You are on page 1of 46

An improved inerter-pendulum tuned mass damper and its

ed
1

2 application in monopile offshore wind turbines


3 ZHU Benruia,b**, Wu Yihana,b, Sun Chaoc*, Sun Daweia,b

iew
4 a. State Key Laboratory of Hydraulic Engineering Simulation and Safety, Tianjin University, Tianjin 300350, China;

5 b. School of Civil Engineering, Tianjin University, Tianjin 300350, China

v
6 c. Department of Civil and Environmental Engineering, Louisiana State University, Baton Rouge, LA 70803, USA

re
7 ** Corresponding author

8 * Corresponding author

9
er
E-mail address: benrui.zhu@tju.edu.cn (B. Zhu). csun@lsu.edu (C. Sun)

10 Abstract: This paper proposes an improved inerter-pendulum tuned mass damper (I-PTMD) to combine the
pe
11 advantages of an inerter unit and the PTMD for bi-directional vibration control. To achieve this objective,

12 two types of I-PTMD, namely SI (series inerter)-PTMD and PSI (series-parallel inerter)-PTMD are

13 developed. The optimal design parameters of the SI-PTMD and PSI-PTMD are derived using the Imperial
ot

14 Competition Algorithm (ICA). Performance of the two types of I-PTMDs are examined via mitigating a
tn

15 single degree-of-freedom system. It is found that the PSI-PTMD outperforms the SI-PTMD in expanding

16 the effective tuning frequency range and enhancing the mitigation effect. To further evaluate the performance
rin

17 of the PSI-PTMD, the National Renewable Energy Lab (NREL) monopile 5 MW baseline offshore wind

18 turbine (OWT) is employed. A 16-DOF(degree-of-freedom) analytical model of the monopile OWT with a
ep

19 PSI-PTMD is established. The generalized aerodynamic, hydrodynamic and seismic loading are derived

20 based on virtual work principle. Research results show that the PSI-PTMD outperforms the traditional
Pr

21 PTMD in mitigating the bi-directional responses of the OWT under multi-hazard conditions.

22

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
23 Key words: inerter-pendulum tuned mass damper; parameter optimization; monopole offshore wind

ed
24 turbines; dynamic response; vibration control

25 1. Introduction

iew
26 With the increasing demand for clean energy, offshore wind turbines (OWTs) are experiencing rapid

27 growth worldwide. They have numerous advantages, including high energy capture, minimal land space

v
28 requirements, and low noise emissions (Bilgili et al., 2011; Gao et al., 2016). However, the complex and

re
29 uncertain environmental conditions, including wind, wave and seismic loads, pose a potential threat to the

30 structural safety of the OWTs (Zhou and Lin, 2013; Liang et al., 2022). The excessive vibrations caused by

31
er
these dynamic loads can jeopardize the structural integrity and operational efficiency of OWTs. To protect

32 OWTs, substantial research efforts have been devoted to developing structural control strategies to ensure
pe
33 the performance and safety of OWTs operating in harsh marine environments.

34 Pitch control is an important method for vibration attenuation during the early stage of OWTs
ot

35 development. Its principle is to reduce the wind load by changing the pitch angle of the blades, thereby

36 avoiding excessive vibration and stresses of the wind turbine components (Jafarnejadsani et al., 2013).
tn

37 Christianse et al. (2011) proposed a linear quadratic regulator (LQR) optimum controller with wind speed

38 estimation and state observer. Research results shown that the controller can maintain the operational safety
rin

39 of the turbine in various harsh wind environments. To improve the control accuracy, Selvam et al. (2009)

40 proposed an independent pitch control method based on feedforward-feedback multivariable linear quadratic
ep

41 Gaussian (LQG) optimum control. Namik and Stol (2010) investigated state-space-based independent pitch

42 control and created a robust controller. Although existing studies have shown that pitch control can reduce
Pr

43 wind turbine loads and mitigate vibrations, it is important to note that pitch control may adversely affect the

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
44 stable power generation of the wind turbines.

ed
45 In addition to pitch control, application of dampers, which is widely-used in traditional civil structures,

46 is another effective approach to reduce wind turbine vibrations. Different types of passive dampers have

iew
47 been developed to alleviate the fluctuation of OWTs, i.e. tuned mass dampers (TMDs) (Murtagh et al., 2008;

48 Lackner et al., 2011; Si et al., 2014), multiple tuned mass dampers (MTMDs) (Dinh et al., 2015; Zuo et al.,

49 2019), tuned liquid dampers (TLDs) (Ghaemmaghami et al., 2013), and tuned liquid column dampers

v
50 (TLCDs) (Colwell et al., 2009; Zhang and Høeg, 2020). The above-mentioned dampers have been shown

re
51 effective in mitigating the vibration of OWTs without structural or environmental varations. However, they

52 may lose their effectiveness when OWTs are structurally degraded or the environment changes (Sun, 2018a).
er
53 In comparison, semi-active control devices with tunable natural frequency and damping ratio are more
pe
54 applicable to systems with time-variant parameters. Sun (2018a, 2018b) adopted a semi-active tuned mass

55 damper (STMD) to control the dynamic response of OWTs. Research findings indicated that the STMD

56 outperforms the passive TMD in controlling the dynamic response of OWTs under multiple hazards and
ot

57 structural damage. Nazokkar (2022) evaluated the effect of semi-active liquid column gas damper on
tn

58 controlling the vibrations of a floating OWT. It was found that using a semi-active damper to control the

59 vibration of floating OWTs under various loads is significantly superior. In addition to semi-active control,

60 active control has been demonstrated to be effective in reducing the vibrations of structures considering
rin

61 system and environmental variations. Fitzgerald et al. (2014) utilized a cable connected active tuned mass

62 damper to control the in-plane vibration of the blades. It was found that the active TMDs can provide better
ep

63 reduction than the passive TMDs. Brodersen et al. (2017) employed an active TMD to reduce the nacelle

64 vibrations of a fixed OWT. The authors found that the active TMD is more effective than the passive TMD
Pr

65 in mitigating the structural responses. However, it should be noted that both the semi-active and active

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
66 control demand more complex mechanisms or actuators to enhance the robustness in retuning natural

ed
67 frequencies and damping properties.

68 The dampers described above primarily target fore-aft response mitigation of the wind turbines,

iew
69 whereas Stewart and Lackner (2014) demonstrated that the wind turbine would be susceptible to multi-

70 dimensional vibrations owing to wind-wave misalignment. To mitigate the bidirectional vibration of OWT

71 nacelle/towers, Sun and Jahangiri (2018, 2019) proposed a three-dimensional pendulum damper (3d-PTMD)

v
72 and compared its performance with that of dual linear TMDs (Lackner et al., 2011). Research findings

re
73 indicated that the bi-directional displacements of the nacelle could be effectively attenuated at various wind,

74 wave conditions and misalignment angles, and the 3d-PTMD was more effective than the dual linear TMDs.
er
75 Zhu et al. (2021) characterized ice-induced frequency lock-in vibration of monopile OWTs under different
pe
76 ice thicknesses and directions. The authors adopted the 3d-PTMD to mitigate the ice-induced resonant

77 responses, and found that the 3d-PTMD can significantly reduce the ice-induced large responses of wind

78 turbines. Furthermore, Jahangiri and Sun (2022a) devised a new three-dimensional non-linear tuned mass
ot

79 damper (3d-NTMD) and apply it to floating OWTs. The results demonstrated that the 3d-NTMD is effective
tn

80 in mitigating the structural motion of the floating OWT in the heave, pitch and roll directions. Leng et al.

81 (2023) used a magnetorheological elastomer (MRE) combined with a TMD. The numerical results revealed

82 that the MRE-based TMD could effectively attenuate the bi-directional dynamic responses of monopile
rin

83 OWT under various operation conditions. To more effectively palliate the fore-aft and side-side vibration of

84 a monopile OWT, Ding et al. (2023) developed a toroidal tuned liquid column damper (TTLCD). Research
ep

85 findings indicated that TTLCD could effectively control lateral vibrations and improved the structural

86 performance of wind turbine against wind and wave loads.


Pr

87 In recent years, researchers have attempted to incorporate "inerter" into the design of dampers for OWTs.

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
88 Smith (2002) first proposed the inerter, which has the advantage of generating an apparent mass significantly

ed
89 larger than its physical mass. Marian and Giaralis (2014) developed a passive tunable mass-damper-inerter

90 (TMDI) by combining an inerter and a TMD. Inerter-based TMDs have been demonstrated effective for

iew
91 vibration control of OWTs, where enhanced vibration mitigation can be accomplished using lightweight

92 devices (Pietrosanti et al., 2020; Sarkar et al., 2020; Zhang et al., 2023). Hu et al. (2018) paralleled the inerter

93 with the TMD and compared the performance of three paralleling methods and found that the paralleling

v
94 system improved the overall performance of the damper. Zhang and Fitzgerald (2020) used a tuned mass-

re
95 damper-inerter (TMDI) to mitigate edgewise vibrations of wind turbine blade. The authors found that TMDIs

96 can control edgewise vibrations in wind turbine blades while requiring significantly less damp strokes than
er
97 classical TMDs. Jahangiri and Sun (2022b) created an analytical model for wind turbine blades using a two-
pe
98 dimensional nonlinear tuned mass damper inerter (2d-NTMDI). The results indicate that the properly

99 designed 2d-NTMDI can effectively reduce the edge-wise and flap-wise responses of the blades. Existing

100 literature reviewed above reveals that integration of an inerter with the TMD has the potential to provide
ot

101 improved performance. Following this principle, performance of the developed 3d-PTMD might be
tn

102 enhanced by introducing an inerter, which, however, remains unknown in existing literature.

103 To fill this gap, the present paper proposes an inerter-pendulum tuned mass damper (I-PTMD) for bi-

104 directional vibration control of OWTs under misaligned wind-wave and seismic loads. Novelty of the present
rin

105 study is twofold. First, two configuration schemes of the I-PTMD, i.e., SI-PTMD and PSI-PTMD are

106 proposed and evaluated. It is found that the PSI-PTMD can improve the performance of the original 3d-
ep

107 PTMD. The optimal design parameters of the PSI-PTMD are derived based on the Imperial Competition

108 Algorithm (ICA). Second, the effectiveness of the proposed PSI-PTMD in mitigating the bi-directional
Pr

109 vibration of the OWT is evaluated. To this end, a mathematical model of the National Renewable Energy

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
110 Laboratory (NREL) 5 MW monopile OWT coupled with the PSI-PTMD is established. The non-collinear

ed
111 wind, wave, and earthquake loads are incorporated in the model. The 3d-PTMD developed by Sun and

112 Jahangiri (2018) are used for comparison. The results show that the PSI-PTMD outperforms the 3d-PTMD

iew
113 in reducing the bi-directional response of the OWT under various loads.

