You are on page 1of 13

Applied Ocean Research 132 (2023) 103483

Contents lists available at ScienceDirect

Applied Ocean Research


journal homepage: www.elsevier.com/locate/apor

Directional effects of correlated wind and waves on the dynamic response


of long-span sea-crossing bridges
Rugang Yang , Yongle Li , Cheng Xu , Yi Yang , Chen Fang *
Department of Bridge Engineering, Southwest Jiaotong University, Chengdu, China

A R T I C L E I N F O A B S T R A C T

Keywords: Sea-crossing bridges are sensitive to wind and wave loads, different wind and wave loading directions will lead to
Sea-crossing bridge significant differences in bridge dynamic performance. The design of such a structure may benefit from
Wind and waves considering the directional characteristics of loading. This study thus proposes a directional wind-wave-bridge
Directional effect
(DWWB) numerical framework by incorporating multivariate statistical dependence, directionality analysis,
Multivariate statistical dependence
Dynamic response
and a wind-wave-bridge system for sea-crossing bridges. The trivariate joint probability distribution is first
simulated by an optimized C-vine copula, and the correlated wind and wave combinations are identified by the
3D environmental contours based on the measured data in the bridge site. Directionality analysis involving the
alignment and misalignment of wind and waves are further investigated. The proposed approach is finally
applied to a real sea-crossing bridge with the mean wind speed, significant wave height, peak wave period, wind
direction, and wave direction as the input variables of DWWB. The nonlinear dynamic response as the output is
solved by equilibrium iterations using the Newmark-β method. The results show that the occurrence probability
of wind–wave misalignment is up to 50%. The dynamic response of the bridge is sensitive to the wind and wave
direction. Neglecting the multivariate statistical dependence and directional effect of wind and waves may lead
to a conservative and unrealistic result.

1. Introduction studied the directional effects of offshore wind turbines under wind and
wave loadings and stated that the extreme response is very sensitive to
The construction of sea-crossing bridges presents a growing trend to the structural orientation. Peng (2010) further studied the directional
meet the needs of transportation on the coast. Wind and wave loads will effect of offshore wind turbines under wind and wave loads and indi­
play an essential role in bridge design since they become the controlling cated that the direction of the wind and wave force affects the struc­
environmental loads, especially in deep water (Zhu et al., 2018). The ture’s failure mechanism and ultimate strength significantly. Philippe
direction of wind and wave is time-varying, and the variety of the et al. (2013) found that natural modes are excited differently regarding
incident angle of wind and wave loads will significantly affect the dy­ wave direction. The directionality analysis of wind and wave loads helps
namic performance of the bridge. A reasonable bridge orientation will to better understand the dynamic behavior of offshore structures and
effectively reduce the construction cost and the possibility of structural select the reasonable orientation.
failure induced by wind and waves. Considering the bridge generally contains thousands of degrees of
Many efforts in the directional effect of wind and wave loads on the freedom, previous studies have assumed that wind and waves are uni­
response, system reliability, fragility curves, natural frequencies, etc., of directional and loaded on the bridge perpendicular to the bridge axis
offshore structures have been carried out (Deepu et al., 2014; Li et al., (Meng et al., 2018; Zhu et al., 2018; Fang et al., 2020a). In principle, it is
2016; Jahangiri et al., 2021; Mo et al., 2021; Alkarem and Ozbahceci, possible that the most likely loading direction does not coincide with the
2021). These studies mostly focused on the directional effect of wind most unfavorable loading direction (Ti et al., 2017). Accurate estimation
(Zhang and Chen, 2015, 2017; Tian and Chen, 2020) or the directional of the dynamic response of the bridge by fully considering the direc­
effect of waves (Haver, 1991; Ji et al., 2015, 2022). Few studied the tional effect of wind and wave loads is essential in bridge design.
directional effects of both wind and wave loads. Wei et al. (2016) Therefore, a framework including the correlated wind and waves with

* Corresponding author.
E-mail addresses: 553656050@qq.com (R. Yang), lele@swjtu.edu.cn (Y. Li), 1127718636@qq.com (C. Xu), fangchenwave@163.com (C. Fang).

https://doi.org/10.1016/j.apor.2023.103483
Received 29 November 2022; Received in revised form 10 January 2023; Accepted 26 January 2023
Available online 1 February 2023
0141-1187/© 2023 Elsevier Ltd. All rights reserved.
R. Yang et al. Applied Ocean Research 132 (2023) 103483

