You are on page 1of 25

2nd Reading

September 27, 2023 16:57 WSPC/165-IJSSD 2450123

International Journal of Structural Stability and Dynamics


(2024) 2450123 (25 pages)
c World Scientific Publishing Company
DOI: 10.1142/S0219455424501232

Failure Probability Analysis of Lattice Wind Turbine


by TONGJI UNIVERSITY on 10/22/23. Re-use and distribution is strictly not permitted, except for Open Access articles.

Towers Under Combined Wind and Earthquake Action

Jia-Xiang Li ∗,†,‡ , Ling-Peng Wang ∗ , Zhuo-Qun Zhang§ ,

Yi Zhao†,¶, and Zhi-Qian Dong‡


∗Department of Civil Engineering
Northeastern University
Shenyang 110819, Shenyang, P. R. China
Int. J. Str. Stab. Dyn. Downloaded from www.worldscientific.com

†Key Laboratory of Disaster Prevention and Structural


Safety of Ministry of Education
Guangxi University, Nanning 530004
Guangxi, P. R. China
‡State Key Laboratory of Coastal and Offshore Engineering
Dalian University of Technology
Dalian, Liaoning 116024, P. R China
§State Grid Economic and Technological Research
Institute Co Ltd, Beijing 102209, P. R. China
¶Schoolof Civil Engineering and Architecture
Guangxi University
Nanning 530004, Guangxi, P. R. China
zhaoyidut@126.com

Received 2 July 2023


Accepted 19 August 2023
Published 28 September 2023

With wind farms expanding into seismically active areas, wind turbines face the combined
hazard of wind and earthquakes. Strong winds often accompany strong earthquakes,
posing a significant threat to wind turbine safety. When subjected to multihazard loads,
ignoring the correlation between loads could underestimate the risk to the structure.
However, there is limited research on wind turbines under the combined effect of wind and
earthquakes. In this paper, based on historical data considering the correlation between
wind and earthquakes, a failure probability analysis method for wind turbine towers
under the joint action of wind and earthquakes is presented. To accurately simulate
the mechanical behavior of lattice-type wind turbine towers, the bolt slippage effect is
taken into account. A finite element model of a lattice wind turbine tower is established,
and the failure probability of the tower under the joint action of wind and earthquakes
is calculated. The study also analyzes the influence of the bolt slippage effect and load

 Corresponding author.

2450123-1
2nd Reading
September 27, 2023 16:57 WSPC/165-IJSSD 2450123

J.-X. Li et al.

correlation on the reliability of lattice-type wind turbine towers. The results indicate that
ignoring load correlation can underestimate the failure probability of the structure, thus
decreasing its safety. Similarly, ignoring the bolt slippage effect can also underestimate
the failure probability of the structure and seriously threaten the safety of wind turbines.

Keywords: Wind turbine towers; fragility; failure probability; bolt slippage; multi-
disaster.
by TONGJI UNIVERSITY on 10/22/23. Re-use and distribution is strictly not permitted, except for Open Access articles.

1. Introduction
To mitigate the environmental problems caused by traditional thermal power gener-
ation, wind energy has gained significant attention as a clean and renewable energy
source.1 With the advancement of wind power technology, wind turbine towers have
become taller, and blade sizes have increased.2 This has resulted in wind turbine
Int. J. Str. Stab. Dyn. Downloaded from www.worldscientific.com

towers becoming more slender and more flexible, making them more susceptible to
lateral action such as earthquakes and wind loads.3 The demand for wind power
has been increasing globally, resulting in more wind farms being located in seismi-
cally active areas.4–6 For instance, the Alta wind farm in California is only 96 km
from the San Andreas Fault, while the Kameda wind farm in Japan is situated in
Fukushima Prefecture in the Pacific Rim seismic zone. Existing research shows that
strong earthquakes are often accompanied by strong winds due to atmospheric pres-
sure changes during earthquakes.7–9 Therefore, wind turbines located in seismically
active areas experience the combined action of wind and earthquakes. This means
that these turbines are more dangerous than wind or earthquakes acting alone.6,7
Consequently, it is necessary to study the risk of wind turbines under the combined
action of wind and earthquakes to ensure their safe operation.
Due to the threat of wind and earthquake coupling to wind turbine towers,
many scholars have been attracted to this topic. Asareh et al.10 established a finite
element model of a cylindrical wind turbine tower using ABAQUS and analyzed the
fragility of the wind turbine tower under the simultaneous action of earthquake and
wind. They found that earthquake excitation played a dominant role in the fragility
of wind turbine towers. In addition, when wind speeds approached the rated wind
speeds, lower intensity earthquake actions could also lead to tower collapse. Yuan
et al.11 used the FAST program to study the dynamic response of large-scale wind
turbines considering the combination of seismic and aerodynamic loads. They ana-
lyzed the fragility of wind turbines under normal operation and parked cases. The
results showed that because of aerodynamic damping during wind turbine opera-
tion, the failure probability of towers was lower than that when the wind turbines
were parked. Therefore, maintaining the operation of the baseline control system
during an earthquake can effectively improve the structural reliability. Martı́n and
Pozos-Estrada12 analyzed the fragility of a 5-MW onshore wind turbine in south-
ern Mexico under the coupling action of strong wind and seismic for parked and
operating cases. The results showed that the fragility of the structure was not only

2450123-2
2nd Reading
September 27, 2023 16:57 WSPC/165-IJSSD 2450123

Failure Analysis of Lattice Wind Towers

related to external loads but also affected by the working state of the wind turbine:
seismic action played a dominant role when the wind turbine was parked, while
wind loads played a dominant role when the wind turbine was in operation. Martı́n
et al.13 used ANSYS to establish a model and study the effect of a passive damping
system on the fragility of wind turbine towers under the simultaneous action of
cyclones and earthquakes. The results showed that the mass of the damping system
by TONGJI UNIVERSITY on 10/22/23. Re-use and distribution is strictly not permitted, except for Open Access articles.

played an important role in controlling the dynamic response of the towers. Asareh
et al.14 studied the effect of the combined action of wind and earthquakes on the
power generation efficiency of a 5-MW wind turbine generator. They found that
earthquake intensity had a significant influence on the power generation efficiency
of wind turbines. However, the above studies were limited to fragility analysis and
did not consider the influence of the probability distribution of wind or earthquakes.
The correlation between wind and earthquakes was also not considered.
Int. J. Str. Stab. Dyn. Downloaded from www.worldscientific.com