114 This study is structured as follows. In Section 2, detailed design schemes of the two types of I-PTMDs

115 are depicted, and the corresponding mathematical models are derived. In Section 3, the optimal parameters

v
116 of the proposed two I-PTMDs are provided via the ICA, and the performance of the two I-PTMDs are

re
117 evaluated. In Section 4, a coupled analytical model of the NREL 5 MW monopile OWT with a PSI-PTMD

118 is established. Results of the PSI-PTMD in mitigating the NREL 5 MW OWT are presented in Section 5.
er
119 Section 6 presents the final conclusions.
pe
120 2. Mathematical model and design of I-PTMD

121 2.1 Inerter damper


ot

122 By comparing forces and currents in mechanics and electricity (Firestone, 1933), the British scholar
tn

123 Smith (Smith, 2002) proposed a mechanical device called ‘inerter’. Compared to traditional mass blocks

124 (single end mechanical components), the prominent advantage of an inerter is that it can generate an apparent

125 mass much greater than its own physical mass. Meanwhile, as a two-terminal device, the equal and opposite
rin

126 forces applied to the nodes (terminals) are proportional to the relative acceleration between the nodes, as

127 shown in Fig. 1.


ep
Pr

128
129 Fig. 1. Schematic diagram of a two-terminal mechanical element

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
130 The inertia force equation for an ideal inerter is:

ed
131 F  b(v2  v1 ) (1)

132 where F is the force applied at the terminals; v1, v2 are the velocities of the terminals and b is the coefficient

iew
133 of inerter with a unit of kilogram. The dot represents derivative, similarly hereinafter.

134 Numerous references have studied the physical manifestation of the ideal inerter since the concept of

135 the inerter was proposed. One of them is the rack-and-pinion inerter, as shown in Fig. 2.

v
re
er
136
pe
137 Fig. 2. Schematic diagram of mechanical model of a rack-and-pinion inerter

138 According to the principle of the mechanism shown in Fig. 2, the inerter mass coefficient b can be

139 derived as (Smith, 2002):


ot

r2 2
140 b( ) m (2)
r1r3
tn

141 where r1 is the radius of pinions 1; r2 is the radius of gear; r3 is the radius of pinions 2;γis the turning radius

142 of the flywheel; m is the mass of the flywheel. It can be found via Eqn. (2) that the apparent mass of the
rin

r 2γ 2
143 flywheel could be magnified by ( ) times, thereby the inerter can achieve considerable mass
r 1r 3

144 amplification effect through reasonable parameter design.


ep

145 2.2 Combination of an inertial element and a pendulum damper

146 This subsection derives the mathematical model of an inerter-PTMD device. Two types of inerter-
Pr

147 pendulum tuned mass dampers (I-PTMDs) are conceived, as shown in Fig. 3. One I-PTMD consists of an

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
148 inerter connected to the PTMD in series with a dashpot, which is referred to as SI-PTMD. The other is to set

ed
149 the inerter in parallel with a dashpot and connect them in series with the PTMD through a spring, referred to

150 as PSI-PTMD.

v iew
151

re
152 (a) SI-PTMD (b) PSI-PTMD
153 Fig. 3. Schematic diagram of the two types of I-PTMDs

154 2.2.1 SI-PTMD with an SDOF structure

155
er
Fig. 4 (a) shows the scheme of a SI-PTMD attached to a single degree of freedom (DOF) structure,

156 where PTMD has a mass of m2, a massless rod with length l, a dashpot with a damping coefficient cp and an
pe
157 inerter with an equivalent mass b. For simplicity, an equivalent TMD can be used to represent the SI-PTMD

158 to derive the mathematical model, which is depicted in Fig. 4 (b), where k2 and c2 are the equivalent stiffness
ot

159 and damping coefficients, respectively.


tn
rin

160
ep

161 (a) SI-PTMD (b) equivalent model

162 Fig. 4. Scheme model of a DOF system with a SI-PTMD


Pr

163 Based on Fig. 4 (b), the equation of motion of the system can be written as follows:

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
m1 
x1  c1 x1  k1 x1  b( 
x1  
x2 )  c2 ( x1  x2 )  k2 ( x1  x2 )  F (t )
164 (3)
m2 
x2  b( 
x2  
x1 )  c2 ( x2  x1 )  k2 ( x2  x1 )  0

ed
165 where m1, k1, and c1 denote the mass, stiffness and damping coefficient of the primary structure, respectively;

iew
166 x1 is the displacements of mass m1; x2 is the displacement of the pendulum, which can be approximated as

167 x2 ≈ θl with a small rotation θ.

168 For convenience, the following parameters are defined:

v
m2 b b 2 k k2
1  , 2  , 21   , 1  1 , 2 
m1 m1 m2 1 m1 m2  mb

re
169 (4)
 c c2
1  2 ,  1  1 ,  2 
1 21m1 22 (m2  b)

170
er
where μ1 is the mass ratio of the PTMD; μ2 is the ratio of the apparent mass generated by the inerter to the

171 mass of the primary structure; μ21 is the ratio of the apparent mass of the inerter to the mass of the PTMD;
pe
172 ω1 and ω2 are the natural frequencies of the primary structure and the SI-PTMD, respectively; ζ1, ζ2 are

173 damping ratios of the primary structure and the SI-PTMD, respectively. λ1 is the frequency ratio of the SI-
ot

174 PTMD.

175 By substituting Eq. (4) into Eq. (3), Eq. (3) can be written in a matrix format, where the mass matrix
tn

176 M, damping matrix C and stiffness matrix K are:

1  2  2 
177 M  m1  (5)
  2 1  2 
rin

 2   2( 1   2 ) 2 11  2( 1   2 ) 2 11 
178 C  m1  1 1  (6)
  2( 1   2 ) 2 11 2( 1   2 ) 2 11 
ep

  2   2 2 (  2 ) 12 12 ( 1   2 ) 
179 K  m1  1 2 1 2 1 1  (7)
 1 1 ( 1   2 ) 12 12 ( 1   2 ) 

180 The state space equation of the system is expressed as:


Pr

U  AU  BF (t )
181  (8)
 y (t )  DU

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
182 where U is the state vector, i.e.

ed
183 U  [ x1 x2 x1 x2 ]T (9)

184 D is direct transfer matrix i.e. [1 0 0 0 ] ; F is system input vector; y denotes the output state; the matrices

iew
185 A and B can be written as:

 O I 
186 A 1  (10)
 M K  M 1C 

v
 O 
 
187 B   1 1 (11)

re
M  
 0
188 Applying Laplace transform to Eq. (8) gives:

189
 sU ( s )  U (0)  AU ( s )  BF ( s )

Y ( s )  DU ( s )
er (12)
pe
190 Ignoring the static response, i.e., U(0) = 0 and solving Eq. (12) yields the system transfer function H(s):

Y (s)
191 H(s)   D(sI  A)1 B (13)
F(s)

192 Then, the displacement frequency response function of the system is given by:
ot

193 H ( j)  H (s) s j (14)


tn

194 Furthermore, the dynamic amplification factor (DAF) can be obtained as:

195   k1 H( j) (15)

196 where ω is the frequency of external excitation force. η is the dynamic amplification coefficient, which can
rin

197 be used as an effective indicator for evaluating the performance of dampers.

198 2.2.2 PSI-PTMD with an SDOF structure


ep

199 Fig. 5 shows the schematic model of a PSI-PTMD attached to a SDOF system.
Pr

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
ed
iew
200
201 Fig. 5. Scheme of an SDOF Structure with a PSI-PTMD

v
202 The equation of motion of the system can be derived as:

re
m1 
x1  c1 x1  k1 x1  k2 ( x1  x2 )  k3 ( x1  x3 )  F (t )
203 m2 
x2  c3 ( x2  x3 )  k2 ( x2  x1 )  b( 
x2  
x3 )  0 (16)
b( 
x3  
x2 )  c3 ( x3  x2 )  k3 ( x3  x1 )  0

204
er
To simplify Eq. (16), the following parameters are defined:

3 g k2 k3 c3
205 2  ,   , 3  ,  2  (17)
23  m2  b 
pe
1 2 l m2 b

206 where ω΄2 is natural frequency of the pendulum; the ω3 is the natural frequency of the inerter; λ2 is the

207 frequency ratio of the inerter to that of the primary structure; and ζ΄2 is the damping ratio of the PSI-PTMD.
ot

208 By substituting Eq. (17) into Eq. (16), Eq. (16) can be written in a matrix format, where the mass matrix
tn

209 M, damping matrix C and stiffness matrix K are:

1 0 0 
 
210 M  m1  0 1  21 21  (18)
 0  21 
rin

 21

 2 11 0 0 
 
211 C  m1  0 221 22 221 22  (19)
 0 221 22 221 22 
ep


 12  12 112  12 2122 12 112 12 2122 
 
212 K  m1  12 112 12 112 0  (20)
 12 2122 0 2 
1 212 
2
Pr


213 Following the same procedure shown in Eqns. (12) through (15), the transfer function, the frequency

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
214 response function, and the DAF of the PSI-PTMD model can be derived, where details are not shown due to

ed
215 space limitation.

216 3. Parameter optimization of the I-PTMD

iew
217 This section presents the optimal design of the I-PTMD for the maximum reduction effect. To improve

218 the accuracy of the optimization, a linearization method based dual criteria is adopted. The Imperial

v
219 Competition Algorithm (ICA) is employed to search for the optimal solution. Relevant details are presented

re
220 in the following subsections.

221 3.1 Linearization method based dual criterion

222
er
For un-damped primary structures, there are numerous optimization approaches that have been used to
pe
223 design TMDs, such as the H∞ optimization (fixed-point theory), H2 optimization and stability maximization.

224 However, when considering the damping of the primary structure, it is difficult to gain the analytical

225 solutions for the optimum parameters of TMDs. To this end, the equivalent linearization method (Anh and
ot

226 Nguyen, 2012) was proposed to approximately analyze dynamical systems. The authors further improved
tn

227 the equivalent linearization method based on dual criterion and verified the accuracy for small and large

228 structural damping. The present paper adopts this dual criterion to optimize the design of the I-PTMD.

229 For a damped SDOF system, the motion equation is given as follows:
rin

230 
x  2 0 0 x   0 2 x  0 (21)

231 where ω0 and ζ0 are the natural frequency and damping ratio of the system; x, x , x are the displacement,
ep

232 velocity and acceleration of the system, respectively.

233 To construct the equivalent un-damped system, the equation of motion can be expressed as:
Pr

234 
x  e 2 x  0 (22)

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
235 in which

ed
236 e2  02   1 (23)

237 where γ1 is an unknown variable; ωe represents the equivalent circular frequency of the system.

iew
238 Substituting Eq. (23) to Eq. (22) and comparing it with Eq. (21), one can find that if the term γ1x is

239 equal to 2ω0ζ0𝑥, these two equations are equivalent. This is what the conventional linearization methods

240 (CLMs) are dedicated to, as detailed in Caughey (1960). However, the error generated by the conventional

v
241 method increases with the complexity of the system, especially when the system is nonlinear. To further

re
242 improve the CLM, the dual criterion A is introduced by Anh and Nguyen (2013), which is given as:

243 
A   20 x  1x
2
  x  2   x 
T
1 2 0
2
er T
 min
1 , 2
(24)

244 in which

1 T
. T  0 .dt
pe
245 (25)
T
246 where T is an integral upper bound, γ2 is the other introduced unknown variable.

247 In Eq. (24), the first term is the conventional replacement while the second term describes its dual
ot

248 replacement. The unknown parameters γ 1 and γ 2 can be determined by the partial derivatives with Eq.
tn

249 (24), that is:

A
0
1
250 (26)
A
rin

0
 2

251 By solving Eq. (26) and substituting the results to Eq. (23), one can obtain the closed-form expression
ep

252 of the equivalent frequency ωe of the system (detailed derivation process can be found in reference by Anh

253 and Nguyen (2013), i.e.


Pr

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
  (1  cos 2 )  2  sin 2   2
2
(1  cos 2 )  2  sin 2  

e  1   2   2   0

ed
2
254   8  2sin 2
2  (1  cos 2 )  8  2sin 2
2  (1  cos 2 ) 2
 (27)
 
 γe0

iew
255 where γe represents the frequency equivalence coefficient.