directional dependences of dynamic response and loading is valuable for understand the bridge response excited by wind and waves. The x-axis
risk and safety assessment of sea-crossing bridges. refers to the direction along the bridge axis, and the y-axis refers to the
During the last two decades, three main approaches have been used direction perpendicular to the bridge axis. Since the terrain of the bridge
to determine directional environmental loads. One approach is to use site is always very complex and surrounded by islands and reefs, the
the modified coefficient method specified in the API code (American dominant directions of wind and waves are varied and uncertain. θwind
Petroleum Institute, 2005). As a simplified approach, the modified co­ represents the direction of wind load Fwind acting on the tower and girder
efficients are calculated by measured data in the Gulf of Mexico. How­ and θwave represents the direction of wave load Fwave acting the sub­
ever, the acquisition of the modified coefficient requires long-term merged structure. Noted that θb represents the orientation of the bridge
measurement and does not supply any probability information of the (i.e., the included angle between the bridge axis and due west direction).
most likely incident direction. A second approach is to generate a sto­
chastic wave based on a multidirectional wave spectrum (Haver, 1991), 2.2. DWWB framework
which is usually obtained by the spectral analysis of the time-domain
calculation results at a specific site. This process is always The Directional Wind–Wave-Bridge (DWWB) numerical framework
time-consuming, especially for nonlinear time-domain analysis. A third provides an efficient nonlinear dynamic approach to evaluate the bridge
approach borrows from both approaches and attempts to get the direc­ response under the correlated wind and waves. This framework involves
tional characteristics by directionality analysis based on observed or the interactions of the following three aspects: (a) wind-bridge aero­
simulated wave data. In this study, the third approach is used to obtain dynamic coupling; (b) wave-bridge hydrodynamic coupling; and (c)
the directional characteristics of wind and waves because of its high statistical dependence among wind and wave parameters.
efficiency and low cost.
The purpose of developing the DWWB numerical framework is to 2.2.1. Wind-bridge aerodynamic coupling
investigate the directional effect of wind and waves on nonlinear bridge The time-varying wind load is caused by natural wind, consisting of
response. It helps to better understand the dynamic behavior of sea- mean and fluctuating wind components. According to different wind
crossing bridges under wind-wave alignment/misalignment condi­ components, wind load is divided into static wind force induced by
tions. The DWWB numerical framework also effectively determines mean wind, buffeting force induced by fluctuating wind and self-excited
wind and waves’ most unfavorable direction for engineering design force induced by wind-bridge interaction. Each wind force is composed
purposes. Firstly, a numerical framework for a directional wind-wave- of drag, lift, and moment.
bridge system is proposed to assess structural response in a more accu­ Detailed expressions of wind loads on the bridge girder can be found
rate, load-response-dependent way. Secondly, the mean wind speed, the in (Zhu and Zhang, 2017). Without loss of generality, only the static
significant wave height, and the peak wave period (3D wind and wave wind drag FstD (θwind ), buffeting drag FbuD
(t, θwind ) and self-excited drag
parameters) are collected from a long-term measuring point. The opti­
Fse (t, θwind ) per unit length are given by:
D
mized C-vine copula and trivariate environmental contours are adopted
to determine the 3D correlated wind and wave parameters. The wind FstD (θwind ) = 0.5ρUd2 Cd (θwind )h (1)
direction and the wave direction are monitored from a temporary
[ ]
measuring point, and the possible incident directions with respect to the h v(t)
D
Fbu (t, θwind ) = 0.5ρUd2 b 2 Cd (θwind ) (2)
structure are statistically treated. Thirdly, the directionality analyses are b Ud
conducted to illustrate the characteristics of alignment and misalign­
ment of wind and waves. Finally, the dynamic responses of a real sea- ∫t
[
crossing bridge in wind-alone, wave-alone, and wind-wave alignment/
D
Fse (t, θwind ) =0.5ρUd2 fDp (t − τ, θwind )p(τ) + fDh (t − τ, θwind )h(τ)
(3)
misalignment conditions are calculated and discussed. − ∞
]
+ fDα (t − τ, θwind )α(τ) dτ
2. Numerical framework
where ρ is the air density; Ud is the mean wind speed at the height of
2.1. General configurations bridge girder; h and b are the height and the width of the girder;
Cd (θwind ) is the drag coefficient of the girder section in the wind direc­
Fig. 1 shows the general configuration of a sea-crossing cable-stayed tion θwind , v(t) is the lateral fluctuation wind speed; p(τ), h(τ), α(τ) are the
bridge composed of three basic components: cable, tower, and girder. It lateral, vertical, and torsional components of time-varying bridge de­
is well known that the bridge tower and pier are supported by large-scale formations, respectively; fDj(j=p,h,α) (θwind ) are impulse functions derived
cushion caps and dozens or even hundreds of piles to resist the impact of by flutter derivatives based on the rational approximation approach in
wind and waves, which is sensitive to the loading direction (Wei et al., the wind direction θwind .
2016). The bridge coordinate system is adopted here to better Regarding the wind loads on the bridge tower, the static wind forces
and the buffeting forces are considered and computed in the same way
as the girder. As the tower is a typical blunt structure, so the self-excited
forces acting on it are ignored. The wind loads on the cables are ignored
due to the following two points. First, the cable is a small-scale structure.
The total applied force is relatively small compared to the wind load
acting on the girder and bridge tower. Secondly, the drag coefficient of
the cable is small. Similar conclusions were found in (Bai et al., 2022;
Zhu et al., 2018).
After the bridge section is determined, the wind load depends on the
mean wind speed and fluctuating wind speed. The former is generally
obtained by observed data or historical meteorological data, while the
latter is often simulated by wind spectrum. The 3D fluctuating wind is
simplified to three one-dimensional wind fields along X, Y and Z-axis. In
terms of each wind field composed of multipoint wind speed, it is
simulated by the spectral representation method, and the corresponding
Fig. 1. Schematics of the sea-crossing bridge. statistical dependence is considered by the Davenport coherence func

2
R. Yang et al. Applied Ocean Research 132 (2023) 103483

tion (Li et al., 2004). The Kaimal spectrum is selected to generate the the cushion cap, the wave diffraction effect needs to be further studied.
fluctuating wind along the X-axis and Y-axis, and the Lumley-Panofsky The wave load acting on the grouped piles foundation could be calcu­
spectrum is taken as the target wind power spectra for vertical fluctu­ lated by the Morison equation. In Morison theory, wave load is treated as
ation along the Z-axis. Their power spectral density functions can be the sum of velocity-dependent drag force and acceleration-dependent
written as follows: inertia force. The horizontal wave force per unit length on a single
slender pile at t time and wave directions of θwave can be expressed as
nSv (f ) 15f
= (4)
u2∗ (1 + 9.5f )5/3 Fws (t, θwave ) = 0.5ρw CD (θwave )Dp (uw (t) − ub (t))|(uw (t) − ub (t))|

nSu (f ) 200f +ρw Au̇w (t) + (CM (θwave ) − 1)ρw A(u̇w (t) − u̇b (t)) (11)
= (5)
u2∗ (1 + 50f )5/3
where, ρw is the water density; CD (θwave ), CM (θwave ) are the drag coef­
nSw (f ) 3.36f ficient and the inertia coefficient in the wave direction θwave ; Dp and A
= (6) are the diameter and cross-sectional area; uw , u̇w are the velocity and
u2∗ (1 + 10f )5/3
acceleration of water particle; ub , u̇b are the velocity and acceleration of
where f = nz/U(z) is the dimensionless frequency, n is frequency and the pile.
U(z) is mean wind speed at height z; u∗ = KU(z)/ln(z /z0 ) is the friction Due to mutual shielding and blocking among piles, the wave load
velocity, K = 0.4 is von Kármán constant and z0 = 0.005 is the sea acting on the isolated pile is significantly different from that on a single
surface roughness for a rough sea surface. pile surrounded by the group piles. Therefore, the interference coeffi­
cient and the shelter coefficient are introduced to estimate the pile group
2.2.2. Wave-bridge hydrodynamic coupling effect (Bonakdar et al., 2015). The shelter coefficient is usually conser­
Because of the standard wave-energy spectrum of irregular wave vatively taken as 1.0, and the interference coefficient depends on the
trains, the desirable irregular wave field is simulated through a linear arrangement and size of the group piles. Make use of the pile group
superposition of a finite number of regular waves with varying ampli­ coefficient obtained by laboratory tests, the total horizontal wave force
tudes, phases, and periods. The wave surface elevation as a function of y per unit length for the group piles at t time and wave directions of
and time t is given by: θwave can be written as:
n ∑
∑ m
∑ (12)
M
Fwb (t, θwave ) = KGjk Fws
jk
(t, θwave )
φ(y, t) = ai cos[wi t − ki y + εi ] (7)
j=1 k=1
i=1