To consider the combined load action on structures during service life, current
engineering design codes often use load combination methods, such as the Ferry–
Borges model, the load combination method proposed by Wen, and the Copula func-
tion joint probability model.15 Among these methods, the copula function model
considers the correlation between variables and is easy to operate.16,17 Among the
various copula functions, the Archimedean copula function is more flexible and
has no restrictions on the marginal distribution. It has been widely used in stud-
ies of multiple hazard coupling effects, such as earthquake coupling with strong
winds, earthquake coupling with floods, and wind coupling with rain.16–19 Zheng
et al.20 established a hazard probability model of wind and earthquake occurring
simultaneously based on the Copula function and used this model to calculate the
annual failure probability of high-rise buildings under the combined action of wind
and earthquake. The results showed that considering the correlation between wind
speed and earthquake, the probability of hazard occurring was closer to the real
situation, and the failure probability of the structure was higher. Therefore, it is
necessary to consider the correlation between wind and earthquakes when calcu-
lating the failure probability of structures under the combined action of wind and
earthquakes. Otherwise, the calculation results will be biased toward danger. How-
ever, there is currently a lack of research on the influence of load correlation on
the safety of wind turbine towers. Therefore, it is necessary to conduct research on
the failure probability of wind turbine towers under coupled wind-seismic action,
considering the correlation between wind and earthquakes.
The angle steel lattice tower is a commonly used support structure for wind tur-
bines, with bolted connections between members. Bolt slippage occurs in the joints
of lattice towers during the load process due to the gap between bolt holes and
bolt rods. The results of full-scale tests have shown that bolt connection slippage
significantly affects the response of lattice towers.21,22 Many scholars have studied
the influence of bolt slippage on lattice tower structures under static loads through

2450123-3
2nd Reading
September 27, 2023 16:57 WSPC/165-IJSSD 2450123

J.-X. Li et al.

experiments or numerical simulations.23–25 The results showed that bolt slippage not
only reduces the lateral stiffness of the tower body but also affects the structural
failure mode. Shooshatari and Yaghoobi26,27 studied the load-deformation curve
of bolt slippage of wind turbine towers based on cyclic loading test results and
analyzed the influencing factors of bolt slippage through parameter analysis. The
results showed that bolt slippage would reduce the stiffness of the tower, causing
by TONGJI UNIVERSITY on 10/22/23. Re-use and distribution is strictly not permitted, except for Open Access articles.

a significant increase in structural deformation. Nezamolmolki et al.28 introduced


a bolt node slippage model into the numerical model of lattice steel wind turbine
towers for static analysis. They found that bolt slippage significantly reduced the
frequencies of wind turbine towers at all levels and increased the displacement of
the tower top under external loads. However, most researchers have focused on the
influence of bolt slippage on the response of lattice tower structures under static
loads. Meanwhile, research on the influence of bolt slippage on the response of lat-
Int. J. Str. Stab. Dyn. Downloaded from www.worldscientific.com

tice tower structures under dynamic loads is also essential. Li et al.29 conducted
experimental studies on the mechanical model of bolt slippage under reciprocating
loads and established a numerical model of a lattice tower considering bolt recip-
rocating slip effects. They investigated the influence of bolt slippage effects on the
seismic response of transmission towers under earthquakes. The results showed that
neglecting bolt slippage effects would significantly overestimate the load-bearing
capacity of lattice towers. Although transmission towers and lattice wind turbine
towers share similarities in structure, the external loads they are subjected to are
vastly different. Currently, there is a lack of research on the effects of bolt slippage
on wind turbine towers under dynamic loads.
The above-mentioned papers reveal that research on the failure probability anal-
ysis of wind turbine towers under the joint effect of wind and earthquakes, consid-
ering load correlation, is limited. Therefore, it is necessary to study the evaluation
method of wind turbine tower failure probability, taking into account the correlation
between wind and earthquakes. Furthermore, accurately simulating the dynamic
response of wind turbine towers requires considering the influence of bolt slippage
effects on the mechanical characteristics of wind turbine towers.

2. Method
A wind turbine tower failure probability evaluation method under wind-seismic cou-
pling action considering the correlation between wind and earthquakes is proposed,
as shown in Fig. 1. The method consists of three parts: establishing a joint hazard
probability model, fragility analysis, and failure probability calculation.
The first part involves collecting historical data and performing regression anal-
ysis to establish a single hazard probability distribution. The optimal distribution
is selected based on R2 and RMSE, where a value closer to 1 for R2 and a smaller
value for RMSE indicate a better fitting effect. Four Archimedean copula functions

2450123-4
2nd Reading
September 27, 2023 16:57 WSPC/165-IJSSD 2450123

Failure Analysis of Lattice Wind Towers


by TONGJI UNIVERSITY on 10/22/23. Re-use and distribution is strictly not permitted, except for Open Access articles.

Fig. 1. Flow chart for calculating the failure probability of wind turbine towers.
Int. J. Str. Stab. Dyn. Downloaded from www.worldscientific.com

(Gumbel copula, Clayton copula, AMH copula, and Frank copula) are selected to
construct the wind-seismic copula joint probability distribution model based on the
single hazard probability model. The optimal Copula function is selected based on
the minimum Akaike Information Criterion (AIC) and Bayesian Information Crite-
rion (BIC).
The second part involves calculating the exceeding probability of the structure
reaching a certain damage limit state under a specific hazard condition. To accu-
rately simulate the dynamic response of lattice-type wind turbine towers under wind
and seismic coupling, this part considers the effect of bolt slippage on the tower’s
dynamic response.
The third part combines fragility with the joint probability distribution to cal-
culate the failure probability of the structure using Eq. (2.1)

P (EDP > C) = P (EDP > C | IM1 , IM2 )dF (IM1 , IM2 ), (2.1)

where P (EDP > C) represents the failure probability of the structure, P (EDP >
C | IM1 , IM2 ) represents the fragility function of the wind turbine tower under wind-
seismic coupling action, and F (IM1 , IM2 ) represents the joint hazard probability
model. The uncertainty of the occurrence probability of the hazard is not considered
in the fragility analysis. By simultaneously considering fragility analysis and the
joint hazard probability model through Eq. (2.1), the structural failure probability
can be more accurately evaluated. This method has been widely used in multiple
hazard coupling research fields, such as high-rise frame structures under strong
wind–seismic coupling and transmission lines under wind-rain coupling.19,20 In this
study, the fragility function is combined with the copula function to provide a
calculation method for the failure probability of wind turbine towers during the
design life under wind–seismic coupling.