256 Furthermore, the expression for the equivalent stiffness ke can be obtained as follows:

257 k e  γe2 k 0 (28)

v
258 where k0 is the structural stiffness coefficient.

re
259 For subsequent optimization of the primary structure with dampers, the corresponding equivalent

260 stiffness will be calculated by Eq. (28).

261
er
3.2 Imperialist competition optimization algorithm

262 Various metaheuristic algorithms have been proposed for parameter optimization such as Genetic
pe
263 Algorithm (Sampson, 1976), Simulated Annealing (Ingbe, 1993), Particle Swarm Optimization (Eberhart

264 and Kennedy, 1995), etc. Recently, a new algorithm named ICA was presented by Atashpaz-Gargari et al.
ot

265 (2007), which was inspired by the socio-political process of imperialistic competition. ICA has been used in

266 many engineering optimization problems since its inception (Kaveh and Talatahari, 2010; Peri, 2019;
tn

267 Dehghani, et al. 2021). In the present study, the ICA is used for the I-PTMD parameter optimization. The

268 main procedure of the algorithm is elucidated here:


rin

269 Step 1: Initialization

270 In ICA, every point of the search space is recognized as a country, which is defined by:
ep

271 country   p1 , p2 ,  pi ,  , pN  (29)

272 where pi represents the ith variable to be optimized. For example, p1 is economic policy, p2 is religion, p3 is
Pr

273 skin color, and so on. The subscript N represents the dimension of the optimization problem. The best country

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
274 to solve the optimization can be obtained by evaluating the cost function f, ie.

ed
275 C  f ( country ) (30)

276 where C represents the cost.

iew
277 To start the optimization algorithm, the initial population is defined. And then selecting Nimp of the most

278 powerful counties to form the empires. The remaining population will be the colonies, as shown in Fig. 6

279 (colonies with the same color as the empire belong to that empire). More details on the colonies assigned to

v
280 each empire can be found in Atashpaz-Gargari et al. (2007).

re
er
pe

281
ot

282 Fig. 6. Generating the initial empires: the more colonies an imperialist possesses, the bigger is its relevant ★ mark.
283 (Reproduced based on Atashpaz-Gargari, et al. (2007))
tn

284 Step 2: Assimilation

285 In each generation, imperialist countries improve their colonies. This fact is modeled by moving all the

286 colonies toward the imperialist, as shown in Fig. 7. In the ICA, it can be expressed as:
rin

287 x ∼ U (0, ξ× d ), ξ >1 (31)

288 where ξ is a number larger than one, d is the distance between colony and its imperialist locations, and U
ep

289 represents a uniform distribution between 0 and 1. If a colony reaches a better location with lower cost than

290 that of its imperialist, the labels of colony and its imperialist will be exchanged.
Pr

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
ed
iew
291
292 Fig. 7 Moving colonies toward their relevant imperialist (Atashpaz-Gargari, et al. (2007)).

v
293 Step 3: Competition

re
294 In the competition phase, the weakest colony in the weakest empire will be more likely to be possessed

295 by the most powerful empires. The normalized total power of the nth empire is defined as

NTCn
er
296 Pn  Nimp
(32)
N
pe
TCi
i 1

297 where NTCi is the relative cost of the ith empire. Nimp is the total number of empires.

298 The assimilation and competition procedure is repeated until only one empire exists. Then the ICA is
ot

299 converged to a state in which all colonies and imperialists have a similar cost.
tn

300 Corresponding to the optimization problem in this paper, the DAF is selected as the cost function, i.e.

301 C  f  k1 H ( j ) max ,   (33)

302 where Θ is the frequency range of the external excitations.


rin

303 The principal design parameters of the I-PTMD are the tuning ratio and damping ratio. Therefore, the

304 variables to be optimized for the SI-PTMD are λ 1 and ζ 2, and that for the PSI-PTMD are λ1, λ2 and ζ΄2
ep

305 respectively. The corresponding optimization problems of SI-PTMD and PSI-PTMD can be formulated as:

Copt1  arg min C (1 ,  2 ) ,for SI-PTMD


 1 , 2
Pr

306  (34)
Copt 2  arg min C (1 , 2 ,  2 ) , for PSI-PTMD
1 ,2 , 2

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
307 3.3 Performance analysis

ed
308 In this subsection, performance analysis is carried out to examine the vibration reduction effect of the

309 SI-PTMD and PSI-PTMD by comparing them with TMD.

iew
310 Considering the constraint of practical applications, the mass ratio μ1 in this paper is chosen as 1%, 2%,

311 3%, 4%, and 5%. The optimal frequency ratio λopt and damping ratio ζopt of traditional TMD are determined

v
312 using Eq. (35) derived by Den Hartog (1956):

re
1 , 31
313 opt   opt  (35)
1  1 8 1  1 

314 For the primary structure considering equivalent stiffness as elucidated in Section 3.1, the calculation

315
er
formula for the optimal frequency ratio can be written as follows:

1
316 eopt  e (36)
1  1
pe
317 To investigate the variation of damping performance of the I-PTMD and the conventional TMD, a

318 damping performance index P is defined as:


ot

319 CI (37)
P
C0

320 where C0 is the cost function of TMD controlled structure under optimal parameters; CI is the cost function
tn

321 of I-PTMD controlled structure under optimal parameters. According to Eq. (37) and the physical meaning

322 of cost function in Eq. (33), if the I-PTMD is more effective than the TMD, then the value of P will be less
rin

323 than 1, vice versa.

324 Fig. 8 (a) and (b) show the analysis results of the performance of the SI-PTMD and PSI-PTMD as a
ep

325 function of the mass ratio μ1 and inerter ratio μ21, respectively. As shown in Fig. 8 (a), P is greater than one,

326 indicating that the performance of the SI-PTMD is worse than the traditional TMD. It also can be found that
Pr

327 for a given μ1, a higher μ21 worsens the mitigation.

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
328 In Fig. 8 (b), P is less than 1, indicating that the PSI-PTMD outperformed traditional TMD. Fig. 8(b)

ed
329 shows that the maximum mitigation effect is achieved when μ1 is 5% and μ21 is 10%, where the mitigation

330 effect of the PSI-PTMD is improved by about 22% compared to a conventional TMD.

v iew
re
331

332
er
Fig. 8. The performance index P of SI-PTMD and PSI-PTMD under different μ1and μ21, respectively

333 Based on the above conclusion, the optimal inerter ratio μ21 is 10% when μ1 is 5%. The DAF of an
pe
334 uncontrolled SDOF system, and the controlled one mitigated by traditional TMD and the optimal PSI-PTMD

335 are illustrated in Fig. 9, where fe is the excitation frequency, and fn the natural frequency of the primary
ot

336 structure.
tn
rin
ep

337
338 Fig. 9. DAF of the primary structure without control, and controlled by TMD and PSI-PTMD, respectively

339 As shown in Fig. 9, the PSI-PTMD can further reduce the structure response as compared to TMD,
Pr

340 especially when the structure undergoes resonance. The peak value of DAF can be lowered by roughly 20%

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
341 when compared with TMD, and the effective tuning frequency range can be extended by around 30%.

ed
342 To facilitate engineering application, the PSI-PTMD design formula of parameters λ1, λ2 and ζ΄2 under

343 different μ1 are proposed through curve fitting with respect to the optimal results obtained by ICA, as shown

iew
344 in Fig. 10. The proposed formulas are:

1  0.714312  0.41291  0.9436

8981  20.471 1.1152, 1  0.02


2
345 2   (38)
151 1.871 1.0962, 1  0.02

v
2

 2  57.142912  8.56861  0.5242

re
346 It should be noted that Eq. (38) is applicable for μ21 being 10%.

er
pe
ot
tn
rin

347
ep

348 Fig. 10. Optimal design parameter fitting of the PSI-PTMD when μ21 is 10%. (a) frequency ratio λ1; (b) frequency ratio λ2;
349 (c) damping ratio ζ΄2

350 4. A numerical case study of PSI-PTMD used in monopile OWTs


Pr

351 This section further evaluates the performance of the PSI-PTMD used for OWT vibration mitigation.

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
352 The NREL (National Renewable Energy Laboratory) 5MW monopile OWT is used. Table 1 lists the specific

ed
353 parameters of the benchmark 5MW OWT.

354 Table 1
355 Parameters of the NREL 5MW baseline wind turbine (Jonkman et al., 2009)

iew
Rating 5 MW
Gross Rotor diameter 126 m
properties Cut-in, rated, cut-out wind speed 3 m/s, 11.4 m/s, 25 m/s
Hub height 90m
Mass 17,740 kg

v
Second moment of inertia 11,776,047 kg·m2
Blade
Damping ratio 0.4775 %

re
Length 61.5m
Mass 347,460 kg
Tower bottom diameter, thickness 6 m, 0.027 m
Tower
Tower top diameter, thickness
er 3.87 m, 0.019 m
Damping ratio 1%
Substructure diameter, thickness 6.0 m, 0.06 m
monopile Pile diameter, thickness 6.0 m, 0.06 m
pe
Pile embedment depth 36 m

356 4.1. Multi-DOF OWT analytical model

357 4.1.1 Model description


ot

358 A 16-DOF coupled dynamic model of the OWT with a PSI-PTMD installed at the tower top is

359 developed in this section. The schematic model of the OWT-PSI-PTMD system under combined loads of
tn

360 winds, waves, and earthquakes is presented in Fig. 11. As shown in Fig. 11, the global coordinate system

361 originates at the intersection of the tower center line and the mean sea level (MSL). The wind-wave
rin

362 misalignment is denoted by β. The direction outside the rotating plane of the wind turbine is denoted as the

363 x-axis, while the direction within the rotating plane of the wind turbine is denoted as the y-axis. Fig. 11 (b)
ep

364 and (c) illustrate the PSI-PTMD installed inside the tower or the nacelle. The origin of the local coordinate

365 system, denoted as Op, corresponds to the initial position of the pendulum when it is at rest. As shown in
Pr

366 Fig. 11 (b) and (c), parameters q13, q14 denote the coordinates of mass block relative to the point Op. q15 and

367 q16 represent the motion of the inerter unit within in-plane and out-of-plane, respectively.

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
ed
v iew
re
368
369 Fig. 11. 5MW monopile OWT with a PSI-PTMD. (a) Side view; (b) PSI-PTMD in in- plane; (c) PSI-PTMD in out-of-plane.