where wi , ki are the frequency and the number of the ith wave compo­ where n and m are the numbers of columns and rows of piles, which are
respectively parallel to and perpendicular to the dominant wave direc­
nent, satisfying the linear dispersion relationship such that w2i =
jk
ki gtanh(ki d); εi is the random initial phase of the ith wave components; tion; Fws is the horizontal wave force of the j,kth pile; the pile group
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ jk
ai = 2S(wi )Δw is the amplitude of the ith wave components; S(wi )are coefficient KG of the j,kth pile is set to 1.0 for the side piles and
the wave spectrum; Δw = [iΔw + (i − 1)Δw]/2, where Δw 1.265 − 0.225ln(a /Dp ) for the middle piles, of which a is the distance
= (wmax − wmin )/M is the frequency resolution, dividing the wave spec­ between two adjacent piles.
trum into M unequal subranges.
The horizontal water velocity u̇w and acceleration üw are expressed 2.2.3. Statistical dependence among wind and wave parameter
by Chakrabarti (1971) as follows: Wind and wave loads depend on several key environmental param­
eters such as wind speed, wave height, wave period, wind direction, and
H coshki z ( )
u̇w (z, t) = w ( ( )) cos ki yj − wt (8) wave direction. Some studies showed strong dependence among wind
2 sinh ki d + φ yj , t
and wave parameters (Bai et al., 2020; Fang et al., 2020b) and high­
lighted that uncorrelated wind and wave assumption is conservative and
H 2 coshki z ( )
üw (z, t) = w ( ( )) sin ki yj − wt (9) unrealistic in practical engineering. Therefore, the multivariate joint
2 sinh ki d + φ yj , t
probability distribution is introduced to simulate the dependence among
wind and wave parameters.
where H is the wave height; yj is the horizontal length from the initial
As a practical analytical approach of the high-dimensional joint
position to the jth pile foundation of the bridge along the wave travel probability distribution, the C-vine copula is used to decompose high-
direction; z is the height from the calculation point to the pile bottom; w dimensional variables from a high-dimensional copula into several
= 2π/T is the wave frequency; d is the still water level; φ(yj , t) is the two-dimensional copulas. In this paper, the dependences of three key
wave surface elevation, which can be generated from the random wave variables involving wind speed, wave height, and wave period are
spectrum by using the wave superposition method. The JONSWAP investigated. Let x1 , x2 and x3 represent wind speed, wave height, and
spectrum is used in this study as the random wave spectrum expressed wave period, respectively, their joint probability density function f(x1 ,
by x2 , x3 ) (JPDF) can be rewritten in terms of conditional probability
[ ( )2 / ] density functions (Montes-Iturrizaga and Heredia-Zavoni, 2015; 2016):
( ) [ ( ) ] exp − w − 1 2σ 2
1 w − 5 5 w − 4 wp
(13)
S(w) = βj Hs2 Tp− 4 exp − Tp ⋅γ (10) f (x1 , x2 , x3 ) = f (x1 )⋅f2|1 (x2 |x1 )⋅f3|1,2 (x3 |x1 , x2 )
2π 2π 4 2π
Where f2|1 (x2 |x1 ) can be decomposed into a pair-copula c12 (u1 , u2 ) and a
0.06238
where βj = 0.230+0.0336γ− 0.185(1.9+γ)− 1 ⋅[1.094 − 0.01915 lnγ]; Tp = marginal probability density function f(x2 ); f3|1,2 (x3 |x1 , x2 ) can be
Ts
0.559 ; Hs ,Ts are the significant wave height and the associated decomposed into a pair-copula c13 (u1 , u3 ), a conditional pair-copula
1− 0.132(γ+0.2)−
wave period, respectively; Tp , wp are the peak wave period and the c23|1 (F2|1 (x2 |x1 ), F3|1 (x3 |x1 )) and a marginal probability density func­
corresponding circular frequency, respectively; σ = 0.07 when w ≤ wp , tion f(x3 ); ui = Fi (xi ) is the marginal cumulative distribution function
or σ = 0.09 when w > wp ; and γ is the peak enhancement factor of 3.3. (JCDF) of xi .
As a result, Eq. (13) can be expressed as:
Because of the high elevation of the sea-crossing bridges, the waves
usually could not impact the bridge deck. When wave height runs over