2450123-5
2nd Reading
September 27, 2023 16:57 WSPC/165-IJSSD 2450123

J.-X. Li et al.

3. Wind–Seismic Joint Probability Model


A united probabilistic model of wind and earthquake is established to consider
the joint effects of wind and earthquake hazards. First, historical data of wind
speeds and peak ground acceleration (PGA) are collected for a specific location,
and the probability density function (PDF) of a single hazard is established using
the probability distribution function that best fits the historical data. Then, the
by TONGJI UNIVERSITY on 10/22/23. Re-use and distribution is strictly not permitted, except for Open Access articles.

exceeding probability function of each single hazard is calculated according to the


following equation:

F (IM > x) = 1 − Fcum (x), (3.1)

where F (IM > x) represents the probability that the hazard intensity measure IM is
greater than a given value x; Fcum (x) represents the cumulative distribution function
Int. J. Str. Stab. Dyn. Downloaded from www.worldscientific.com

(CDF) of a single hazard, which is obtained by integrating the PDF. Subsequently,


the exceeding probability functions of wind and earthquake are combined using the
Copula function to establish a joint exceeding probability distribution model as a
hazard risk model.

3.1. Wind speed exceeding probability model


Historical wind speed data are obtained from the website of the National Oceanic
and Atmospheric Administration (NOAA). These data include the daily maximum
wind speed at 10 m above the ground from 1970 to 2022, resulting in a total of
18 732 wind speed data points. Four PDFs, including the lognormal, Gaussian,
Gumbel, and Weibull distributions, are used to model the probability distribution
of wind speed, as shown in Fig. 2(a). The coefficients of these models, as shown in

(a) Probability density distribution (b) Exceeding probability

Fig. 2. Probability model of wind speed.

2450123-6
2nd Reading
September 27, 2023 16:57 WSPC/165-IJSSD 2450123

Failure Analysis of Lattice Wind Towers

Table 1. Wind speed regression analysis results.

PDF a b R2 RMSE
Lognormal 2.645 0.214 0.822 0.0203
Gaussian 13.82 2.989 0.7907 0.0234
Gumbel 13.27 2.681 0.7788 0.0241
Weibull 2e-07 5.514 0.8343 0.01839
by TONGJI UNIVERSITY on 10/22/23. Re-use and distribution is strictly not permitted, except for Open Access articles.

Table 1, are fitted according to the historical data. The coefficient of R2 and root
mean square error (RMSE) are also calculated to select the best-fitted PDF. From
Table 1, the Weibull model has the R2 value closest to 1 and the smallest RMSE
value. Therefore, the Weibull model is chosen as the probability model for wind
speed, and its exceeding probability distribution is shown in Fig. 2(b).
Int. J. Str. Stab. Dyn. Downloaded from www.worldscientific.com

3.2. Seismic exceeding probability model


The PGA serves as the intensity measure of seismic action. The earthquake infor-
mation is obtained from the earthquake catalog established by the U.S. Geological
Survey (USGS). These earthquakes occurred between 1972 and 2022 with a magni-
tude greater than four and an epicentral distance less than 300 km from the target
location. A total of 832 PGA values are calculated based on the ground motion
prediction equations (GMPE) of the NGA-West2 project using the information col-
lected from the earthquake catalog.30 The statistical results are shown in Fig. 3(a).
Four PDFs identical to those in Sec. 3.1 are used to fit the PGA data. The
parameters of each PDF and goodness-of-fit statistics are presented in Table 2.
According to Table 2, the Gumbel distribution model has the highest goodness-of-
fit, and thus, the Gumbel distribution function is selected as the probability model
of the PGA. The exceeding probability distribution of PGA is shown in Fig. 3(b).

(a) Probability density distribution (b) Exceeding probability

Fig. 3. Probability model of PGA.

2450123-7
2nd Reading
September 27, 2023 16:57 WSPC/165-IJSSD 2450123

J.-X. Li et al.

Table 2. PGA regression analysis results.

PDF a b R2 RMSE
Lognormal 2.695 0.2075 0.8999 0.01548
Gaussian 14.61 3.174 0.8967 0.02195
Gumbel 13.85 2.644 0.9216 0.01348
Weibull 7e-07 5.141 0.8080 0.02110
by TONGJI UNIVERSITY on 10/22/23. Re-use and distribution is strictly not permitted, except for Open Access articles.

3.3. Wind-seismic joint exceeding probability model


The correlation between earthquakes and wind hazards during the service life of
structures is considered based on copula theory. A coupled probability distribution
model of earthquakes and wind is constructed using the single-hazard probability
model established in the previous sections. Four Archimedean copula functions,
Int. J. Str. Stab. Dyn. Downloaded from www.worldscientific.com

namely, the Gumbel copula, Clayton copula, AMH copula, and Frank copula, are
selected in this study, as shown in Table 3.
The probability distributions of PGA and wind speed V are not independent but
have a certain correlation with each other.20 The value of parameter θ of the copula
functions needs to be determined based on historical data to consider the correlation
between wind speed V and PGA. The data of the same-day recorded PGA and wind
speed V are extracted to calculate the Kendall rank correlation coefficient τ .17 After
determining τ , the coefficient θ can be calculated based on the relationship between
τ and θ, as shown in Table 4. Afterwards, the best-fitting Copula function is selected
based on the minimum AIC and minimum BIC. According to the calculation results
shown in Table 4, the Clayton copula function has the smallest AIC and BIC values,

Table 3. Bivariate Archimedean Copula function.

Copula C(u, v) The relationship between τ and θ The range of θ


1
θ
GH e−[(−lnu1 ) +(−lnu2 )θ ] θ τ =1− 1
[1, ∞)
θ
−1 θ
Clayton (u−θ + v−θ − 1) θ τ = θ+2
(0, ∞)
» – h i
(e−θu −1)(e−θv −1) R
Frank − θ1 ln 1+ e−θ−1
τ =1− 4
θ
− 1θ θ0 t
exp(t)−1
dt −1 (−∞, ∞)\0

uv
` 2
´ 2
` ´
1 2
AMH 1−θ(1−u)(1−v)
τ = 1− 3θ
− 3
1− θ
ln(1 − θ) [−1, 1)

Table 4. The values of θ, AIC, and BIC


for the adopted Achimedean copulas.