370
er
Fig. 12 illustrates the coordinates of the blades (edgewise and flapwise) and the nacelle. The DOFs of

371 the wind turbine nacelle in fore-aft and side-side directions are represented by q7 and q8, respectively. The
pe
372 in-plane and out-of-plane DOFs of the three blades are denoted by q1 to q6, respectively.
ot
tn
rin

373
374 Fig. 12. Coordinates of the turbine blades and the nacelle
ep

375 The interaction between the foundation and soil is considered in the present study, using an equivalent

376 mechanical model established by Carswell, et al. (2015). In the equivalent model, two sets of horizontal
Pr

377 springs, horizontal dampers, rotational springs, and rotational dampers are introduced to simulate the pile-

378 soil interaction, as shown in Fig. 13. Variables q9 and q10 represent the translational and rotational DOFs of

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
379 the foundation within the xoz plane, and q11 and q12 represent the translational and rotational DOFs within

ed
380 the yoz plane.

v iew
re
381

382 Fig. 13. Simplified foundation model of the offshore wind turbine.

383 4.1.2 Fully coupled OWT analytical model

384
er
The mathematical model of the OWT with PSI-PTMD is derived using the Euler-Lagrange equation,

385 which can be expressed as:


pe
d L L
386 ( )  Qi (39)
dt qi qi

387 where L is the Lagrange operator, which is the difference between the kinetic energy and potential energy
ot

388 of the system; qi is the ith DOF of the system; Qi is the generalized force corresponding to the ith DOF.
tn

389 As indicated in Fig. 11 to Fig. 13, the absolute displacements of the wind turbine nacelle in the fore-aft

390 and side-side directions unacfa, unacss can be approximately represented as follows:

unacfa  q7  q9  hq10
rin

391 (40)
unacss  q8  q11  hq12

392 where h is the nacelle height with reference to the sea bed.
ep

393 Then, the velocity of the nacelle in the fore-aft direction vnacfa and the side-side direction vnacss can be

394 calculated as:


Pr

vnacfa  q7  q9  hq10


395 (41)
vnacss  q8  q11  hq12

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
396 The absolute position of the pendulum can be formulated as:

ed
xp  unacfa  xr
397 yp  unacss  yr (42)

zp  l  l 2  xr2  yr2

iew
398 where xr, yr and zr denote the coordinates of the pendulum in the local coordinate system, respectively; l

399 denotes the length of the pendulum.

v
400 The kinetic energy of the PSI-PTMD system TPI can be written as:

re
  q13 q13  q14 q14  
2
1
 m p   v nacfa   2 v nacfa q13  q13   v nacss   2 v nacss q14  q14 
2 2 2 2
TPI 
401 2  l 2  q132  q142  (43)
1 
b  q15  q 7    q16  q8  
2 2

2  

402
er
where mp is the mass of the PTMD; b is the mass of the inerter.
pe
403 The potential energy of the PSI-PTMD system VPI is given by:

404 V PI  m p g l  
l 2  x r2  y r2 
1 
k b  q15  q 7    q16  q 8  
2 
2 2


(44)

405 where kb is the stiffness of the spring between the damper and the structure, as shown in Fig. 11(c).
ot

406 Based on the above equations, the total kinetic energy T and potential energy V of the OWT coupled
tn

407 with a PSI-PTMD can be written as follows:

408 T  Tb  Tnac  Ttow  T f  TPI (45a)

409 V  Vb  Vtow  V f  VPI (45b)


rin

410 where Tb and Vb denote the kinetic energy and potential energy of the blade, respectively; Tnac denotes the

411 kinetic energy of the nacelle; Ttow and Vtow denote the kinetic energy and potential energy of the tower,
ep

412 respectively; Tf and Vf denote the kinetic energy and potential energy of the foundation, respectively. Those

413 parameters are explained in details in Refs. (Sun et al., 2019).


Pr

414 By substituting Eqns. (45a) and (45b) into Eq. (39), the equation of motion of the system can be derived

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
415 and written in a matrix format as follows:

ed
416 Mq  Cq  Kq  Qwind  Qwave  Qseismic  Qdamp  Fp (46)

417 where M is the mass matrix; C is the damping matrix; Cdamp is the damping matrix obtained by rearranging

iew
418 the generalized force vector caused by damping forces to the left-hand side of the equation; K is the stiffness

419 matrix. The dimension of the mass, damping and stiffness matrices is 16 × 16 for OWT coupled with the

420 PSI-PTMD. Details of, M, C and K can be found in the Appendix. Qwind, Qwave, Qseismic are the generalized

v
421 forces caused by wind, wave and seismic loads, respectively; Fp denotes the generalized force caused by the

re
422 nonlinearity of the pendulum.

423 4.2 Generalized forces er


424 The generalized force vectors can be obtained using the virtual work principle, which is expressed as:
pe
 W
425 Qi  (47)
 qi

426 where δW is the virtual works. In the following subsections, the calculation of generalized aerodynamic
ot

427 loads, hydrodynamic loads, seismic loads and damping loads are described.

428 4.2.1 Aerodynamic loading


tn

429 The aerodynamic loads on the wind turbine blades can be determined using the Blade Element

430 Momentum (BEM) theory. The BEM theory divides the wind blade into N elements and assumes that there
rin

431 is no dependency along the blade span. Therefore, the aerodynamic force on each element can be calculated

432 independently according to lift and drag coefficients of each airfoil, as shown in Fig. 14.
ep
Pr

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
ed
iew
433

v
434 Fig. 14. Local velocity and aerodynamic force of the blade element (Modified based on Ref. Zhu et al. (2021). In this figure,
435 R refers to the rotor radius; dr and c(r) are the element span length and the chord length at the element mid-span,

re
436 respectively; FL and FD lift force FL are lift and the drag force, respectively; ф is the flow angle; Ω is the rotation velocity of
437 the blade; c is the chord length at the element mid-span; α is the wind attack angel and θ is the summation of the pitch angle
438 and the twist angle which is predetermined by the airfoil. r is the distance of the ith blade element to the rotor.

439 The normal force FN and the tangential forces FT can be calculated by:
er
1
FN  Vr2 cCN
440 2 (48)
1
FT  Vr2 cCT
pe
2

441 where ρ is the air density; CN and CT are the normal and tangential coefficients, respectively; Parameter Vr

442 represents the relative wind speed which can be calculated by:
ot

Vr  v 1  a     r 1  b  
2 2
443 (49)
tn

444 where parameters a and b denote the axial speed and tangential speed induction factors, respectively; 𝑣 is

445 the turbulent wind velocity.

446 In the presented study, the normal and tangential wind loads on the blades are iteratively calculated by
rin

447 a code which is developed based on the algorithm proposed in Ref. (Hansen, 2015), where Prandtl’s tip loss

448 factor and the Glauert correction are considered. The turbulent wind field is simulated using the Turbsim
ep

449 software based on IEC Kaimal spectral model (IEC, 2009). Fig. 15 demonstrates the aerodynamic loading

450 on a single blade, where the mean wind speed is 11.4 m/s the turbulence intensity is 10% with a rotational
Pr

451 speed of 12.1 rpm.

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
ed
v iew
re
452

453 Fig. 15. Aerodynamic loads on a single blade

454
er
Based on the obtained wind loads and Eq. (47), the generalized force vector for the wind loads can be

455 derived as follows:


pe
3
Qj,wind   FT (r,t)1edr , Qj+3,wind   FN (r,t)1f dr , Q7,wind   FN (r,t)dr , j 1,2,3
R R R

0 0 0
j1
456 (50)
3
Q8,wind   FT (r,t)cosj dr , Q9,wind  Q7,wind , Q10,wind  hQ7,wind , Q11,wind  Q8,wind , Q12,wind  hQ8,wind
R

0
j1
ot

457 where Qn,wind represents the generalized force of the wind load in the nth degree of freedom; 𝜙1e and 𝜙1f
tn

458 denote the edgewise and flapwise mode shapes of blades, respectively, which can be found in the design

459 document (Jonkman et al., 2009); Ψj is the azimuthal angle of the jth blade.

460 4.2.2 Hydrodynamic loading


rin

461 Wave and current loading on the monopile wind turbines can be estimated using the Morison’s equation

462 (Faltinsen, 1993), where the wave elevation time histories can be generated using the JONSWAP spectrum.
ep

463 The virtual work δWwave done by the wave loads can be expressed as:

 
464 Wwave   dFwave utow  dFwave  (1t q7   q9  z q10 )cos   (1t q8   q11  z q12 )sin   (51)
0 0
Pr

465 where dFwave is the wave force on a microelement segment dz calculated by Morrison's equation, as detailed

466 in Ref. (Sun, 2018a). 𝜙1t denotes the mode shape of tower, which also can be found in the design document

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
467 (Jonkman, et al., 2009); δutow is virtual displacement of the tower; β is the angle between the incident

ed
468 direction of the wave and the positive x-axis, and z is the vertical ordinate from mean water level.

469 Substituting Eq. (51) into Eq. (47), yielding to:

iew
Q7, wave  Fwave ,1 cos  , Q8, wave  Fwave ,1 sin  , Q9, wave  Fwave ,2 cos 
470 (52)
Q10, wave  Fwave ,3 cos  , Q11, wave  Fwave ,2 sin  , Q12, wave  Fwave ,3 sin 

471 where Qn,wave represents the generalized force of the wave load in the nth degree of freedom; the expressions

v
472 for Fwave,1, Fwave,2 and Fwave,3 are detailed in Ref. (Sun and Jahangiri, 2018).

re
473 4.2.3 Seismic loading

474 The total virtual work done by the seismic loading on the wind turbine system consists of three parts,

475 which can be expressed as: er


476  Wseismic   Wseismic ,bl   Wseismic ,nac   Wseismic ,tow (54)
pe
477 in which

 Wseismic,bl   j 1    magx (1 f  q j 3   unacfa )dr    magy (1e q j cos j   unacss )dr 
3 R R

 0 0 
478  Wseismic,nac   M nac agx unacfa  M nac agy unacss
h h
 Wseismic,tow    agx M ( q7tow   q9   q10l )dl   agy M  q8to1 w dl
ot

1
0 0

479 where δWseismic, bl , δWseismic,nacl, and δWseismic, tow are the virtual work done by seismic loading on blades,
tn

480 nacelle, and tower, respectively; 𝑎gx and 𝑎gy denote the seismic acceleration components in x and y

481 directions, respectively; Mnac denotes the mass of the nacelle; 𝑚 and 𝑀 denote the mass density of the
rin

482 blade and the tower respectively, the unit is kg/m; dr is an infinitesimal unit of the blade; denotes the mass

483 density of the tower; ϕ1tow denotes the first mode shape of the tower, and h is the distance from the sea bed
ep

484 to the top of the tower.

485 Substituting Eq. (54) into Eq. (47) yields the earthquake induced generalized forces:
Pr

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
Q j , seismic   m1e cos( j ) agy , Q j  3, seismic   m1 f agx , j  1, 2, 3
Q7 , seismic   (3 m 0  M nac  M 1tow ) agx , Q8, seismic   ( m 0 cos( j )  M nac  M 1tow ) agy

ed
486 (55)
Q9, seismic   (3 m 0  M nac  M 0 tow ) agx , Q10, seismic   (3 hm 0  hM nac  M 2 tow ) agx
Q11, seismic   (3 m 0  M nac  M 0 tow ) agy , Q12, seismic   (3 hm 0  hM nac  M 2 tow ) agy

iew
487 in which

R R
m1e   m1e2 dr , m1 f   m1f2 dr
0 0
488 h h h
M 0tow   M dl , M 1tow   M1t dl , M 2tow   Mldl
0 0 0

v
489 where m0 is the mass of a single blade.

re
490 4.2.4 Damping loading

491 For the generalized force vector due to damping, the recommended damping ratio from the 5MW wind

492
er
turbine technical manual is used for calculation (Jonkman et al., 2009). The virtual work done by the damping

493 force can be expressed as follows:


pe
 W damp  Fdamp u   ceg q1 q1  ceg q 2 q 2  ceg q 3 q3  cfp q 4 q 4  cfp q 5 q5
 cfp q 6 q 6  c7 q 7 q 7  c8 q8 q8  c x q 9 q9  c xθ q10 q10
494 (56)
 c y q11 q11  c yθ q12 q12  c px q13 q13  c py q14 q14  c px q15 q15
 c py q16 q16  c px q 7 q15  c px q15 q 7  c py q8 q16  c py q16 q8
ot

495 Similarly, the generalized damping force for each DOF of the wind turbine can be obtained using Eq.
tn

496 (47), which is derived as:

Q1,damp   ceg q1 , Q2,damp   ceg q 2 , Q3,damp   ceg q3 , Q4,damp   cfp q 4
Q5,damp   cfp q5 , Q6,damp   cfp q 6 , Q7,damp   c7 q 7 , Q8,damp   c8 q8
rin

497 Q9,damp   cx q9 , Q10,damp   cxθ q10 , Q11,damp   c y q11 , Q12,damp   c yθ q12 (57)
Q13,damp   cpx q 7  cpx q15 , Q14,damp   cpy q8  cpy q16 , Q15,damp   cpx q15  cpx q 7
Q16,damp   cpy q16  cpy q8
ep

498 where, ceg and cfp represent the damping in the flap-wise and edge-wise directions of the blades, respectively,

499 with both values set to 0.48%; c7 and c8 represent the damping in the fore-aft and side-side directions of the
Pr

500 nacelle and tower in value of 1%, respectively; cx and cxθ represent the damping in the fore-aft translation

501 and rotation of the wind turbine foundation, respectively; and cy and cyθ represent the damping in the side-

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
502 side translation and rotation of the wind turbine foundation, respectively, with a damping ratio of 0.6%

ed
503 chosen for all of them (Carswell et al., 2015).