3
R. Yang et al. Applied Ocean Research 132 (2023) 103483

( )
f (x1 ,x2 ,x3 ) = f (x1 )⋅ f (x2 )c12 (u1 ,u2 )⋅f (x3 )c13 (u1 ,u3 )c23|1 F2|1 (x2 |x1 ),F3|1 (x3 |x1 )
Mb üb (t) + Cb u̇b (t) + K b ub (t) = Fsb (t, θwind ) + Fbub (t, θwind ) + Fseb (t, θwind )
(14)
+ Fwb (t, θwave )
Where c12 and c13 are the copula density functions for the variables (u1 , (19)
u2 ) and (u1 , u3 ), respectively; c23|1 is the conditional copula density,
which consists of the marginal conditional cumulative distribution where, Mb , Cb , K b are the mass, damping and stiffness matrixes of the
functions of F2|1 (x2 |x1 )and F3|1 (x3 |x1 ). bridge, respectively; üb (t), u̇b (t), ub (t) are the acceleration, velocity and
⃒ displacement vectors of the bridge node at t time, respectively;
The Fi|j (xi ⃒xj ) in Eq. (14) is obtained by the following recursive
Fsb (t, θwind ), Fbub (t, θwind ), Fseb (t, θwind ) are the static wind forces, the
formula.
buffeting forces and the self-excited forces at t time and a load direction
( ( )) ( )
( ⃒ ) ∂Cij Fi (xi ), Fj xj ∂Cij ui , uj of θwind , respectively; and Fwb (t, θwave ) is the wave forces acting on the
Fi|j xi ⃒xj = ( ) = (15) group piles at t time and a load direction of θwave .
∂Fj xj ∂uj
The nonlinear dynamic response induced by wind and wave loads
When xi , xj are uniform random variables, the following function is could be obtained using this governing equation. The general flowchart
defined: is shown in Fig. 2. Both large deformation and sag effects are considered.
( ⃒ ) ( ⃒ ) ( ) The following procedure is performed to complete the DWWB analysis.
Fi|j xi ⃒xj = Ci|j ui ⃒uj = h ui , uj , θ12 (16)
(a) Determine the correlated wind and wave parameters of wind
Where xj is the conditioning variable corresponding to the variable xi ; speed, wave height, and wave period based on an optimized C-
θ12 is the parameter of the copula of bivariate (u1 , u2 ); the h-function vine copula model (Fang et al., 2022);
h(ui , uj , θ12 ) is a simplified form of the marginal conditional cumulative (b) Establish the full bridge finite element model and simulate the
distribution function. fluctuating wind field and the random wave field;
For three random variables, the conditional probability distributions (c) Calculate the time-varying aero- and hydrodynamic forces in
are given by: ANSYS based on the wind and wave elevation time histories
(
∂c23|1 F2|1 (x2 |x1 ), F3|1 (x3 |x1 )
) simulated by Matlab as excitation sources acting on the bridge;
F3|12 (x3 |x1 , x2 ) = (17) (d) Perform a wind-wave-bridge analysis to obtain the solution
∂F2|1 (x2 |x1 )
through equilibrium iterations. Since the wave force and self-
Then, replacing Eqs. (15) and (16) in Eq. (17): excited force contain bridge motion items and are motion
dependent, the iterative method is used to reach the convergence
∂c23|1 [h(u1 , u2 , θ12 ), h(u1 , u3 , θ13 )]
F3|12 (x3 |x1 , x2 ) = (18) condition at each time step.
∂h(u1 , u2 , θ12 )
(e) Set θwind = θwind,i i ∈ [1, nθ ] and θwave = θwave,j j ∈ [1, nθ ] based on
Some commonly used JCDF and h-function in Eq. (18) are summa­ the directionality analysis and repeat from step (c), incrementing
rized in Table 1. The Gumbel, Clayton, and Frank copulas belong to the the value of i and j so that all desired load directions are
class of the Archimedean copulas, which belong to one-parameter considered.
functions, whereas the Gaussian copula is the Elliptical copulas whose
dependence is fitted using a symmetric and positive definite matrix. In In the present study, the linear dynamic system has been conducted
terms of tail dependences for the four pair-copulas, the Clayton copula for the preliminary design of structures where it provides flexibility to
and Gumbel copula have more probability concentrated in lower-tail evaluate the bridge response at a low computational cost. Nevertheless,
and upper-tail, respectively. The Frank copula models both lower-tail the full nonlinearity including material nonlinearity of the bridge under
and upper-tail simultaneously. The Gaussian copula is neither lower- the extreme wind and wave loads needs further investigation.
tail nor upper-tail dependent, which is also approximated as the Nataf
transformation. These four pair-copulas have been widely used in wind 3. Correlated wind and wave parameters
and ocean engineering.
3.1. Measured data at the bridge site
2.3. Governing equation of motion for the DWWB framework
Fig. 3 shows a sea-crossing bridge located on the southeast coast of
China. As an important channel connecting the mainland and islands, it
With all the pair interactions between wind, wave, and bridge, the
is susceptible to joint actions of wind and waves. It can be clearly
governing equation of the coupled wind-wave-bridge system is given by:

Table 1
Four commonly used JCDF and h-function.
Copulas c12 (u1 , u2 ) h-function
⎛ ⎞
Gaussian ϕ(x1 , x2 |θ12 ) xi = ϕ − 1
(ui ) − 1
⎜ϕ (u1 ) − θ12 ϕ− 1 (u2 ) ⎟
ϕ⎝ √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ ⎠
1 − θ212
Clayton (u−1 θ12 + u−2 θ12 − 1)
− 1/θ12 − 1− 1/θ12
u−2 θ12 − 1 (u−1 θ12 + u−2 θ12 − 1)
Gumbel exp( − ((− lnu1 ) θ12
+ (− lnu2 ) θ12 1/θ12 1
) ) c12 (u1 , u2 ) (− logu2 )θ12 − 1
u2
1
− 1
× {(− logu1 )θ12 + (− logu2 )θ12 }θ12
[ ]
Frank 1 (e− θ12 u1
− 1)(e − θ12 u2
− 1) e− θ12 u2
− ln 1 +
θ12 (e− θ12 − 1) 1 − e− θ12
+ e− θ12 u2 − 1
1 − e− θ12 u1

4
R. Yang et al. Applied Ocean Research 132 (2023) 103483

Fig. 2. Simulation procedure for the DWWB numerical framework.

Fig. 3. Geographical location of the sea-crossing bridge and corresponding measuring points.

5
R. Yang et al. Applied Ocean Research 132 (2023) 103483

observed from Fig. 3 that more than 20 islands surround the bridge site decomposed into two bivariate copula density functions and the corre­
and reefs, and the environmental characteristics of wind and waves in sponding conditional copula density. In order to obtain a high-precision
this region are different from that in the open sea. In order to more copula model, the C-vine copula is used to estimate the 3D JPDF of mean
accurately estimate the statistical dependence between wind and wave wind speed U10, significant wave height Hs and peak wave period Tp.
in the bridge site, the measured data monitored from two nearby Firstly, the R Largest Order Statistics (R-LOS) method (An and Pandey,
measuring points are adopted. 2007) is used to select the samples of extreme values with an R of 6, that
Point 1 is a long-term measuring point with 5 km offshore and 30.5 m is, 6 extreme events are considered per year. According to independent
water depth. Ten years of data from 2007 to 2016, including mean wind and identically distributed requirements, nine typhoon events are
speed, significant wave height, and peak wave period, are collected from removed from the wind data. The six maximum values of the mean wind
measuring point 1. The mean wind speed U10 refers to the maximum 10- speed and the corresponding other two variables are selected from
min mean wind speed within one hour at 10 m height above sea level. one-year data. The selected six maximum values of (U10, Hs, Tp) in each
The significant wave height Hs is the maximum 30-min significant wave year should be spaced more than 72 h apart to make sure that they can
height within one hour. The peak wave period Tp is the period of the be assumed as independent events (Valamanesh et al., 2015). Secondly,
wave corresponding to the greatest power spectral density within one a total of 60 groups of (U10, Hs, Tp) are selected as the sample of extreme
hour. Point 2 is a temporary measuring point near the bridge axis. Only values. Since all data is recorded hourly, only a 1-hr database is avail­
wind direction and wave direction data in 2018 are available. Wind able. Thirdly, the one-step optimization method is used to obtain the
direction and wave direction refer to the incoming direction of wind and optimal marginal distributions and the optimal pair-copulas, in which
waves in the X-Y plane. Since the recording time of the two measuring the maximum log-likelihood (MLogL) values and the root mean square
points does not coincide, the directions of wind and waves do not strictly error (RMSE) are used to evaluate the fitting accuracy of the marginal
correspond to the selected wind speed and wave height. Instead, some of distributions and pair-copulas, respectively.
the most likely directions are discussed and considered in the following Fig. 4 shows the JPDF of optimal bivariate and trivariate copula. The
analysis. Nevertheless, the large values of wind and wave parameters marginal distributions of mean wind speed, significant wave height, and
and corresponding directions need to be selected simultaneously for peak wave period are estimated by the four commonly-used marginal
more accurate and reasonable statistics when conditions permit. density functions, which include Weibull, Lognormal, Gamma, and
General Extreme Value (GEV) functions. It can be observed from the left
side of Fig. 4 that the GEV distribution is most suitable for the variables
3.2. 3D statistical dependence of wind speed, wave height, and wave U10 and Tp while the lognormal distribution is the best for the variable
period H s.
Furthermore, the Gumbel copula and the Frank copula are used to
3.2.1. Optimized C-vine copula simulate the bivariate copula of (U10–Hs) and (U10-Tp) according to the
As mentioned above, the trivariate copula density functions could be