Copula θ AIC BIC


GH 1.043 3052.46 3055.88
Clayton 0.086 2787.90 2791.31
Frank 2.893 3012.74 3010.29
AMH 0.177 3058.62 3062.01

2450123-8
2nd Reading
September 27, 2023 16:57 WSPC/165-IJSSD 2450123

Failure Analysis of Lattice Wind Towers


by TONGJI UNIVERSITY on 10/22/23. Re-use and distribution is strictly not permitted, except for Open Access articles.
Int. J. Str. Stab. Dyn. Downloaded from www.worldscientific.com

Fig. 4. Joint exceeding Probability Model

and thus, the Clayton copula function is used to establish the wind-earthquake
coupled probability distribution model, which is expressed as follows:
1
F (IM1 , IM2 ) = C(u, v) = (u−θ + v −θ − 1)− θ , (3.2)
where C(u, v) represents the Clayton Copula function, u represents the exceeding
probability function of PGA, which is a Gumbel distribution, and v represents the
exceeding probability function of wind speed, which is a Weibull distribution. The
joint exceeding probability function is plotted in Fig. 4, which shows that the joint
probability value is relatively large for small PGA and small wind speed values due
to the influence of the probability distributions of PGA and wind speed.

4. Fragility Analysis of Wind Turbines


4.1. Fragility analysis method
In the fragility analysis of wind turbine towers when subjected to hazards such as
earthquakes and wind, the ultimate displacement of the tower top is usually used to
assess the tower’s carrying capacity.3,4,10 In this study, the maximum displacement
of the tower top is also described as the engineering demand parameter (EDP),
while the mean value of the EDP can be expressed as
ln(EDP ) = aln(IM1 ) + bln(IM2 ) + c, (4.1)
where IM1 and IM2 represent the intensity measures of the hazards, which corre-
spond to the mean wind speed V and PGA in m/s and g, respectively. Equation (4.1)
is known as the structural demand model, as it describes the relationship between
the median value of EDP and the intensity measures of the hazards.

2450123-9
2nd Reading
September 27, 2023 16:57 WSPC/165-IJSSD 2450123

J.-X. Li et al.

Assuming that the EDP follows a log-normal distribution in both earthquake


and wind engineering.20 The fragility of the wind turbine structure under a sudden
disaster can be expressed mathematically as
 
ln(C) − ln(EDP)
Pf (EDP > C | PGA, V) = 1 − Φ  , (4.2)
2
βEDP 2
+ βC

by TONGJI UNIVERSITY on 10/22/23. Re-use and distribution is strictly not permitted, except for Open Access articles.


[ln(di ) − ln(EDP)]2
βEDP = , (4.3)
n−2
where Φ represents the standard normal CDF; C represents the carrying capac-
ity of the tower; di represents the maximum value of the tower top displacement;
βEDP represents the logarithmic standard deviation of the engineering requirement
parameter EDP; and βC denotes the inherent uncertainty of the structural capacity.
Int. J. Str. Stab. Dyn. Downloaded from www.worldscientific.com

4.2. Finite element model


A 2.0 MW 60 m high wind turbine is taken as an example. The dimensional
schematic of the tower is shown in Fig. 5, and a summary of the general char-
acteristics of the turbine is presented in Table 5. The tower structure is modeled
and analyzed using ANSYS, where the main members of the tower is Q420 with
an elastic modulus of 210 GPa, Poisson’s ratio of 0.3, and a density of 7850 kg/m3 .

Fig. 5. Wind turbine tower size diagram and finite element model.

2450123-10
2nd Reading
September 27, 2023 16:57 WSPC/165-IJSSD 2450123

Failure Analysis of Lattice Wind Towers

Table 5. Parameters of 2.0 MW wind


turbine tower.
Properties Value
Rating 2.0 MW
Cut-in wind speed 3 m/s
Rated wind speed 12 m/s
by TONGJI UNIVERSITY on 10/22/23. Re-use and distribution is strictly not permitted, except for Open Access articles.

Cut-out wind speed 25 m/s


Hub height 60.50 m
Blade length 38.75 m
Rotor mass 33 640 kg
Nacelle mass 52 300 kg
Elastic modulus 210 GPa
Poisson’s ratio 0.3
Density of tower 7850 kg/m3
Int. J. Str. Stab. Dyn. Downloaded from www.worldscientific.com

To simplify the calculation and improve the computational efficiency, the rotor and
nacelle are modeled as two lumped masses by employing the elements MASS 21,
as shown in Fig. 5.5,11,12 The tower members and the blades are modeled by using
element BEAM 188. The mass distribution of the blade model is referenced to the
actual blades to realistically simulate the inertial forces of the blades during the
dynamic analysis.
To obtain the accurate dynamic response of the wind turbine tower, the bolt
slippage effect is considered in the finite element modeling of the wind turbine
tower in ANSYS. The finite element model which includes the bolt slippage effect
is established by adding the nonlinear spring elements COMBIN39 to the nodes of
the original tower numerical model, as shown in Fig. 6. To make the added spring
elements simulate the mechanical behavior of the bolt slippage under reciprocating
load, the hysteresis curves of bolt slip are obtained by numerical simulation. Then,
the force-displacement skeleton curves of bolt slip are extracted according to the

Fig. 6. Schematic diagram of the bolt-slippage model.

2450123-11
2nd Reading
September 27, 2023 16:57 WSPC/165-IJSSD 2450123

J.-X. Li et al.
by TONGJI UNIVERSITY on 10/22/23. Re-use and distribution is strictly not permitted, except for Open Access articles.
Int. J. Str. Stab. Dyn. Downloaded from www.worldscientific.com

Fig. 7. Schematic diagram of the tower with bolt slippage models.

hysteresis curves. The parameters of the spring element COMBIN39 are set based
on the skeleton curves of the nodes so that the spring element can simulate the
reciprocal sliding of the bolted nodes.29 The bolt slippage effect of three kinds
of nodes, including main members to main members, main members to auxiliary
members and auxiliary members to auxiliary members, is considered. A schematic
diagram of the tower with bolt slippage models is shown in Fig. 7.
Table 6 lists the vibration frequencies of the wind turbine tower in different
directions of orders 1 and 2. As shown in Table 6, the frequencies of each order
of the tower decreased after bolt slippage is considered. This is because the bolt
slippage effect decreases the stiffness of the wind turbine tower, and the frequency of

Table 6. Parameters of 2.0MW wind turbine.