504 5. Results and discussions

iew
505 As listed in Table 3, four typical environmental conditions are used to evaluate the effectiveness of the

506 PSI-PTMD in mitigating the wind, wave and seismic response of the OWT. In Table 3, the wave parameters

v
507 correspond to a one-year return period and a fifty-year return period, which are normal operating and extreme

re
508 condition, respectively. The other two conditions are considered as far-field and near-field ground motions

509 under extreme wave loads. The rated wind speed is used under the four loading conditions.

510
511
er Table 3
The combined load conditions of wind, wave and seismic excitation.
Conditions Description
pe
Operating loading Mean wind velocity v0=11.4 m/s, Turbulent intensity TI=10%, Significant wave height Hs=3.55 m,
case 1 (OLC1) Wave period Tp=5.79 s
Operating loading Mean wind velocity v0=11.4 m/s, Turbulent intensity TI=10%, Significant wave height Hs=7.26 m,
case 2 (OLC2) Wave period Tp=7.60 s
Operating loading operational wind-wave loading OLC2 + Far-field seismic loading (Imperial Valley-02, 1940,
ot

case 3 (OLC3) Mw=6.95, PGA=0.4g, Rjb=6.09 km)


Operating loading operational wind-wave loading OLC2 + Near-field seismic loading (Imperial Valley-03, 1951,
case 4 (OLC4) Mw=5.6, PGA=0.4g, Rjb=24.58 km)
tn

512 Based on the results of subsection 3.3, the optimal inerter mass ratio μ21 is equal to 10% when the

513 damper mass ratio μ1 is taken as 5%. The standard deviation of the tower top displacement is chosen as the
rin

514 objective optimization function when applying the ICA to obtain the optimum λ1, λ2 and ζ΄2 in the coupled

515 OWT model with PSI-PTMD. The total simulation time is 300s and the step size is 0.02s. Specific analyses
ep

516 of the OWT dynamic response are presented as follows.

517 5.1. Response of controlled OWT under wind and wave misalignment
Pr

518 The dynamic responses of the NREL 5 MW OWT under OLC1 are calculated, four representative

519 misaligned angles β of wind and wave loads are considered to examine the bi-directional vibration control

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
520 performance of the PSI-PTMD. For comparison, the response results of uncontrolled and controlled by

ed
521 PTMD are also presented in Fig. 16.

522 It can be found from Figs. 16 (a)-(d) that both the PTMD and the PSI-PTMD are effective in reducing

iew
523 the bi-directional vibration of the OWT in all the simulated scenarios. As shown in Fig. 16 (a), the

524 displacement of the nacelle in fore-aft (normal) direction is the largest when the wind and wave loads are

525 aligned. In this scenario, when the PSI-PTMD is used, the standard deviation of the tower top fore-aft

v
526 response is reduced by about 77% and its peak by approximately 32% compared to the uncontrolled case.

re
527 Due to the minimum fluctuation in side-side (lateral) direction, the control effect of the PSI-PTMD is slight.

528 With the misalignment angle β increases, the tower top normal displacement gradually decreases, and the er
529 lateral response increases, as shown in Fig. 16 (b)-(d). One can observe that the damping effect of PSI-
pe
530 PTMD in side-side direction gradually manifests as β increases. When β is 90°, the PSI-PTMD can reduce

531 the standard deviation of the tower top lateral displacement by 40% and the peak by 46%, and the vibration

532 mitigation remains significant in normal direction. From the thumbnails in each scenario, it can be found
ot

533 that the bi-directional damping effect of the PSI-PTMD is better than that of the PTMD. Quantitatively, in
tn

534 Fig. 16 (d), the side-side standard deviation reduction of the response is increased by around 8% when using

535 the PSI-PTMD. A detailed comparison is shown in Fig. 18.


rin

2 0.2
Nacelle fore-aft disp(m)

2 1 0.4 0.02
No control
1.8
2 0.4 No control
0.15 PTMD
0.01No control
(m)

0.8 0.3 PSI-PTMD PTMD


(m)

disp(m)(m)

1.6 0.3 PTMD PSI-PTMD


disp(m)(m)

displacement

0.1 0PSI-PTMD
displacement

0.2
displacement

1.5 0.6
displacement

1.4
1.5 0.2
-0.01
180 185 190 195 200 0.05 180 185 190 195 200
1.2 Time(s) 0.1
0.1
side-side
fore-aft

ep

1
11 000
side-side
side-side
fore-aft
fore-aft

0.8 0.1 0.1


Nacelle

PTMD -0.1
-0.05 PTMD
Nacelle

-0.1
0.05 PSI-PTMD 0.05 PSI-PTMD
0.6 -0.2
Nacelle
Nacelle

0.5
-0.2
-0.1
Nacelle
Nacelle

0.5
0.4 0 -0.3 0
-0.3
-0.15
0.2 -0.05 -0.05
0
𝛽 = 0° -0.4
𝛽 = 0°
Pr

0 50 100 150 200 250 300 0 50 100 150 200 250 300
0 -0.1 -0.4
-0.2 -0.1 Time(s)
Time(s) 150 00 5050 100100
00 5050 100100
0.6 150 0.8200 200 250
1 250300 300 0.6 150 150 0.8
200 200 250
1 250300 300
Time(s)
Time(s) Time(s)
Time(s)
(a)
536

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
22 0.4
0.2

Nacelle fore-aft disp(m)


0.04 No control

ed
0.4

(m)(m)
1.8
2 0.9
0.3
0.4 No control PTMD
disp(m) (m) (m)

0.02 No control
PTMD PSI-PTMD
0.3
0.15 No control
PTMD

displacement
0.8 PSI-PTMD

disp(m)(m)
1.6 0.3
0.2 PTMD
PSI-PTMD
displacement

1.5 0

displacement
0.7 PSI-PTMD
0.2

displacement
displacement

1.4
1.5 0.2
0.6 0.1 -0.02
0.1
0.1
1.2 0.5 0.1 270 275 280 285 290

side-side
180 185 190 195 200
fore-aft

0.1 00
0.1
11 Time(s)

side-side
0.05
0

side-side
PTMD PTMD

iew
side-side
fore-aft
fore-aft

-0.1
0.8 0.05 PSI-PTMD -0.1
-0.1 PSI-PTMD
0.05
Nacelle

Nacelle
-0.2 0

Nacelle
Nacelle

0.6
0.5 -0.2

Nacelle
0 -0.2

Nacelle
0
Nacelle

0.5 -0.3
0.4 -0.3
-0.05
-0.05 -0.3
-0.4 -0.05
0.2
0 -0.4 0 50 100 150 200 250 300
0 50
-0.1
100 150
Time(s)
200 250 𝛽300= 30° -0.4
0 50
-0.1
100 150
Time(s)
Time(s)
200 250 𝛽300= 30°
00 -0.1
00 50
50 0.6
100
100 150150 0.8200200 2501250 300 300 00 5050 100
0.6
100 150150 0.8200 200 2501 250 300 300
Time(s) Time(s) Time(s)
Time(s)
Time(s)
Time(s)
(b)

v
537
2 0.2

Nacelle side-side disp(m)


Nacelle fore-aft disp(m)

0.05

re
0.4
2 0.9 0.4 No control
1.8 PTMD No control
2 No control

(m)
2 0.8 0.3
0.15 PSI-PTMD 0 PTMD
disp(m) (m)
0.3 PTMD
(m)(m)

1.6 0.7 PSI-PTMDPSI-PTMD


displacement(m)

displacement
displacement
0.2
0.2
disp(m)
displacement

0.6
1.4
1.5 -0.05
1.5
displacement

1.5 0.5 0.1 270 275 280 285 290


180 185 190 195 200 0.1 Time(s)
1.2 Time(s)
0.1
side-side
fore-aft

0.1 0 0.1
11 0.05
side-side

PTMD 0 PTMD
11
er
side-side
fore-aft

-0.1
fore-aft

PSI-PTMD PSI-PTMD
fore-aft

0.8 0.05 0.05


Nacelle

Nacelle

-0.1
-0.2 0
Nacelle

0.6
Nacelle

0.5
0 0
Nacelle

0.5
-0.2
Nacelle

-0.3
Nacelle

0.5
0.4
-0.05
-0.05 -0.4
-0.30 -0.05
0.2
0 50 100 150 200 250 300
pe
0
0
0
50
50
-0.1
100
100
150
150
Time(s)
200
200
250
250 𝛽 = 60°
300
300
-0.4 -0.1
Time(s) 𝛽 = 60°
00 0 50
Time(s)
0.6
100 150 0.8 200 1 250 300
-0.1
00 5050 1000.6
100 150 150 0.8
200 200 250 1 250 300 300
0 50 100 150 200 250 300
Time(s)
Time(s) Time(s)
Time(s)
Time(s)
Time(s)
(c)
538
2 0.25
Nacelle side-side disp(m)
Nacelle fore-aft disp(m)

0.05
0.4 No control
1.8
2 0.9 0.4 PTMD
2 0.2
ot

No No control
0.8 PSI-PTMD 0 control
(m)

0.3 PTMD
disp(m) (m)

1.6 0.3 PTMD


disp(m)(m)
(m)

0.7 PSI-PTMDPSI-PTMD
0.15
displacement
displacement

1.4 0.2
0.2
displacement

-0.05
displacement

1.5 0.6
1.5 270 275 280 285 290
0.5 0.1
0.1
1.2 180 185 190 195 200 0.1 Time(s)
side-side
fore-aft

Time(s)
tn

0
11 0.05
side-side

0
fore-aft

1
side-side

-0.1
0.8
fore-aft
Nacelle

-0.1 0
Nacelle

-0.2
Nacelle
Nacelle

0.6
0.5
-0.2
-0.05
Nacelle

-0.3
Nacelle

0.5
0.4
-0.3
-0.4
-0.1
0
0.2 0 50 100 150 200 250 300
rin

0 50 100 150
Time(s)
200 250
𝛽 = 90°
300
-0.4
Time(s) 𝛽 = 90°
00 -0.15
00 5050 100100 150 150 200 200 250
00 50
50 100
100 150150 200 200 250 250 300 300 250 300 300
Time(s) Time(s)
Time(s)
Time(s)
(d)
539
540 Fig. 16. Nacelle fore-aft and side-side displacement responses under uncontrolled, controlled by PTMD and by PSI-PTMD,
ep

541 respectively, under OLC1. (a) β=0°, (b) β=30°, (c) β=60°, (d) β=90°.