Fig. 4. JPDF contour slices of the optimized C-vine copula and the corresponding marginal distributions.

6
R. Yang et al. Applied Ocean Research 132 (2023) 103483

Fig. 5. 3D environmental contour of (U10–Hs-Tp) and the corresponding 2D contours at 50-year return period.

criterion of minimum RMSE value between the empirical and theoretical combination values of (U10, Hs, Tp) are 24.29 m/s, 4.91 m, and 8.11 s
JCDF values. Finally, the trivariate joint probability distribution can be when the U is taken as the maximum value. Obviously, the accompa­
solved by Eq. (9). As shown in the right side of Fig. 4, the conditional nying Hs and Tp will not reach the design value (6.35 m and 9.70 s) of the
JPDF of (Tp-Hs) for U10 is divided into 15 contour slices from 6 m/s to 20 50-year return period. Thus, multivariate statistical dependence be­
m/s at an interval of 1 m/s. The scatter represents the empirical results tween wind and wave parameters needs to be considered to avoid overly
at the measuring point. The clustering characteristics of the trivariate conservative results.
copula match the empirical results very well.
3.3. Wind and wave directions based on measured data
3.2.2. Trivariate environmental contours
Once the trivariate joint probability distribution is obtained, the 3D Historical measurements of offshore conditions can be used to
environmental contours could be obtained by the Inverse First Order investigate the directional distribution of wind and waves, which is
Reliability Method (IFORM). With a given return period, a correlated utilized as input for the DWWB analysis. The directional characteristics
combination of (U10, Hs, Tp) can be determined based on the 3D envi­ of wind and waves are extracted from the measured data at point 2 and
ronmental contours. In consideration of the limit of only ten years of the occurrence probability is calculated by a direct frequency-based
measurement data, the return periods are set as 20 years, 30 years, and estimation. Fig. 6(a) shows the wind direction distribution and the
50 years respectively. Without loss of generality, the 3D environmental corresponding occurrence probability. Possible wind directions are
contour of (U10–Hs-Tp) with 50 years return period is taken as an concentrated on the range from 55◦ - 120◦ and 270◦ - 300◦ with the
example to illustrate its combination characteristics as shown in Fig. 5 occurrence probability from 1% - 24%. The wind direction from 75◦ - 90◦
(a). The isosurface of (U10–Hs-Tp) presents an ellipsoid covering all is a range with the highest occurrence probability of 24%. The second
possible combinations of wind and waves. The environmental contours highest occurrence probability of 17% and the third highest occurrence
of Hs and Tp with a given value of U10 ranging from 5 m/s to 20 m/s at an probability of 11% occur at the range of wind direction from 65◦ - 75◦
interval of 5 m/s are shown in Fig. 5(b). A higher wave height is more and 270◦ - 280◦ , respectively. Similarly, the wave direction distribution
likely to co-occur with a higher wave period. Fig. 5(c) shows the envi­ and the corresponding occurrence probability are presented in Fig. 6(b).
ronmental contours of U10 and Hs for a given value of Tp ranging from 3 s It is also interesting that the wave directions are concentrated in the
to 11 s at an interval of 2 s. Similarly, a significant positive correlation range from 30◦ - 150◦ The highest occurrence probability of 25% occurs
between U10 and Hs can be found. It is worth noting that the maximum at the range of wave direction from 75◦ - 85◦ , which is almost coincident
values of the three variables do not coincide. For instance, the with the most likely range of wind direction.

Fig. 6. Directionality analysis of wind and waves based on measured data: (a) wind direction; (b) wave direction.

7
R. Yang et al. Applied Ocean Research 132 (2023) 103483

mass elements. All piles and piers are fixed at the bottom. The self-
vibration characteristics of the bridge are first analyzed to master its
natural frequency and corresponding vibration mode. The natural fre­
quencies of the first-order lateral bending, vertical bending, and
torsional modes of the bridge are 0.211 Hz, 0.278 Hz, and 0.435 Hz,
which are used to solve the coefficient of Rayleigh damping. It is worth
noting that the fundamental frequency of lateral bending mode is
smaller than the fundamental frequency of vertical bending mode. This
indicates that the bridge is more sensitive to lateral wind and wave
loads.

4.2. Environmental conditions

In order to assess the directional effects of 3D correlated wind and


waves on the dynamic response of the sea-crossing bridge, the correlated
mean wind speed U10, significant wave height Hs and peak wave period
Fig. 7. Included angle between wind direction and accompanying
Tp are extracted through 3D environmental contours. The maximum
wave direction.
value of U10 and the corresponding values of Hs and Tp for three different
return periods are used as the input values of the DWWB framework. In
It is important to analyze the characteristics of alignment and
addition, the wind-alone and wave-alone cases are also calculated for
misalignment between wind direction and wave direction. Fig. 7 shows
comparison. All wind and wave parameters of (U10–Hs-Tp) are listed in
the included angle between wind direction and accompanying wave
Table 2.
direction. Both directions are recorded simultaneously. As observed, the
The wind and wave parameters of (U10–Hs-Tp) have here been
highest occurrence probability of 50.4% occurs at the range of included
assumed independent of the direction in the absence of data strongly
angle from 0◦ - 6◦ , indicating the wind and wave are almost in the same
indicating otherwise. According to the directional distribution of wind
direction during half of the statistical time. Nevertheless, the probability
and waves, several directions with high occurrence probability are
of wind–wave misalignment is nearly 50%. The second highest occur­
selected as input values for the DWWB numerical framework. Since the
rence probability of 11% occurs at the range of included angle from
bridge is an axisymmetric structure, the direction ranges are set as 0◦ to
150◦ - 156◦ Neglecting the wind–wave misalignment will lead to an
90◦ for wind-wave, wind-alone, and wave-alone cases. The range of θ
unrealistic result.
has been discretized into a series of 15◦ intervals, as summarized in
Table 3.
4. Example bridge and environmental conditions