Mode Without joint-slippage With joint-slippage
1st tower fore-aft 0.815 Hz 0.587 Hz
1st tower side-side 0.812 Hz 0.588 Hz
2nd tower fore-aft 5.148 Hz 2.541 Hz
2nd tower side-side 5.524 Hz 3.097 Hz

2450123-12
2nd Reading
September 27, 2023 16:57 WSPC/165-IJSSD 2450123

Failure Analysis of Lattice Wind Towers


by TONGJI UNIVERSITY on 10/22/23. Re-use and distribution is strictly not permitted, except for Open Access articles.
Int. J. Str. Stab. Dyn. Downloaded from www.worldscientific.com

Fig. 8. Modal shapes of wind turbine tower.

the tower also decreases. The vibration modes associated with the tower are shown
in Fig. 8.

4.3. Aerodynamic loads


To simulate the aerodynamic loads, a three-dimensional turbulent wind field is
generated by the software GH BLADED, with statistical properties representative
of real atmospheric turbulence. The wind field is discretized into a finite number
of grid points in the longitudinal and vertical directions, as shown in Fig. 9. The
turbulence generation at each point adopts the von Karman spectrum. The spectral
density of the longitudinal turbulence component is given by
nSuu (n) 4nu
2
= 5 , (4.4)
σu (1 + 70.8n2u ) 6
where Suu is the autospectrum of wind speed variation; n is the frequency of vari-
ation; σu2 is the standard deviation of wind speed variation, and nu is a nondimen-
sional frequency parameter given by
nx L u
nu = , (4.5)
U

2450123-13
2nd Reading
September 27, 2023 16:57 WSPC/165-IJSSD 2450123

J.-X. Li et al.
by TONGJI UNIVERSITY on 10/22/23. Re-use and distribution is strictly not permitted, except for Open Access articles.
Int. J. Str. Stab. Dyn. Downloaded from www.worldscientific.com

Fig. 9. Grid discretization of turbulence field.

where x Lu is the length scale of longitudinal turbulence and U is the mean wind
speed.
The wind shear model uses the exponential model to define the change in the
steady-state mean wind speed with height
 α
h
V (h) = Vhub , (4.6)
hhub
where h is the height above the ground, α is the power law exponent (assumed to
be 0.2), and Vhub is the mean wind speed at the hub height.
The aerodynamic loads on the wind turbine blade are calculated using blade ele-
ment momentum (BEM) theory. Each blade is divided into several blade elements,
and at each radial position, the rate of change of axial and tangential momentum is
equal to the thrust and torque generated by each blade element.32 The thrust dT
and torque dQ produced by a blade element with a length of dr and located at a
distance of r from the rotor center can be given by the following equations:
1
dT = ρV 2 (CL cos ∅ + CD sin ∅)cdr, (4.7)
2
1
dQ = ρV 2 r(CL sin ∅ − CD cos ∅)cdr, (4.8)
2
where ρ is the air density, V is the magnitude of the apparent wind velocity vector
at the blade element, ∅ is known as the inflow angle and defines the direction of the
apparent flow speed vector relative to the plane of rotation of the blade, as shown

2450123-14
2nd Reading
September 27, 2023 16:57 WSPC/165-IJSSD 2450123

Failure Analysis of Lattice Wind Towers


by TONGJI UNIVERSITY on 10/22/23. Re-use and distribution is strictly not permitted, except for Open Access articles.
Int. J. Str. Stab. Dyn. Downloaded from www.worldscientific.com

Fig. 10. Force acting on the blade cross-section.

in Fig. 10; c is the chord of the blade element, and CL and CD are the lift and drag
coefficients, respectively. Finally, the total forces and moments acting on the rotor
can be calculated by summing all forces and moments along all blade spans. Then,
they are applied to the rotor mass points of the finite element model to simulate
the wind turbine operating loads.
To calculate the wind loads applied on the tower, the wind turbine tower is
divided into different sections along the tower height. The wind loads on each section
are applied as a concentrated force, calculated as follows:
1
F = ρCd V02 (t)A, (4.9)
2
where ρ is the air density; Cd is the drag coefficient; A is the windward area of a
section, and V0 (t) is the relative wind speed, taken as the wind speed time range
at the grid division point at the center point of the node. The wind loads on the
tower are applied on the nodes of the main legs at the corresponding heights.

4.4. Seismic action simulation


According to the recommendation of the FEMAP695 report, 80 seismic records,
each containing two horizontal components, are selected from the PEER NGA
database.35 To match the spectral characteristics of the selected seismic records
with the target response spectrum of the selected site, SeismoMatch software is
used to match each seismic wave with the target spectrum at the location of the
tower. The relationship between the square root of the 5% damped response spec-
trum acceleration (Sa,5%) and the target spectrum for each record is shown in
Fig. 11. As the interaction between the soil and structure is not considered in this
study, the bottom of the wind turbine tower is simplified as a fixed connection.31

2450123-15
2nd Reading
September 27, 2023 16:57 WSPC/165-IJSSD 2450123

J.-X. Li et al.
by TONGJI UNIVERSITY on 10/22/23. Re-use and distribution is strictly not permitted, except for Open Access articles.
Int. J. Str. Stab. Dyn. Downloaded from www.worldscientific.com

Fig. 11. Response spectra from some of the ground motion records simulated.

4.5. Estimation of limit states


To evaluate the safety of the wind turbine tower under earthquake and wind loads,
two limit states are defined based on the structural response: the serviceability
limit state (SLS) and the ultimate limit state (ULS). When the displacement at
the top of the tower exceeds 1.25% of the tower height, the SLS is considered to
be reached,10–12 as this can affect the power generation of the wind turbine and
even result in the failure of some components of the turbine. The ULS is defined
as the displacement at the top of the tower exceeding the buckling displacement,
which is determined using static pushover analysis. The static pushover analysis
is conducted by gradually increasing the horizontal force applied to the top of the
tower.10,11 The load-displacement curve is shown in Fig. 12. It can be seen from
Fig. 12 that the inflection points of the curves are very obvious and these inflection
points are defined as buckling points. The limit state parameters for each limit state
are listed in Table 7. Considering the slip effect, the buckling displacement of the
structure increased by 137.9%. However, the difference in the buckling load between
the two cases did not exceed 5%. The results indicate that the node slip effect
significantly affects the deformation of the tower under static pushover analysis
but has little effect on the bearing capacity of the tower, which is consistent with
previous studies.21,22

4.6. Fragility analysis under combined wind and earthquake action


To simulate the action of wind and earthquakes, dynamic time history analysis is
performed under wind and seismic coupling. The total duration of each analysis

2450123-16
2nd Reading
September 27, 2023 16:57 WSPC/165-IJSSD 2450123

Failure Analysis of Lattice Wind Towers


by TONGJI UNIVERSITY on 10/22/23. Re-use and distribution is strictly not permitted, except for Open Access articles.
Int. J. Str. Stab. Dyn. Downloaded from www.worldscientific.com

Fig. 12. Pushover analysis curves.