542 Figs. 17 (a)-(d) depict the nacelle fore-aft and side-side displacement time histories under OLC2. It can

543 be observed that under combined wind and extreme wave loads, the dynamic responses of the turbine in
Pr

544 uncontrolled scenarios are larger than that under OLC1. In the controlled case, the omnidirectional responses

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
545 of the wind turbine are dramatically suppressed. Compared with the uncontrolled results, the maximum

ed
546 reduction of standard deviation of the side-side tower top displacement with the PSI-PTMD increases to

547 69%, indicating that the PSI-PTMD has better mitigation performance in harsh environments. Furthermore,

iew
548 comparison between the mitigation effects of PSI-PTMD and PTMD indicates that the PSI-PTMD

549 outperforms the PTMD, as shown in the thumbnails in each sub-figure. In Fig. 17 (d), the side-side standard

550 deviation reduction of the response can be improved by around 19% using the PSI-PTMD. In addition, it is

v
551 worth noting that when the wind turbine undergoes severe fluctuations under external loads, the PSI-PTMD

re
552 can reduce the structural vibration more rapidly than the PTMD, as shown in the dashed box in Figs.17 (b)-

553 (d). A detailed performance comparison between the PSI-PTMD and PTMD is shown in Fig. 18. er
2 0.4
Nacelle fore-aft disp(m)

1 0.05
2 0.4 No control
1.8 PTMD
0.3 No control
0.8
pe
PSI-PTMD
disp(m) (m)

0.3 0 PTMD
disp(m) (m)

1.6
0.2 PSI-PTMD
0.6
displacement
displacement

1.5
1.4 0.2 -0.05
180 185 190 195 200 180 185 190 195 200
Time(s) 0.1 Time(s)
1.2 0.1
side-side
fore-aft

1 0
1 0
side-side
fore-aft

0.8
NacelleNacelle

-0.1
NacelleNacelle

-0.1
0.6
ot

0.5 -0.2
-0.2
0.4
-0.3
-0.3
0.2

00
𝛽 = 0° -0.4
𝛽 = 0°
-0.4
00 5050 100
100 150150 200 200 250 250 300 300 0
0 50
50 100
100 150150 200 200 250 250 300 300
tn

Time(s)
Time(s) Time(s)
Time(s)
(a)
554
2
Nacelle side-side disp(m)
Nacelle fore-aft disp(m)

2 0.4 No control 0.06


1.8 0.9 0.5
PTMD 0.04 No control
0.8
rin

disp(m) (m)

0.3 PSI-PTMD 0.02 PTMD


disp(m) (m)

1.6 0.4
0.7 0 PSI-PTMD
displacement
displacement

0.6 0.3
1.5 0.2 -0.02
1.4 0.5 -0.04
180 185 190 195 200 180 185 190 195 200
0.2
1.2 Time(s) 0.1 Time(s)
side-side
fore-aft

0.1
1 0
side-side

0
fore-aft

ep

0.8
Nacelle

-0.1
Nacelle

-0.1
0.6
Nacelle

0.5 -0.2
Nacelle

-0.2
0.4 -0.3
-0.3
0.2 -0.4
0 𝛽 = 30° -0.4 𝛽 = 30°
00 50 100 150 200 250 300 -0.5 0 50 100 150 200 250 300
0 50 100 Time(s)
150 200 250 300 0 50 100 Time(s)
150 200 250 300
Pr

Time(s) Time(s)
(b)
555

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
2

Nacelle fore-aft disp(m)


2 0.4 0.1
0.9
No control
1.8 0.5 No control
PTMD

ed
0.05

disp(m) (m)
0.8 0.3 PTMD
disp(m) (m)

PSI-PTMD
1.6 0.4 PSI-PTMD
0.7

displacement
0
displacement

1.5 0.6 0.2


0.3
1.4 -0.05
0.5 180 185 190 195 200
180 185 190 195 200 0.2
0.1 Time(s)
1.2 Time(s)

side-side
fore-aft

0.1
11 0

side-side
fore-aft

iew
0.8 -0.1
Nacelle

Nacelle
-0.1
0.6
Nacelle

0.5 -0.2

Nacelle
-0.2
0.4
-0.3
-0.3

0.2
0 𝛽 = 60° -0.4
-0.4 𝛽 = 60°
0 50 100 150 200 250 300 0 50 100 150 200 250 300
0 -0.5
0 50 100 Time(s)
150 200 250 300 0 50 100 Time(s)
150 200 250 300
Time(s) Time(s)
(c)

v
556
2

Nacelle side-side disp(m)


2 0.4 0.15

re
No control
1.8 0.9 0.5 No control
PTMD 0.1

disp(m) (m)
0.8 0.3 PSI-PTMD PTMD
disp(m) (m)

1.6 0.4 0.05


0.7
PSI-PTMD
displacement
displacement

1.5 0.6 0.2


0.3 0
1.4
0.5 -0.05
0.2 180 185 190 195 200
1.2 180 185 190 195 200 0.1
Time(s)
side-side

Time(s)
fore-aft

0.1
11 er 0
side-side

0
fore-aft

0.8
Nacelle

-0.1
Nacelle

-0.1
0.6
-0.2
-0.2
Nacelle

Nacelle

0.5
0.4 -0.3
-0.3
pe
0.2 -0.4
0
𝛽 = 90° -0.4 𝛽 = 90°
00 50 100 150 200 250 300 -0.50 50 100 150 200 250 300
0 50 100 150 200 250 300 0 50 100 150 200 250 300
Time(s) Time(s)
Time(s) Time(s)
(d)
557
558 Fig. 17. Nacelle fore-aft and side-side displacement responses under uncontrolled, controlled by PTMD and by PSI-PTMD,
559 respectively, under OLC1. (a) β=0°, (b) β=30°, (c) β=60°, (d) β=90°.
ot

560 Fig. 18 demonstrates the bi-directional response mitigation improvement of PSI-PTMD when compared
tn

561 to the PTMD under OLC1 and OCL2 conditions. The variable Rstd in the y-axis of Fig. 18 (a) and Fig. 18

562 (b) represents the improved percentage of reduction of standard deviation of tower top displacement of the

563 PSI-PTMD in comparison with the PTMD. Variable Rpeak in Fig. 18 (c) and Fig. 18 (d) refers to the
rin

564 improvement of peak reduction. The results show that the PSI-PTMD outperforms the PTMD. Under OLC1,

565 the Rstd can be further improved by around 4% and 8% in the fore-aft and the side-side direction,
ep

566 respectively. Under OLC2, Rstd can be improved by approximately 11% in the normal direction and

567 approximately 19% in the lateral direction, which shows that the mitigation effect of the PSI-PTMD is more
Pr

568 pronounced in harsher environments. For peak response mitigation, the improved reduction effect Rpeak of

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
569 the PSI-PTMD is minimal in the fore-aft direction, while it is considerable in the side-side direction. The

ed
570 maximum of Rpeak in Fig. 18 (d) is approximately 11% and 5% under OLC1 and OLC2, respectively.

(a) Fore-aft
Fore-aftStd
Std
(b) Side-side
Side-sideStd
Std
20 20

iew
20 20
Rstd (%)
Rstd (%)

10 10
10 10

0 0
0 OLC1 OLC2 0 OLC1 OLC2

v
OLC1 OLC2 OLC1 OLC2
(c) Fore-aft
Fore-aftPeak
Peak
(d) Side-side
Side-sidePeak
Peak
20 20
20 20

re
Rpeak (%)
Rpeak (%)

10 10
10 10

0
0 OLC1
OLC1
er
OLC2
OLC2
β=0° β=30°
0
0 OLC1
OLC1
β=60°
OLC2
OLC2
β=90°
571
pe
572 Fig. 18. Performance evaluation of the PSI-PTMD relative to PTMD under OLC1 and OLC2.

573 5.2. Responses under misaligned wind-wave and seismic loading

574 For the OWT in earthquake-prone areas, the normal operation and structural safety can be affected by
ot

575 seismic loads in addition to the combined wind and wave loads. Therefore, this section further evaluates the
tn

576 performance of the proposed PSI-PTMD under multi-hazard conditions. Two scenarios, i.e. near-field

577 seismic condition (OLC3) and far-field seismic condition (OLC4), are simulated and analyzed in this
rin

578 research. Fig. 19 illustrates the acceleration time histories of the selected seismic motion from the PEER

579 database (Ancheta et al., 2014) and the corresponding displacement response spectrum. From Fig. 19 (c), it
ep

580 can be observed that the displacement response spectrum of the two earthquakes are different. The near field

581 earthquake causes a larger displacement response at the 1st natural period of the NREL 5MW OWT.
Pr

582 For simplicity, only the controlled responses by PSI-PTMD and PTMD of the OWT under the

583 conditions of OLC3 and OLC4 are presented. In the simulations, the misalignment angle between the wind

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
584 and wave loads β is set to be 30°. Assuming that the seismic load acts along the y-axis, starting from t = 50

ed
585 s. The calculation results of the nacelle displacement responses are shown in Fig. 20 and Fig. 21, respectively.

v iew
re
er
586
587 Fig. 19. The two selected earthquake ground motions. (a) Acceleration time history of the near-field earthquake; (b)
pe
588 acceleration time history of the far-field earthquake; (c) displacement response spectrum.

1.5
No control
1.5 PTMD
(m)(m)

No control
1 PSI-PTMD
PTMD
ot
Fore-aft

1 PSI-PTMD
Fore-aft

0.5
0.5
0
tn

589 00 50 100 150 200 250 300


0 50 100 150 200 250 300
590 (a)
0.4
0.4
(m)(m)

0.2
0.2
rin Side-side

0
Side-side

0
-0.2
Earthquake
-0.2
-0.4 Earthquake
-0.4 0 50 100 150 200 250 300
ep

0 50 100 Time(s)
150 200 250 300
591 Time(s)
592 (b)
593 Fig. 20. Nacelle displacement time-histories comparison between the PTMD and the PSI-PTMD under OLC3.
Pr

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
1.5

ed
No control
1.5

(m)(m)
PTMD
No control
1 PSI-PTMD
PTMD

Fore-aft
1 PSI-PTMD

Fore-aft
0.5
0.5

iew
0
594 00 50 100 150 200 250 300
0 50 100 150 200 250 300
595 0.4
(a)

0.4
(m)(m)

0.2
Side-side

0.2

v
0
Side-side

0
-0.2

re
Earthquake
-0.2
-0.4 Earthquake
0 50 100 150 200 250 300
-0.4
0 50 100 Time(s)
150 200 250 300
596 Time(s)
597 (b)
598
er
Fig. 21. Nacelle displacement time-histories comparison between the PTMD and the PSI-PTMD under OLC4.

599 From Fig. 20 and Fig. 21, one can find that the response of nacelle in fore-aft direction is identical due
pe
600 to the same loads under OCL3 and OCL4. In this direction, the control performance of the PIS-PTMD is

601 better than that of the PTMD. The side-side displacement of the uncontrolled wind turbine under near-field
ot

602 seismic loading is higher than that under far-field seismic loading, which is consistent with the displacement

603 response spectrum, as shown in Fig. 19(c). Comparing the vibration reduction of the PSI-PTMD and PTMD,
tn

604 one can observe that the PSI-PTMD provides better mitigation effect for seismic responses. Moreover, as

605 illustrated in Fig. 21(b), the PSI-PTMD can mitigate the response of the nacelle to a low level more quickly
rin

606 than the PTMD when an earthquake occurs.