5. Results and discussion


4.1. Load and structure modeling

This section describes the directional effects of sea-crossing bridges


The sea-crossing bridge investigated here is a cable-stayed bridge
under correlated wind and waves. Once the wind and wave parameters
with a span arrangement of (90+145+560+145+90) m. As shown in
are determined, the DWWB analysis is performed to solve the nonlinear
Fig. 8, the bridge is composed of a girder, two towers, four auxiliary
dynamic response of the selected bridge. As mentioned, the mean wind
piers, and a large number of cables. The prestressed concrete box-girder
speed, the significant wave height, the peak wave period, the wind di­
with a size of 22 m × 6 m is divided into 116 segments, and the wind
rection, and the wave direction are taken as the input variables. The
loads are applied on the 117 nodes marked from G1 to G117. The con­
output data contain the displacements of all nodes of the bridge. Spe­
crete tower with a height of 161 m consists of two columns above the
cifically, the Newmark-β method is used to solve the equation of motion
still water level (SWL) and a submerged foundation supported by 38
piles with a radius of 3.4 m. The wind loads acting tower column are
marked from T1 to T21, and the wave loads acting group piles are Table 2
Wind and wave parameters for wind-wave, wind-alone, and wave-alone cases.
marked from P1 to P5 (5 rows). Since the SWL is lower than the cushion
cap, the wave loads are only applied to the piles. All auxiliary piers are Type Return Period U10/(m/s) Hs/m Tp/s
built on artificial reefs above the SWL. Wind-Wave 20 20.51 4.50 7.85
The finite element model of the bridge is established by ANSYS 30 22.33 4.69 7.97
software. The girder, the towers, and the piers are simulated by 3D beam 50 24.29 4.91 8.11
Wind alone 50 24.29
elements, whereas the cables are built by only-tension link elements.
Wave alone 50 4.91 8.11
The secondary loads acting on the bridge nodes are modeled by adding

Fig. 8. Loading diagram of sea-crossing bridges (unit: cm).

8
R. Yang et al. Applied Ocean Research 132 (2023) 103483

Table 3 5.1. Response evaluation in wind or wave alone


Wind and wave directions for wind-wave, wind-alone, and wave-alone cases.
Type θwind /◦ θwave /◦ The directional effect of bridges under wind alone or wave alone is
first investigated to better understand the contribution of wind or wave
Wind alone 60,75,90 ——
Wave alone —— 30,45,60,75,90 to bridge response. Fig. 9 shows the time-history dynamic analysis of the
Wind-wave alignment 60,75,90 sea-crossing bridge under wind alone for a 50-year return period. All
Wind-wave misalignment 90 30,45,60,75 results are obtained from DWWB numerical framework. As observed in
30,45,60,75 90 Fig. 9(a), the periodic buffeting response in the middle of the main span
(mid-span) obtained from three different wind directions can be seen on
and reach the prescribed convergence. According to the preliminary the left side of the figure, while the corresponding root mean square
sensitivity analyses, the time step of 0.05 s is adopted in the DWWB (RMS) value is shown in the right side of the figure. The displacement
analysis to provide accurate response results with reasonable time response gradually increases with the wind direction from 60 ◦ to 90 ◦
consumption. The dynamic responses with a duration of 250 s for just The RMS value reaches 0.012 m at the wind direction of 90 ◦ The lon­
one numerical simulation need to be calculated for 18.5 h through one gitudinal (X-axis) displacement Dtx and lateral (Y-axis) displacement Dty
workstation, which is equipped with an Intel Xeon E5–2650v5 3.40 GHz at top tower are presented in Figs. 9(b) and 9(c). The Dty increases slowly
CPU and 32 GB memory size. with the wind direction from 60 ◦ to 90 ◦ , while the Dtx drops rapidly to
0 with the wind direction from 60 ◦ to 90 ◦ The tower vibrates signifi­
cantly in both directions under the yawed wind. Figs. 9(d) and (e)

Fig. 9. Dynamic response of the sea-crossing bridge under wind alone of 50-year return period: (a) lateral displacement at mid-span; (b) longitudinal displacement at
top tower; (c) lateral displacement at top tower; (d) base shear; and (e) base bending moment.

9
R. Yang et al. Applied Ocean Research 132 (2023) 103483

Fig. 9. (continued).

Fig. 10. RMS displacements of the sea-crossing bridge under wave alone of 50-year return period: (a) lateral displacement at mid-span; (b) longitudinal displacement
at top tower; (c) lateral displacement at top tower.

present the base shear along the Y-axis and the base bending moment the girder. In addition, the maximum RMS value of the bridge occurs at
around the Z-axis. The Fy increases with the wind direction from 60 ◦ to both wind and wave directions 90 ◦ , indicating that the most unfavor­
90 ◦ , while the MZ remains almost unchanged. able loading direction of the bridge is 90 ◦ , that is, the inflow loads are
Fig. 10 shows the directionality analysis of the sea-crossing bridge loaded perpendicular to the bridge axis system.
under wave alone of a 50-year return period. The trend of girder and
tower responses with wave direction is the same as that with wind di­ 5.2. Response evaluation in wind-wave alignment
rection. For the lateral displacement at mid-span as shown in Fig. 10(a),
the wave-induced response with the RMS value of 0.0022 m is equiva­ The dynamic responses of the girder and tower with the wind and
lent to 18% of wind-induced response with the RMS value of 0.0119 m wave directions from 45 ◦ to 90 ◦ are present in which wind-wave
when the wind and wave direction is taken as 90 ◦ While the lateral and misalignment is assumed to be zero. Fig. 11 shows the RMS values of
vertical wave-induced responses account for 52.3% and 47.6% of the the lateral displacement response of all girder and tower nodes in the 50-
lateral and vertical wind-induced responses at the direction of 90 ◦ Since year return period. The maximum response of the girder occurs in the
the wave is directly loaded on the group piles of the bridge tower, the middle of the span, and the lateral responses gradually decrease from the
influence of waves on the bridge tower is more significant than that on mid-span to both side spans, as shown in Fig. 11(a). By comparing four

10
R. Yang et al. Applied Ocean Research 132 (2023) 103483

Fig. 11. Displacement responses of the bridge for four different directions in the 50-year return period: (a) lateral RMS value along the girder; (b) lateral RMS value
along the tower.