Table 7. Limit states of the top displacement of the tower.

Damage state With joint-slippage Without joint-slippage


SLS 0.763 0.763
ULS 0.812 1.832

is set as 300 s, with wind load acting alone for the first 140 s and seismic action
starting from 140 s. As IEC 61400-1 and other research suggested, to eliminate
the start-up inaccuracy caused by transient effects, the results of the first 50 s are
excluded from all nonlinear time history analyses.12,33
In the dynamic time history analysis, the Rayleigh damping is considered. For
the wind turbine in shutdown condition, the total damping is set to 1% because
only the structural damping and the soil damping are considered.10,12 Whereas for
the wind turbine in operating condition, the total damping is assumed to be 5%
due to the aerodynamic damping caused by the rotating blades.10,12,14
To consider more wind-seismic combinations, a Monte Carlo simulation is used
to randomly generate 80 average wind speeds between 3 m/s and 25 m/s using
MATLAB. Then, 80 wind speeds are combined with 80 earthquake ground motions
randomly using MATLAB, resulting in 80 sets of load combinations, as shown in
Fig. 13. For all 80 load combinations, dynamic time history analysis is performed
on the models with/without considering bolt slippage. The parameters of the engi-
neering demand model in Eq. (4.1) are obtained through regression analysis, as
presented in Table 8. The demand model is established as shown in Fig. 14. The

2450123-17
2nd Reading
September 27, 2023 16:57 WSPC/165-IJSSD 2450123

J.-X. Li et al.
by TONGJI UNIVERSITY on 10/22/23. Re-use and distribution is strictly not permitted, except for Open Access articles.
Int. J. Str. Stab. Dyn. Downloaded from www.worldscientific.com

Fig. 13. Random combination of wind speed and seismic intensity.

Table 8. Regression analysis results of EDPs.

Model a b c R2 RMSE
With joint-slippage −0.612 1.020 0.141 0.7372 0.3156
Without joint-slippage −1.069 1.019 0.064 0.8047 0.2579

fitting results in Fig. 14 and the determination coefficient R2 in Table 8 show that
the two-dimensional linear equation has good accuracy in simulating the EDPs. By
bringing Eq. (2.1) into Eqs. (4.2) and (4.3), the failure probability of the structure
under given combined wind and seismic action, i.e. fragility, can be calculated. The
fragility surfaces calculated for each load combination are shown in Fig. 15, indi-
cating that seismic action has a controlling effect on the fragility of wind turbine
towers.
To show the fragility differences of the structure under different load combi-
nations more clearly, the seismic fragility curves corresponding to wind speeds of
10 m/s and 25 m/s are plotted as shown in Fig. 16. The fragility curves for each limit
state shift to the left when bolt slippage effects are considered, indicating that the
failure probability is higher under given wind and seismic loads when bolt slip effects
are considered. The effect is most significant for the SLS because considering bolt
slip effects results in a significant increase in the deformation of the tower under
external loads, leading to a higher probability of wind turbine generator failure.
Therefore, if bolt slip effects are ignored in the design phase, the danger of external
loads to the normal operation of wind turbine generators will be underestimated,
resulting in an unsafe design.

2450123-18
2nd Reading
September 27, 2023 16:57 WSPC/165-IJSSD 2450123

Failure Analysis of Lattice Wind Towers


by TONGJI UNIVERSITY on 10/22/23. Re-use and distribution is strictly not permitted, except for Open Access articles.

(a) Considering bolt slippage (b) Not considering bolt slippage

Fig. 14. Engineering demand model.


Int. J. Str. Stab. Dyn. Downloaded from www.worldscientific.com

(a) ULS without bolt slippage (b) SLS without bolt slippage

(c) ULS without bolt slippage (d) SLS without bolt slippage

Fig. 15. Fragile surfaces.

2450123-19
2nd Reading
September 27, 2023 16:57 WSPC/165-IJSSD 2450123

J.-X. Li et al.
by TONGJI UNIVERSITY on 10/22/23. Re-use and distribution is strictly not permitted, except for Open Access articles.

(a) V = 10 m/s (b) V= 25 m/s


Int. J. Str. Stab. Dyn. Downloaded from www.worldscientific.com

Fig. 16. Fragility curves for a given PGA and V .

The fragility surfaces are horizontally cut at a probability of 10%, and the curves
that intersect with them are defined as the critical failure curves, i.e. the load com-
binations that exceed the critical failure curves are considered to cause structural
failure.34 The critical failure curves for each limit state are shown in Fig. 17. When
considering bolt slip effects, each curve shifts to the left, indicating a decrease in the
bearing capacity of the tower. Especially for the SLS, when the wind speed is con-
stant, the PGA on the critical failure curve decreases by nearly 60% when bolt slip
effects are considered. This shows that ignoring bolt slip effects will overestimate

Fig. 17. Critical curves at 10% fragility.

2450123-20
2nd Reading
September 27, 2023 16:57 WSPC/165-IJSSD 2450123

Failure Analysis of Lattice Wind Towers


by TONGJI UNIVERSITY on 10/22/23. Re-use and distribution is strictly not permitted, except for Open Access articles.

(a) Wind acting only (b) Wind and earthquake both


Int. J. Str. Stab. Dyn. Downloaded from www.worldscientific.com

Fig. 18. Time history curves of tower top displacement.

the limit loads for wind turbine generators during normal operation, which is very
unfavorable for wind turbine generators.
As shown in Fig. 18, the time history curves of the tower top displacement along
the fore-aft direction are plotted, where the PGA is 1.01 g and the average wind
speed is 17.13 m/s. Figure 18(a) shows the time history curves of tower top displace-
ment during the period from 50 s to 80 s with only wind load acting. The vibration
amplitude of the tower top displacement increases significantly after considering
the bolt slippage effect. Additionally, it can be seen in Fig. 18(b) that the vibration
amplification is greater after the seismic action is applied from 140 s. This indicates
that the bolt slippage effect on the tower under the combined action of wind and
earthquake is greater than that under the wind load only.