607 Detailed performance comparison between the PSI-PTMD and PTMD under OLC3 and OLC4 is
ep

608 illustrated in Fig. 22. It can be found that PSI-PTMD can improve the mitigation of the nacelle bi-directional

609 displacement, especially in the side-side direction. The Rstd can be further improved by a maximum of 17.5%
Pr

610 and the Rpeak by 22.9% under the OLC3, and the Rstd by 17.5% and Rpeak by of 20.8% under OLC4,

611 respectively.

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
(a) Near-field (b) Far-field
20

ed
Rstd
Reduction(%) Rpeak

iew
10

0
Fore-aft Side-side
612

v
613 Fig. 22. Performance evaluation of the PSI-PTMD relative to PTMD under OLC3 and OLC4.

re
614 6. Conclusions

615 In this paper, an improved I-PTMD (inerter-pendulum tuned mass damper), which is an extension of
er
616 the PTMD, is proposed to enhance the mitigation effectiveness for the bi-directional vibration of wind
pe
617 turbines under multi-hazard conditions. Two types of I-PTMDs (SI-PTMD and PSI-PTMD respectively) are

618 proposed. The optimum design parameters are derived using the Imperial Competition Algorithm. To

619 evaluate the performance of the I-PTMD, a multi DOF mathematical model of the NREL 5 MW OWT with
ot

620 a PSI-PTMD is established, and the control performance of the PSI-PTMD under different wind, wave and
tn

621 earthquake conditions is evaluated. Based on the simulation results and discussions, the following

622 conclusions are drawn.


rin

623 (1) The performance of the SI-PTMD is worse than that of traditional TMDs. Instead, the PSI-PTMD

624 can improve the mitigation effect and effective tuning frequency range of traditional TMDs, which shows
ep

625 great potential to be applied in real engineering. For a μ1 of 5%, the optimum level of vibration reduction is

626 achieved when the design parameter μ21 of the PSI-TMD is 10%.
Pr

627 (2) The PSI-PTMD can effectively mitigate the bi-directional vibration of the nacelle displacement of

628 the NREL 5 MW OWT under misaligned wind and wave loadings. It exhibits better performance than

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
629 conventional PTMD, with a maximum mitigation improvement of approximately 11% for peak reduction

ed
630 and 19% for standard deviation mitigation.

631 (3) Under combined actions of multiple hazards including winds, waves, and earthquakes, PSI-PTMD

iew
632 has better bi-directional vibration reduction effect then the PTMD, and exhibits good robustness in both

633 near-field and far-field earthquakes. In comparison with the PTMD, the standard deviation of the nacelle

634 displacement of the wind turbine can be further reduced by 17%, and the peak response by more than 20%.

v
635 As a summary, the proposed PSI-PTMD is more effective than the 3d-PTMD in mitigating the bi-

re
636 directional vibration of the OWTs under misaligned loading. However, there are many ways to introduce

637 inerter into a damper and only two of them are investigated in this paper. Therefore, other configurations of
er
638 the I-PTMD is being studied by the authors to find the optimal one.
pe
639 Declaration of competing interest

640 The authors declare that they have no known competing financial interests or personal relationships
ot

641 that could have appeared to influence the work reported in this paper.
tn

642 Acknowledgements

643 The study was financially supported by the National Natural Science Foundation of China Project
rin

644 [No.51509184], the China Scholarship Council [No. 201906255007] and Tianjin University Independent

645 Innovation Fund Project [No. 2020XT-0027].


ep

646

647 Appendix
Pr

648 As for the OWT with PSI-PTMD system, it is a 16 DOFs system, and the mass matrix M in Eq. (46)

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
649 can be derived as:

ed
 m1 0 0 0 0 0 0 m18 0 0 m18 hm18 0 0 0 0 
 
 0 m1 0 0 0 0 0 m28 0 0 m28 hm28 0 0 0 0 
 0 0 m1 0 0 0 0 m38 0 0 m38 hm38 0 0 0 0 
 
 0 0 0 m2 0 0 m47 0 m47 hm47 0 0 0 0 0 0 

iew
 0 0 0 0 m2 0 m47 0 m47 hm47 0 0 0 0 0 0 
 
 0 0 0 0 0 m2 m47 0 m47 hm47 0 0 0 0 0 0 
 0 0 0 m47 m47 m47 M7 0 m7 hm7 0 0 mp 0 0 0 
 
m m28 m38 0 0 0 0 M8 0 0 m8 hm8 0 mp 0 0 
M   18 
 0 0 0 m47 m47 m47 m7 0 M9 hm7 0 0 mp 0 0 0 
 0 0 0 hm47 hm47 hm47 hm7 0 hm7 M 10 0 0 hmp 0 0 0 
 

v
 m18 m28 m38 0 0 0 0 m8 0 0 M 11 hm8 0 mp 0 0
 hm18 hm28 hm38 0 0 0 0 hm8 0 0 hm8 M 12 0 hmp 0 0
 

re
 0 0 0 0 0 0 mp 0 mp hmp 0 0 m13 m1314 b 0 
 0 0 0 0 0 0 0 mp 0 0 mp hmp m1314 m14 0 b 
 
 0 0 0 0 0 0 0 0 0 0 0 0 b 0 b 0
650  0 (58)
 0 0 0 0 0 0 0 0 0 0 0 0 b 0 b 

651 in which
er
R R R R
m1   m1e2 dr , m2   m1f2 dr , m j 8   m1e dr cos  j , j  1, 2,3 m47   m1f dr
pe
0 0 0 0
h
M 1t   mt 1t  dz , m7  m8  3m0  mhub  mnac  M 1t
2
0
h
652 m9  m11  3m0  mhub  mnac  mt  mf , mt   mt zdz
0

m10  m12  h m7  I f
2
ot

 q2   q2  q q
m13  1  2 213 2  mp  bm14  1  2 214 2  mp  b , m1314  2 132 14 2 mp
 l  q13  q14   l  q13  q14  l  q13  q14
tn

M 7 =m7 + m p , M 8 =m8 + m p , M 9 =m9 + m p , M 11 =m11 + m p


653
M 10 =m10 + h 2 m p , M 12 =m12 + h 2 m p

654 where m0 denotes the mass of the wind turbine blade; 𝑚 denotes the mass density function of the blade per
rin

655 unit length; 𝑚t denotes the tower mass density function per unit length; mt denotes the mass of the wind

656 turbine tower; mf denotes the mass of the wind turbine foundation; If denotes the rotational inertia of the
ep

657 wind turbine foundation; and h denotes the nacelle-to-mud surface distance.

658 The damping matrix C in Eq. (46) can be expressed as:


Pr

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
 c1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 
 
 0 c1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 

ed
 0 0 c1 0 0 0 0 0 0 0 0 0 0 0 0 0 
 
 0 0 0 c4 0 0 0 0 0 0 0 0 0 0 0 0 
 0 0 0 0 c4 0 0 0 0 0 0 0 0 0 0 0 
 
 0 0 0 0 0 c4 0 0 0 0 0 0 0 0 0 0 
 0 

iew
0 0 0 0 0 0 c7 0 0 0 0 0 cpx 0 0
 
 2  2
m81 2 2 m82 2 2 m83 0 0 0 0 c8 0 0 0 0 0 cpy 0 0 
C 
 0 0 0 0 0 0 0 0 cx 0 0 0 0 0 0 0 
 0 0 0 0 0 0 0 0 0 cxθ 0 0 0 0 0 0 
 
 2  2
m81 2 2 m82 2 2 m83 0 0 0 0 0 0 0 cy 0 0 0 0 0 
 2h 2 m81 2h 2 m82 2h 2 m83 0 0 0 0 0 0 0 0 cyθ 0 0 0 0 
 

v
 0 0 0 0 0 0 0 0 0 0 0 0 cpx 0 cpx 0 
 0 0 0 0 0 0 0 0 0 0 0 0 0 cpy 0 cpy 
 

re
 0 0 0 0 0 0 0 0 0 0 0 0 cpx 0 cpx 0 
659  (59)
 0 0 0 0 0 0 0 0 0 0 0 0 0 cpy 0 cpy 

660 in which

661
R
m8, j   m1e dr sin  j , j  1, 2,3
0
er
662 where, parameters 𝑐1 and 𝑐4 denote the edgewise and flapwise damping of blades, 𝑐7 and 𝑐8 are the fore-aft
pe
663 and lateral damping of the tower together with monopile, cx and 𝑐𝑥𝜃 denote the fore-aft translational and

664 rotational damping of the foundation, cy and 𝑐𝑦𝜃 denote the side-side translational and rotational damping
ot

665 of the foundation, and 𝑐𝑝𝑥 and 𝑐𝑝𝑦 denote the fore-aft and side-side damping of the damper, respectively.

666 The damping matrix K in Eq. (46) can be expressed as:


tn

 k11, 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 
 0 k 0 0 0 0 0 0 0 0 0 0 0 0 0 0 
 1,2

 0 0 k1,3 0 0 0 0 0 0 0 0 0 0 0 0 0 
 
 0 0 0 k2,1 0 0 0 0 0 0 0 0 0 0 0 0 
rin

 0 0 0 0 k2,2 0 0 0 0 0 0 0 0 0 0 0 
 
 0 0 0 0 0 k2,3 0 0 0 0 0 0 0 0 0 0 
 0 0 0 0 0 0 ktfa  kb 0 kb hkb 0 0 0 0 kb 0 
 2 
  m18  m28  m38 0 0 0 ktss  kb 0 0 0 kb 
2 2
0 0 0 kb hkb
K 
0 0 0 0 0 0 kb 0 kx  kb hkb 0 0 0 0 kb 0 
ep

 
 0 0 0 0 0 0 hkb 0 hkb kx  h2kb 0 0 0 0 hkb 0 
 2m 2m28 2m38 0 0 0 0 kb 0 0 ky  kb hkb 0 0 0 kb 
 18

 h m18 h m28 h m38 0 0 0


2 2 2
0 hkb 0 0 hkb ky  h kb 0 0 0 hkb 
2

 
 0 0 0 0 0 0 0 0 0 0 0 0 k13 0 0 0 
667  0 0 
(59)
Pr

0 0 0 0 0 0 0 0 0 0 0 0 k13 0
 
 0 0 0 0 0 0 kb 0 kb hkb 0 0 0 0 kb 0 
 0 0 0 0 0 0 0  k 0 0  k  hk 0 0 0 k 
 b b b b 

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
668 in which

ed
mp g
669 k1, j  k1  k 2  k 3 cos  j   m1 , k 2 , j  k 4  k 5  k 6 cos  j , k13 
l 2  q132  q142

R R R R R
k1   EI e 1e  dr , k2   2   m    d    dr , k  g   m   d    dr
2 2 2

iew
0 0 r 1e 3 0 r 1e
670
R R R R R
k4   EI f 1f  dr , k5   2   m    d    dr , k  g   m   d    dr
2 2 2
0 0 r 1f 6 0 r 1f

671 The pendulum generalized force Fp is expressed as:

v
672 Fp=[0 0 0 0 0 0 0 0 0 0 0 0 F13,p F13,p 0 0]T (60)

re
673 where

q13 q13
2  13
 q13 q13  q14 q14  q 2  q142 
2
F13,P   
l  l  q13  q14
2 2 2
2
q q
2
13
2
14
674
F14,P  
q14
 q13 q13  q14 q14 
er 2

q14
2  13
q 2  q142 
l  l  q13  q14
2 2 2
2
q q
2
13
2
14
pe
675 References

676 Ancheta T. D., Darragh R. B., Stewart J. P., et al, 2014. NGA-West2 database. Earthquake Spectra, 30(3): 989-1005.
ot

677 Anh N. D., Nguyen N. X., 2012. Extension of equivalent linearization method to design of TMD for linear damped systems.