Fig. 12. RMS response of the bridge under wind-wave alignment for different return periods: (a) lateral displacement at mid-span; (b) lateral displacement at top
tower; (c) base shear force.

different wind and wave directions, the maximum response occurs at found for four directions. The tower and base responses show a similar
θwind = θwave = 90 ◦ The mid-span response of 0.013 m is lower than the trend in Figs. 12(b) and 12 (c). The RMS value of top tower of correlated
simple superposition of the wind-induced response of 0.012 m and the cases increased by 17% to 20% compared with that of uncorrelated case.
wave-induced response of 0.002 m in the 50-year return period. This The RMS value of the base shear force of correlated cases increased by
shows that the joint action of wind and wave loads causes a slight 20% to 23% compared with that of uncorrelated case. Obviously, the
reduction of the mid-span response. As observed in Fig. 11(b), the assumption of independence is conservative and unrealistic under the
maximum responses calculated from four wind and wave directions actual bridge site. Moreover, the wind-induced shear accounts for 48%
occur at θwind = θwave = 90 ◦ When the wind and wave direction is taken of the total shear at the direction of 90◦ , indicating that both wind and
as 90 ◦ , the maximum response of 0.006 m at top tower is also slightly wave have significant contributions to the internal force of the base.
lower than the sum of the wind-induced response of 0.005 m and the
wave-induced response of 0.002 m. The simple superposition of the
5.3. Response evaluation in wind-wave misalignment
wind-induced and wave-induced responses will overestimate the real
responses of the bridge.
Since the 90◦ directions is the most unfavorable wind and wave
Fig. 12 shows the RMS values of lateral displacement and internal
loading direction, the wind-wave misalignment analysis is set to the
efforts responses for three correlated cases and one uncorrelated case.
following two conditions. One is that the wind direction is fixed at 90◦
Three correlated cases marked by a solid dot refer to the DWWB analysis
and the wave direction changes from 30◦ to 75◦ at 15◦ intervals. The
of trivariate wind and wave parameters in three return periods, while
other is that the wave direction is fixed at 90◦ and the wind direction
the uncorrelated case marked by a hollow dot refers to the DWWB
changes from 30◦ to 75◦ at 15◦ intervals. Fig. 13 shows the RMS response
analysis of univariate wind and wave parameters in the 50-year return
of the girder and the tower in four different directions in the 50-year
period. It is observed from Fig. 12(a) that the RMS value increases with
return period. When the wind direction is fixed, the variety of girder
the increase of the return period. Compared with correlated cases in the
displacement can be almost neglected as shown in Fig. 13(a). Only 5.6%
50-year return period, a 17% to 21% increase in mid-span response is
differences are observed by comparing the mid-span response at θwave =

11
R. Yang et al. Applied Ocean Research 132 (2023) 103483

Fig. 13. Displacement response of the bridge for four different directions in the 50-year return period: (a) lateral RMS value along the girder fixed the wind direction;
(b) lateral RMS value along the tower fixed the wind direction; (c) lateral RMS value along the girder fixed the wave direction; (d) lateral RMS value along the tower
fixed the wave direction.

30 ◦ and θwave = 75 ◦ However, the wind direction significantly changes cable-stayed bridge has been performed, and the directional effects on
the response of the girder as shown in Fig. 13(c). The response of the the dynamic response of the bridge have been discussed. The main
girder is reduced by nearly five times from θwind = 75 ◦ and θwind = 30 ◦ conclusions from this study can be summarized as follows:
This implies that the wind controls the displacement of the girder. Un­
like the girder, the bridge tower is significantly affected by wind and (a) The significant statistical dependences among wind speed, wave
wave direction. When the wind direction is fixed, the lateral response of height, and wave period are found based on measured data in the
top tower at θwave = 75 ◦ is reduced by 38.7% than that at θwave = 30 ◦ bridge site area. The optimized C-vine copula provides a feasible
When the wave direction is fixed, the lateral response of top tower at way to simulate the 3D statistical dependence of wind and wave
θwind = 75 ◦ is reduced by 69.2% than that at θwind = 30 ◦ It can be seen parameters and identify the possible wind and wave combina­
from Figs. 13(b) and 13(d) that the tower response is dominated by both tions in different return periods.
wind and wave. (b) Compared with bridge responses under the combined action of
wind and waves in alignment, the simple superposition of wind-
6. Conclusion induced and wave-induced responses will overestimate the
bridge response. Wind and wave loads need to be considered
A numerical framework for a directional wind-wave-bridge system is simultaneously for the sea-crossing bridge.
proposed to evaluate the dynamic performance of a sea-crossing bridge. (c) By comparing only wind load, only wave load, and wind-wave
The directional distribution and statistical dependence of wind and load induced responses, the 90◦ directions is the most unfavor­
waves are investigated according to the measured data in the bridge site able wind and wave loading direction. Adjusting the orientation
area for more accurate and realistic purposes. The 3D correlated wind of the bridge, that is, avoiding the direction of the bridge axis
speed, wave height, and wave period are estimated by the optimized C- system perpendicular to the most likely direction of wind and
vine copula, and the most likely wind and wave combinations are waves, can effectively reduce the construction cost and risk
defined by the trivariate environmental contours. The characteristics of possibility.
alignment and misalignment between wind direction and wave direc­ (d) The occurrence probability of wind–wave misalignment is up to
tion are further investigated. A case study using a real sea-crossing 50% according to the directional statistics. The evaluation results