5. Calculation of the Failure Probability


The fragility of wind turbine towers under wind and seismic coupling action is
calculated in Sec. 4. However, fragility only represents the failure probability of
towers under certain situations, and the uncertainty and the correlation of external
loads are not considered. To calculate the failure probability of wind turbine towers
more accurately, the load joint probability model established in Sec. 3 and the
fragility in Sec. 4 are combined to calculate the failure probability of wind turbine
towers through Eq. (2.1). The results are shown in Fig. (19).
To analyze the influence of load correlation on the failure probability of wind
turbine towers, the failure probability without considering load correlation is also
calculated for comparison. Specifically, the copula function is substituted by the
product of the marginal distribution functions, as shown in the following equation:

F (IM1 , IM2 ) = u · v, (5.1)

2450123-21
2nd Reading
September 27, 2023 16:57 WSPC/165-IJSSD 2450123

J.-X. Li et al.
by TONGJI UNIVERSITY on 10/22/23. Re-use and distribution is strictly not permitted, except for Open Access articles.

Fig. 19. Failure Probability of various conditions.


Int. J. Str. Stab. Dyn. Downloaded from www.worldscientific.com

where u and v represent the exceeding probability functions of PGA and wind speed
V , which are the same in Sec. 3.3. Then, the failure probability of the structure
without considering load correlation is calculated in the same way.
From Fig. 19, it can be seen that the failure probability of the towers increases
for every limit state when load correlation is considered. Especially for the ULS
state considering bolt slippage, the failure probability increases by 28%. This is
because the pressure wave generated during an earthquake may cause a significant
change in wind loads, which will affect the probability distribution of wind.7–9
If earthquakes and wind are considered independent variables and the influence
of their interaction is ignored, the probability of wind and earthquakes occurring
together will be underestimated. Therefore, the impact of load correlation on the
failure probability of wind turbine towers cannot be ignored.
In addition, Fig. 19 shows that the influence of the bolt slippage effect on the
failure probability of the structure is significant for the SLS limit state, increasing
the failure probability by 19.15 times. Therefore, ignoring the bolt slippage effect
could seriously underestimate the danger to wind turbine generators.

6. Conclusion
Based on the copula function model, the correlation between wind load and earth-
quake is considered. A failure probability prediction framework for wind turbine
towers under wind and seismic coupling is proposed. The failure probability of a
60 m high, 2.0 MW wind turbine tower is analyzed, and the main conclusions are
as follows:

(1) Under the joint action of wind and earthquakes, ignoring the correlation
between them will underestimate the failure probability of the structure. There-
fore, to ensure the safety of the structures, the correlation between loads should
be taken into account.

2450123-22
2nd Reading
September 27, 2023 16:57 WSPC/165-IJSSD 2450123

Failure Analysis of Lattice Wind Towers

(2) Under the joint action of wind and earthquake, compared with the wind loads,
the influence of seismic action on the fragility of wind turbine towers is more
significant. Therefore, more attention should be given to the seismic design of
wind turbine towers in earthquake-prone areas.
(3) The effect of bolt slippage on the dynamic response of wind turbine towers is
significant. Considering the bolt slip effect will reduce the stiffness of the tower
by TONGJI UNIVERSITY on 10/22/23. Re-use and distribution is strictly not permitted, except for Open Access articles.

and increase the tower top displacement under external loads. Under the same
load, bolt slippage will make wind turbine towers more prone to failure and
wind turbine generators more prone to breaking down.
(4) Ignoring the effect of bolt slip will lead to a biased unsafe analysis of wind
turbine towers under wind and seismic coupling action. Under the serviceability
limit state of wind turbine towers, the failure probability considering bolt slip
is much higher than that without considering bolt slippage.
Int. J. Str. Stab. Dyn. Downloaded from www.worldscientific.com

Acknowledgments
The Fundamental Research Funds for the Central Universities (Grant No.
N2301012), the Systematic Project of Guangxi Key Laboratory of Disaster Pre-
vention and Engineering Safety (Grant No. 2021ZDK017), Open fund of State Key
Laboratory of Coastal and Offshore Engineering, Dalian University of Technology
(Grant No. LP2218).

ORCID
Jia-Xiang Li https://orcid.org/0000-0003-2152-6960
Ling-Peng Wang https://orcid.org/0009-0003-9310-7682

References
1. C. C. W. Chang, T. J. Ding, T. J. Ping, K. C. Chao and M. A. S. Bhuiyan, Getting
more from the wind: Recent advancements and challenges in generators development
for wind turbines, Sustain. Energy Technol. Assessments 53 (2022) 102731.
2. M. V. Mendes, G. B. Colherinhas, M. V. Girão de Morais and L. J. Pedroso, Optimum
TLCD for mitigation of offshore wind turbine dynamic response considering soil–
structure interaction, Int. J. Struct. Stab. Dyn. (2023) 2350187.
3. H. Zuo, K. Bi, H. Hao, Y. Xin, J. Li and C. Li, Fragility analyses of offshore wind
turbines subjected to aerodynamic and sea wave loadings, Renew. Energy 160 (2020)
1269–1282.
4. A. Patil, S. Jung and O. S. Kwon, Structural performance of a parked wind turbine
tower subjected to strong ground motions, Eng. Struct. 120 (2016) 92–102.
5. Y. Sun, C. Xu, C. Cui, M. H. E. Naggar and X. Du, Seismic response of monopile-
supported OWT structure considering effect of long-term cyclic loading, Int. J. Struct.
Stab. Dyn. 23(09) (2023) 2350099.
6. K. Iqbal, C. Xu, Y. Han, Q. A. Motalleb, N. Ijaz and P. Dou, Numerical study on
seismic response of offshore wind turbine monopile in multi-layered soil profile of
Arabian Sea, Int. J. Struct. Stab. Dyn. 23(05) (2023) 2350049.

2450123-23
2nd Reading
September 27, 2023 16:57 WSPC/165-IJSSD 2450123

J.-X. Li et al.

