/TP-500-38060). Golden, CO: In order to achieve time domain descriptions of aerodynamic and
National Renewable Energy Laboratory. hydrodynamic forces exerted on the system, the wind velocity
Kelley, N.D., & Jonkman, B.J. (2007). Overview of the TurbSim and water level variations in time domains are required.
stochastic inflow turbulence simulator. Version 1.21 (NREL/ To calculate time history of wind speed, Kaimal spectrum is
TP-500-41137). Golden, CO: National Renewable Energy adopted which can be described by Equation (A-1). The Kaimal
Laboratory. spectrum is a function of frequency ( f), average wind velocity
Structure and Infrastructure Engineering 15
Downloaded by [Colorado College] at 18:17 24 February 2015

Figure A-1. SimuLink model.

(Vmean), length scale (l), and the turbulence intensity factor It of water particles in different depths are calculated and are used
(Hansen, 2008): in the Morrison formulation to obtain the corresponding
hydrodynamic forces:
I 2t V mean l
Sðf ÞKaimal ¼ 5 : ðA  1Þ  
ð1 þ 1:5 V mean
fl
Þ3 5 H 2s 5  24
Sðf ÞPierson – Moskowitz ¼ exp 2 fT p : ðA  2Þ
16 T 4p f 5 4
In this study, using Turbsim (Kelley & Jonkman, 2007)
which is one of the subsets of FAST code, the turbulence
intensity factor, shear effect, and other contributing factors are Appendix B
considered so as to extract the time domain wind velocity for Figure A-1 illustrates the developed SimuLink model. This
different mean wind velocities. Note that Turbsim uses random model is built based upon the governing equation of motion of the
phases with uniform distribution to obtain wind speed time series. JOWT– TLCGD system, Equation (14). The derived model is
Having wind speeds in time domain, FAST is used to calculate indeed the MDOF model of the JOWT and TLCGD while taking
the aerodynamic force based on the so called BEM theory. into account the interaction between water and the
In order to calculate the time domain of water level, Pierson – support structure. In Figure A-1(a), a screen shot from the
Moskowitz spectrum, which is expressed in Equation (A-2), is model is depicted in which the main three blocks, called main
used. This spectrum is a function of the significant wave height structure, TLCGD and hydrodynamic, are represented. Details of
(Hs), the wave peak period (Tp), and frequency ( f). Using the each block are depicted in Figures A-1(b) – (d). Note that
above-mentioned spectrum and the Inverse Fast Fourier Trans- these blocks are correlated with each other through the closed
form (IFFT), the time history of water level variations would be loops. Direct inputs of the model are the time domain vectors of
obtained. Again, during the IFFT, random phases with uniform the aerodynamic force (obtained from FAST) as well as
distribution are adopted. Subsequently, adopting the so called water level variations (obtained from Pierson – Moskowitz
Airy theory (Dean & Dalrymple, 1992), velocity and acceleration spectrum).

You might also like