678 Structural Control and Health Monitoring, 19(6): 565-573.


tn

679 Anh N. D., Nguyen N. X., 2013. Design of TMD for damped linear structures using the dual criterion of equivalent

680 linearization method. International Journal of Mechanical Sciences, 77: 164-170.


rin

681 Atashpaz-Gargari E., Lucas C., 2007. Imperialist competitive algorithm: an algorithm for optimization inspired by imperialistic

682 competition. IEEE congress on evolutionary computation. IEEE: 4661-4667.


ep

683 Bilgili M., Yasar A., Simsek E., 2011. Offshore wind power development in Europe and its comparison with onshore

684 counterpart. Renewable and Sustainable Energy Reviews, 15(2): 905-915.


Pr

685 Brodersen M L, Bjørke A S, Høgsberg J., 2017. Active tuned mass damper for damping of offshore wind turbine vibrations.

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
686 Wind Energy, 20(5): 783-796.

ed
687 Carswell W., Johansson J., Løvholt F., et al, 2015. Foundation damping and the dynamics of offshore wind turbine monopiles.

688 Renewable energy, 80: 724-736.

iew
689 Caughey TK, 1960. Random excitation of a system with bilinear hysteresis. Trans ASME, J Appl Mech, 27(1): 649–52.

690 Christiansen S., Knudsen T., Bak T., 2011. Optimal control of a ballast-stabilized floating wind turbine. IEEE international

691 symposium on computer-aided control system design (CACSD). IEEE: 1214-1219.

v
692 Colwell S., Basu B., 2009. Tuned liquid column dampers in offshore wind turbines for structural control. Engineering

re
693 structures, 31(2): 358-368.

694 Dehghani M., Mashayekhi M., Sharifi M., 2021. An efficient imperialist competitive algorithm with likelihood assimilation
er
695 for topology, shape and sizing optimization of truss structures. Applied Mathematical Modelling, 93, 1-27.
pe
696 Ding H., Altay O., Wang J. T., 2023. Lateral vibration control of monopile supported offshore wind turbines with toroidal

697 tuned liquid column dampers. Engineering Structures, 286: 116107.

698 Dinh V. N., Basu B., 2015. Passive control of floating offshore wind turbine nacelle and spar vibrations by multiple tuned
ot

699 mass dampers. Structural Control and Health Monitoring, 22(1): 152-176.

700
tn

Eberhart R., Kennedy J., 1995. A new optimizer using particle swarm theory. MHS'95. Proceedings of the sixth international

701 symposium on micro machine and human science. IEEE: 39-43.

702
rin

Faltinsen O., 1993. Sea loads on ships and offshore structures. Cambridge university press.

703 Firestone F. A., 1933. A new analogy between mechanical and electrical systems. The Journal of the Acoustical Society of

704 America, 4(3): 249-267.


ep

705 Fitzgerald B., Basu B., 2014. Cable connected active tuned mass dampers for control of in-plane vibrations of wind turbine

706 blades. Journal of Sound and Vibration, 333(23): 5980-6004.


Pr

707 Gao Z., Tang C., Zhou X., et al, 2016. An overview on development of wind power generation. Chinese Control and Decision

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
708 Conference (CCDC). IEEE: 435-439.

ed
709 Ghaemmaghami A., Kianoush R., Yuan X. X., 2013. Numerical modeling of dynamic behavior of annular tuned liquid

710 dampers for applications in wind towers. Computer‐Aided Civil and Infrastructure Engineering, 28(1): 38-51.

iew
711 Hansen, M. O., 2015. Aerodynamics of wind turbines. Routledge.

712 Hu Y., Wang J., Chen M. Z. Q., et al, 2018. Load mitigation for a barge-type floating offshore wind turbine via inerter-based

713 passive structural control. Engineering Structures, 177: 198-209.

v
714 IEC, 2009. Wind turbines. Part 3: design requirements for offshore wind turbines. IEC61400-3 (ed. 1), Geneva, Switzerland:

re
715 International Electrotechnical Commission.

716 Ingber L., 1993. Simulated annealing: Practice versus theory. Mathematical and Computer Modelling, 18(11): 29-57.
er
717 Jonkman J., Butterfield S., Musial W., et al, 2009. Definition of a 5-MW reference wind turbine for offshore system
pe
718 development. National Renewable Energy Lab. (NREL), Golden, CO (United States).

719 Jonkman J., Michalakes J., Jonkman J.M., et al., 2013. NWTC Programmer's Handbook: A guide for software development

720 within the fast computer-aided engineering tool. Technical Report.


ot

721 Jonkman J., Musial W., 2010. Offshore code comparison collaboration (OC3) for IEA Wind Task 23 offshore wind technology

722
tn

and deployment. National Renewable Energy Lab. (NREL), Golden, CO (United States).

723 Jafarnejadsani H., Pieper J., Ehlers J., 2013. Adaptive control of a variable-speed variable-pitch wind turbine using radial-

724
rin

basis function neural network. IEEE transactions on control systems technology, 21(6): 2264-2272.

725 Jahangiri V., Sun C., 2022a. A novel three dimensional nonlinear tuned mass damper and its application in floating offshore

726 wind turbines. Ocean Engineering, 250: 110703.


ep

727 Jahangiri V., Sun C., 2022b. A novel two dimensional nonlinear tuned mass damper inerter and its application in vibration

728 mitigation of wind turbine blade. arXiv preprint arXiv:2206.14328.


Pr

729 J. P. Den Hartog, Mechanical Vibrations, 4th edn, McGraw-Hill, New York, 1956.

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
730 Kaveh A., Talatahari S., 2010. Imperialist competitive algorithm for engineering design problems. Asi. J. Civil Eng. (Build

ed
731 &.Hous).

732 Lackner M. A., Rotea M. A., 2011. Passive structural control of offshore wind turbines. Wind energy, 14(3): 373-388.

iew
733 Leng D., Wang R., Yang Y., et al, 2023. Study on a three-dimensional variable-stiffness TMD for mitigating bi-directional

734 vibration of monopile offshore wind turbines. Ocean Engineering, 281: 114791.

735 Liang F., Yuan Z., Liang X., et al, 2022. Seismic response of monopile-supported offshore wind turbines under combined

v
736 wind, wave and hydrodynamic loads at scoured sites. Computers and Geotechnics, 144: 104640.

re
737 Marian L., Giaralis A., 2014. Optimal design of a novel tuned mass-damper–inerter (TMDI) passive vibration control

738 configuration for stochastically support-excited structural systems. Probabilistic Engineering Mechanics, 38: 156-164.
er
739 Murtagh P. J., Ghosh A., Basu B., et al, 2008. Passive control of wind turbine vibrations including blade/tower interaction and
pe
740 rotationally sampled turbulence. Wind Energy: An International Journal for Progress and Applications in Wind Power

741 Conversion Technology, 11(4): 305-317.

742 Namik H., Stol K., 2010. Individual blade pitch control of floating offshore wind turbines. Wind Energy: An International
ot

743 Journal for Progress and Applications in Wind Power Conversion Technology, 13(1): 74-85.

744
tn

Nazokkar A., Dezvareh R., 2022. Vibration control of floating offshore wind turbine using semi-active liquid column gas

745 damper. Ocean Engineering, 265: 112574.

746
rin

Peri D., 2019. Hybridization of the imperialist competitive algorithm and local search with application to ship design

747 optimization. Computers & Industrial Engineering. Volume 137, 106069.

748 Pietrosanti D., De A. M., Giaralis A., 2020. Experimental study and numerical modeling of nonlinear dynamic response of
ep

749 SDOF system equipped with tuned mass damper inerter (TMDI) tested on shaking table under harmonic excitation.

750 International Journal of Mechanical Sciences, Volume 184, 105762, ISSN 0020-7403.
Pr

751 Sampson J. R. 1976. Adaptation in natural and artificial systems (John H. Holland).

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
752 Smith M. C., 2002. Synthesis of mechanical networks: the inerter. IEEE Transactions on automatic control, 47(10): 1648-

ed
753 1662.

754 Sarkar S., Fitzgerald B., 2020. Vibration control of spar‐type floating offshore wind turbine towers using a tuned mass‐

iew
755 damper‐inerter. Structural Control and Health Monitoring, 27(1): e2471.

756 Selvam K., Kanev S., Wingerden J. W. van, et al, 2009. Feedback–feedforward individual pitch control for wind turbine load

757 reduction. International Journal of Robust and Nonlinear Control: IFAC‐Affiliated Journal, 19(1): 72-91.

v
758 Si Y., Karimi H. R., Gao H., 2014. Modelling and optimization of a passive structural control design for a spar-type floating

re
759 wind turbine. Engineering structures, 69: 168-182.

760 Stewart G. M., Lackner M. A., 2014. The impact of passive tuned mass dampers and wind–wave misalignment on offshore
er
761 wind turbine loads. Engineering structures, 73: 54-61.
pe
762 Sun C., 2018a Semi-active control of monopile offshore wind turbines under multi-hazards. Mechanical Systems and Signal

763 Processing, 99: 285-305.

764 Sun C., 2018b Mitigation of offshore wind turbine responses under wind and wave loading: Considering soil effects and
ot

765 damage. Structural Control and Health Monitoring, 25(3): e2117.

766
tn

Sun C., Jahangiri V., 2018. Bi-directional vibration control of offshore wind turbines using a 3D pendulum tuned mass damper.

767 Mechanical Systems and Signal Processing, 105: 338-360.

768
rin

Sun C, Jahangiri V, 2019. Fatigue damage mitigation of offshore wind turbines under real wind and wave conditions[J].

769 Engineering Structures, 178: 472-483.

770 Sun C., Jahangiri V., Sun H., 2019. Performance of a 3D pendulum tuned mass damper in offshore wind turbines under
ep

771 multiple hazards and system variations. Smart Struct. Syst, 24(1): 53-65.

772 Zhang Z., Chen B., Hua X., 2023. Closed-form optimization of tuned mass-damper-inerter (TMDI) in flexible structures.
Pr

773 Journal of Building Engineering, Volume 72, 106554.

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628
774 Zhang Z., Fitzgerald B., 2020. Tuned mass-damper-inerter (TMDI) for suppressing edgewise vibrations of wind turbine blades.

ed
775 Engineering Structures, 221: 110928.

776 Zhang Z., Høeg C., 2020. Dynamics and control of spar-type floating offshore wind turbines with tuned liquid column dampers.

iew
777 Structural Control and Health Monitoring, 27(6): e2532.

778 Zhou J. F., Lin Y. F., 2013. Essential mechanics issues of offshore wind power systems. SCIENTIA SINICA Physica,

779 Mechanica & Astronomica, 43(12): 1589.

v
780 Zhu B., Sun C., Jahangiri V., 2021. Characterizing and mitigating ice-induced vibration of monopile offshore wind turbines.

re
781 Ocean Engineering, 219: 108406.

782 Zuo H., Bi K., Hao H., 2019. Mitigation of tower and out-of-plane blade vibrations of offshore monopile wind turbines by
er
783 using multiple tuned mass dampers. Structure and Infrastructure Engineering, 2019, 15(2): 269-284.
pe
ot
tn
rin
ep
Pr

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4578628

You might also like