7. S. Watada, T. Kunugi, K. Hirata, H. Sugioka, K. Nishida, S. Sekiguchi et al., Atmo-


spheric pressure change associated with the 2003 Tokachi-Oki earthquake, Geophys.
Res. Lett. 33(24) (2006) L24306.
8. S. Jin, L. Han and J. Cho, Lower atmospheric anomalies following the 2008 Wenchuan
Earthquake observed by GPS measurements, J. Atmos. Solar-Terrestrial Phys. 73 (7-
8) (2011) 810–814.
9. R. P. Singh, W. Mehdi and M. Sharma, Complementary nature of surface and atmo-
by TONGJI UNIVERSITY on 10/22/23. Re-use and distribution is strictly not permitted, except for Open Access articles.

spheric parameters associated with Haiti earthquake of 12 January 2010, Nat. Hazards
Earth Syst. Sci. 10(6) (2010) 1299–1305.
10. M. A. Asareh, W. Schonberg and J. Volz, Fragility analysis of a 5-MW NREL wind
turbine considering aero-elastic and seismic interaction using finite element method,
Fin. Elem. Anal. Des. 120 (2016) 57–67.
11. C. Yuan, J. Chen, J. Li and Q. Xu, Fragility analysis of large-scale wind turbines
under the combination of seismic and aerodynamic loads, Renew. Energy 113 (2017)
1122–1134.
Int. J. Str. Stab. Dyn. Downloaded from www.worldscientific.com

12. J. O. Martin del Campo and A. Pozos-Estrada, Multi-hazard fragility analysis for
a wind turbine support structure: An application to the Southwest of Mexico, Eng.
Struct. 209 (2020) 109929.
13. J. O. Martin del Campo, A. Pozos-Estrada and O. Pozos-Estrada, Development of
fragility curves of land-based wind turbines with tuned mass dampers under cyclone
and seismic loading, Wind Energy 24(7) (2021) 737–753.
14. M. A. Asareh, W. Schonberg and J. Volz, Effects of seismic and aerodynamic load
interaction on structural dynamic response of multi-megawatt utility scale horizontal
axis wind turbines, Renew. Energy 86 (2016) 49–58.
15. H. N. Li, G. Li, X. W. Zheng, X. Fu, Y. Liu, Z. Q. Dong et al., Advances in the study of
engineering structures under multi-hazard coupling effects, J. Civil Eng. 54(5) (2021)
1–14.
16. W. Bi, L. Tian, C. Li, Z. Ma and H. Pan, Wind-induced failure analysis of a transmis-
sion tower-line system with long-term measured data and orientation effect, Reliab.
Eng. Syst. Safety 229 (2023) 108875.
17. H. N. Li, X. W. Zheng and C. Li, Copula-based approach to construct a joint prob-
abilistic model of earthquakes and strong winds, Int. J. Struct. Stab. Dyn. 19(04)
(2019) 1950046.
18. Y. Xu, X. S. Tang, J. P. Wang and H. Kuo-Chen, Copula-based joint probability
function for PGA and CAV: A case study from Taiwan, Earthq. Eng. Struct. Dyn.
45(13) (2016) 2123–2136.
19. X. Fu, H. N. Li, G. Li, Z. Q. Dong and M. Zhao, Failure analysis of a transmission line
considering the joint probability distribution of wind speed and rain intensity, Eng.
Struct. 233 (2021) 111913.
20. X. W. Zheng, H. N. Li, Y. B. Yang, G. Li, L. S. Huo and Y. Liu, Damage risk
assessment of a high-rise building against multihazard of earthquake and strong wind
with recorded data, Eng. Struct. 200 (2019) 109697.
21. D. Nezamolmolki and A. Shooshtari, The effect of nonlinear behavior of bolted connec-
tions on dynamic analysis of steel transmission towers, Int. J. Steel Struct. 21 (2021)
634–649.
22. R. Balagopal, A. Ramaswamy, G. S. Palani and N. P. Rao, Simplified bolted connec-
tion model for analysis of transmission line towers, in Structures, Vol. 27 (Elsevier,
2020), pp. 2114–2125.

2450123-24
2nd Reading
September 27, 2023 16:57 WSPC/165-IJSSD 2450123

Failure Analysis of Lattice Wind Towers

23. Y. Zhan, B. Li, Z. Wu, L. S. Cunningham, G. Wu and Y. Yang, Modeling of galvanized


lattice steel structures incorporating the effect of joint slip, J. Construct. Steel Res.
173 (2020) 106252.
24. A. M. Taha, M. A. Dabaon, M. H. El-Boghdadi and M. F. Hassanein, Experimental
testing and evaluation of real-scale lap-splice bolted connections used in typical lattice
steel transmission towers, Thin-Walled Struct. 171 (2022) 108790.
25. R. R. de Souza, L. F. F. Miguel, G. McClure, F. Alminhana and J. Kaminski-Jr, Reli-
by TONGJI UNIVERSITY on 10/22/23. Re-use and distribution is strictly not permitted, except for Open Access articles.

ability assessment of existing transmission line towers considering mechanical model


uncertainties, Eng. Struct. 237 (2021) 112016.
26. S. Yaghoobi and A. Shooshtari, Joint slip investigation based on finite element mod-
elling verified by experimental results on wind turbine lattice towers, Front. Struct.
Civil Eng. 12 (2018) 341–351.
27. A. Ebrahimi, S. Yaghoobi and A. Shooshtari, Joint slip formulation for members with
double angle section based on experimental results in wind turbine lattice towers,
Arabian J. Sci. Eng. 47(10) (2022) 13699–13709.
Int. J. Str. Stab. Dyn. Downloaded from www.worldscientific.com

28. D. Nezamolmolki and A. Shooshtari, Investigation of nonlinear dynamic behavior of


lattice structure wind turbines, Renew. Energy 97 (2016) 33–46.
29. J. X. Li, J. P. Cheng, C. Zhang, C. X. Qu, X. H. Zhang and W. Q. Jiang, Seismic
response study of a steel lattice transmission tower considering the hysteresis charac-
teristics of bolt joint slippage, Eng. Struct. 281 (2023) 115754.
30. K. W. Campbell and Y. Bozorgnia, NGA-West2 Campbell-Bozorgnia ground motion
model for the horizontal components of PGA, PGV, and 5%-damped elastic pseudo-
acceleration response spectra for periods ranging from 0.01 to 10 sec, Pacific Earthq.
Eng. Res. Center (2013).
31. L. Tian and K. Liu, Uncertainty analysis of the dynamic responses of a transmis-
sion tower-line system subjected to cable rupture, J. Aerospace Eng. 34(1) (2021)
04020088.
32. E. A. Bossanyi, GH bladed theory manual, GH & Partners Ltd 2, 56–58. International
Electrotechnical Commission. (2005). Wind turbines-part 1: design requirements, IEC
61400-1-Ed. 3.0 (2003).
33. X. Fu, H. N. Li, G. Li and Z. Q. Dong, Fragility analysis of a transmission tower under
combined wind and rain loads, J. Wind Eng. Ind. Aerodyn. 199 (2020) 104098.
34. Applied Technology Council, Quantification of building seismic performance factors,
US Department of Homeland Security, FEMA (2009).

2450123-25

You might also like