12
R. Yang et al. Applied Ocean Research 132 (2023) 103483

in wind-wave misalignment indicate that wind load dominates Bonakdar, L., Oumeraci, H., Etemad-Shahidi, A., 2015. Wave load formulae for
prediction of wave-induced forces on a slender pile within pile groups. Coast. Eng.
the girder response, and the tower response is controlled by both
102, 49–68.
wind and wave loads. Chakrabarti, S.K., 1971. Discussion of nondeterministic analysis of offshore structures.
(e) The responses of the girder and tower in different directions in­ J. Eng. Mech. Div. 97 (3), 1028–1029.
crease by 17% to 21% for the correlated cases compared with Deepu, S.P., Prajapat, K., Ray-Chaudhuri, S., 2014. Seismic vulnerability of skew bridges
under bi-directional ground motions. Eng. Struct 71, 150–160.
those for the uncorrelated cases. The bridge response will thus be Fang, C., Tang, H., Li, Y., Wang, Z., 2020a. Effects of random winds and waves on a long-
overestimated if the statistical dependence between wind and span cross-sea bridge using Bayesian regularized back propagation neural network.
waves is not considered. Adv Struct Eng 23, 733–748.
Fang, C., Tang, H., Li, Y., Zhang, J., 2020b. Stochastic response of a cable-stayed bridge
under non-stationary winds and waves using different surrogate models. Ocean Eng.
CRediT authorship contribution statement 199, 106967.
Fang, C., Xu, Y.-.L., Li, Y., 2022. Optimized C-vine copula and environmental contour of
joint wind-wave environment for sea-crossing bridges. J Wind Eng Ind Aerodyn. 225,
Rugang Yang: Software, Investigation, Writing – original draft. 104989.
Yongle Li: Resources, Supervision. Cheng Xu: Data curation, Software. Haver, S., 1991. Effects of wave directionality and choice of wave spectrum on the
Yi Yang: Visualization, Investigation. Chen Fang: Conceptualization, extreme response of a deep water jacket. Marine Struct. 4, 503–531.
Jahangiri, V., Sun, C., Kong, F., 2021. Study on a 3D pounding pendulum TMD for
Methodology, Software, Funding acquisition, Writing – review & edit­ mitigating bi-directional vibration of offshore wind turbines. Eng. Struct. 241,
ing, Supervision. 112383.
Ji, X., Liu, S., Li, J., Jia, W., 2015. Experimental investigation of the interaction of
multidirectional irregular waves with a large cylinder. Ocean Eng. 93, 64–73.
Declaration of Competing Interest Ji, X., Zou, L., Yang, Z., Wang, D., Bingham, H.B., 2022. Numerical research on the
interaction of multi-directional random waves with an offshore wind turbine
foundation. Ocean Eng. 250, 111029.
The authors declare that they have no known competing financial Li, J., Chen, Y., Pan, S., Pan, Y., Jiang, J., Sowa, D.M.A., 2016. Estimation of mean and
interests or personal relationships that could have appeared to influence extreme waves in the East China Seas. Appl. Ocean Res. 56, 35–47.
Li, Y., Liao, H., Qiang, S., 2004. Simplifying the simulation of stochastic wind velocity
the work reported in this paper.
fields for long cable-stayed bridges. Comput. Struct. 82, 1591–1598.
Meng, S., Ding, Y., Zhu, H., 2018. Stochastic response of a coastal cable-stayed bridge
Data availability subjected to correlated wind and waves. J. Bridge Eng. 23, 04018091.
Montes-Iturrizaga, R., Heredia-Zavoni, E., 2016. Multivariate environmental contours
using C-vine copulas. Ocean Eng. 118, 68–82.
Data will be made available on request. Montes-Iturrizaga, R., Heredia-Zavoni, E., 2015. Environmental contours using copulas.
Appl. Ocean Res. 52, 125–139.
Mo, R., Cao, R., Liu, M., Li, M., 2021. Effect of ground motion directionality on seismic
dynamic responses of monopile offshore wind turbines. Renew. Energy 175,
Acknowledgments 179–199.
Peng, L., 2010. Analysis and design of offshore jacket wind turbine.
The works described in this paper are financially supported by the Philippe, M., Babarit, A., Ferrant, P., 2013. Modes of response of an offshore wind
turbine with directional wind and waves. Renew. Energy, Sel. paper. World Renew.
National Natural Science Foundation of China (Grants No. 52208504) Energy Congress - XI 49, 151–155.
and the Fundamental Research Fund for Central Universities Ti, Z., Wei, K., Qin, S., Li, Y., Mei, D., 2017. Numerical simulation of wave conditions in
(XJ2022002001). Any opinions and conclusions presented in this paper nearshore island area for sea-crossing bridge using spectral wave model. Adv. Struct.
Eng., 136943321773249
are entirely those of the authors. Tian, J., Chen, X., 2020. Evaluation of wind directionality on wind load effects and
assessment of system reliability of wind-excited structures. J. Wind Eng. Ind.
References Aerodyn. 199, 104133.
Valamanesh, V., Myers, A.T., Arwade, S.R., 2015. Multivariate analysis of extreme
metocean conditions for offshore wind turbines. Struct. Saf. 55, 60–69.
Alkarem, Y.R., Ozbahceci, B.O., 2021. A complemental analysis of wave irregularity Wei, K., Arwade, S.R., Myers, A.T., Valamanesh, V., 2016. Directional effects on the
effect on the hydrodynamic response of offshore wind turbines with the semi- reliability of non-axisymmetric support structures for offshore wind turbines under
submersible platform. Appl. Ocean Res. 113, 102757. extreme wind and wave loadings. Eng. Struct. 106, 68–79.
American Petroleum Institute, 2005. Recommended Practice for Planning, Designing, Zhang, X., Chen, X., 2017. Refined process upcrossing rate approach for estimating
and Constructing Fixed Offshore Platforms-Working Stress Design, 21st ed. API probabilistic wind load effects with consideration of directionality. J. Struct. Eng.
Publishing Services, Washington DC. 143, 04016148.
An, Y., Pandey, M.D., 2007. The r largest order statistics model for extreme wind speed Zhang, X., Chen, X., 2015. Assessing probabilistic wind load effects via a multivariate
estimation. J. Wind Eng. Ind. Aerodyn. 95, 165–182. extreme wind speed model: a unified framework to consider directionality and
Bai, X., Jiang, H., Li, C., Huang, L., 2020. Joint probability distribution of coastal winds uncertainty. J. Wind Eng. Ind. Aerodyn. 147, 30–42.
and waves using a log-transformed kernel density estimation and mixed copula Zhu, J., Zhang, W., 2017. Numerical simulation of wind and wave fields for coastal
approach. Ocean Eng. 216, 107937. slender bridges. J. Bridge Eng. 22, 04016125.
Bai, X., Jiang, H., Song, G., Li, X., 2022. Extreme responses of sea-crossing bridges Zhu, J., Zhang, W., Wu, M.X., 2018. Coupled dynamic analysis of the vehicle-bridge-
subjected to offshore ground motion and correlated extreme wind and wave. Ocean wind-wave system. J. Bridge Eng. 23, 04018054.
Eng. 247, 110710.

13

You might also like