5 - Hydroelastic Analysis of A Floating Bridge Under Spatially Inhomogeneous Waves, With Emphasis On The Effect of Drift Force Modeling - Shuai Li

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

Applied Ocean Research 139 (2023) 103666

Contents lists available at ScienceDirect

Applied Ocean Research


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

Research paper

Hydroelastic analysis of a floating bridge under spatially inhomogeneous


waves, with emphasis on the effect of drift force modeling
Shuai Li a,b , Torgeir Moan c,d , Shixiao Fu a,b ,∗, Shiyuan Zhang a,b , Yuwang Xu a,b
a
State Key Laboratory of Ocean Engineering, Shanghai Jiao Tong University, Shanghai, 200240, China
b
Collaborative Innovation Centre for Advanced Ship and Deep-Sea Exploration, Shanghai, 200240, China
c
Department of Marine Technology, Norwegian University of Science and Technology, Trondheim, 7491, Norway
d
Faculty of Maritime and Transportation, Ningbo University, Ningbo, China

ARTICLE INFO ABSTRACT

Keywords: The pontoon-supported floating bridge is an alternative concept for crossing a fjord with a width of up to 5
Floating bridge km. Due to the huge span, the first two natural periods of the bridge are more than 60 s, indicating that the
Hydroelastic analysis corresponding modes can be excited by slowly-varying drift forces on the pontoons. A simplified approach for
Beam-connected-discrete-modules method
determining the drift forces is to assume that the pontoons are independent of the floating bridge, i.e., fixed
Slowly-varying drift force
or freely floating. However, the first-order motions of the pontoons, which are restricted by the deformation
Inhomogeneous wave field
Newman’s approximation
of the floating bridge, have a significant effect on the drift forces. To evaluate the uncertainties implied
by the simplification of drift forces, in this study, a second-order beam-connected-discrete-modules (BCDM)
hydroelastic method is adopted to assess the effect of different drift force models on the floating bridge. In
this method, the first-order response amplitude operators (RAOs) of bridge-restricted pontoons are first solved
in the frequency-domain. Based on the RAOs, the mean drift forces for each pontoon are then determined by
using potential theory. Considering the spatial inhomogeneity of the wave field, the time series of the first- and
second-order wave forces are generated by using the linear transfer functions and Newman’s approximation,
respectively. Then, the method is applied to investigate the hydroelastic responses of a straight side-anchored
floating bridge for the crossing site of Bjørnafjord. The results show that the horizontal displacement is very
sensitive to the different force models implied by various pontoon boundary conditions, i.e., free, fixed, and
bridge-restricted. The drift forces on the bridge-restricted pontoons are approximately 10%–20% larger than
those for the fixed pontoons. Moreover, wave inhomogeneity results in increased vertical displacements and
weak axis bending moments.

1. Introduction eight fjords on route E39 (Moan and Eidem, 2020). Two main types
of floating bridge concepts supported by discrete pontoons are under
A floating bridge is a continuous bridge girder supported by the consideration: the curved end-anchored bridge and the straight side-
buoyancy of several pontoons or floating rafts. Compared with con- anchored bridge. Compared with traditional floating structures, the
ventional pier-supporting bridges, floating bridges are more suitable
design and analysis of the floating bridge crossing a wide and deep fjord
when the water is very deep or the seabed is extremely soft. Moreover,
are very challenging. For instance, the crossing location of Bjørnafjord
employing buoyancy to support the bridge deck can overcome the lim-
itation of the maximum span length for suspension bridges (Watanabe, is approximately 4600 m wide, with a water depth in the middle of
2003). Because of these benefits, floating bridges have been widely more than 500 m. Because of such a huge span, the floating bridge is
constructed worldwide to cross rivers or bays. Examples of floating more flexible, resulting in significant deformations under waves. The
bridges include the Evergreen Point Floating Bridge and the Hood strong coupling between deformation and waves, known as hydroe-
Canal Bridge in the USA, the Nordhordland Bridge and the Bergsøysund lastic responses, needs to be accounted for in the design and safety
Bridge in Norway, the William R. Bennett Bridge in Canada, and the assessment of the bridge. According to a side-anchored bridge concept
Yumemai Bridge in Japan (Xu, 2020). introduced by COWI (COWI, 2016), the bridge has numerous vibration
Currently, the Norwegian Public Roads Administration (NPRA) is
modes with natural periods ranging from 2 to 80 s, suggesting that both
planning to use floating structures, e.g., floating bridges, for crossing

∗ Corresponding author at: State Key Laboratory of Ocean Engineering, Shanghai Jiao Tong University, Shanghai, 200240, China.
E-mail address: shixiao.fu@sjtu.edu.cn (S. Fu).

https://doi.org/10.1016/j.apor.2023.103666
Received 13 May 2023; Received in revised form 14 June 2023; Accepted 13 July 2023
Available online 7 August 2023
0141-1187/© 2023 Elsevier Ltd. All rights reserved.
S. Li et al. Applied Ocean Research 139 (2023) 103666

first- and second-order wave forces should be considered. Moreover, However, these numerical methods cannot be easily applied to ad-
due to the complex topography of the Bjørnafjord, the real wave field dress the second-order wave force on the floating bridge in inhomoge-
in the fjord is spatially inhomogeneous, which indicates that waves neous waves. Recently, based on the linear beam-connected-discrete-
along the bridge might have different wave spectra, wave directions, modules (BCDM) method (Wei et al., 2017, 2018), Li et al. (2022a)
and phase angles (Cheng et al., 2019). first proposed a second-order hydroelastic method for box-type floating
Owing to these design and analysis challenges, floating bridges structures under inhomogeneous wave conditions.
have attracted great research interest, and numerous studies have been In this paper, the method presented by Li et al. (2022a) is applied
carried out. Xiang et al. (2017) investigated the effect of the pontoon’s to investigate the hydroelastic responses of a floating bridge subjected
flange part on the responses of the bridge and found that the viscous to first- and second-order wave loads under inhomogeneous waves.
damping caused by the flange significantly reduces the peaks of the In this method, the first-order frequency-domain hydroelastic equation
responses. The impact of the hydrodynamic interaction between pon- of the bridge is first established to obtain the response amplitude
toons on the global responses of the bridge was also examined by Xiang operators (RAOs) for each pontoon that are restricted by the bridge.
et al. (2018). Cheng et al. (2018a) built a numerical model using the By incorporating these RAOs into the second-order hydrodynamic for-
software SIMO-RIFLEX to compare the responses of the bridge caused mulation, the mean drift forces on each pontoon are calculated. Fi-
by different hydrodynamic load components, including the viscous nally, the time-domain hydroelastic equation is established by applying
drag force, the first-order wave force, and the second-order difference- first- and second-order wave forces onto each pontoon separately,
frequency force. The results revealed that the horizontal displacement considering the spatial wave inhomogeneity. The effect of drift force
is dominated by the second-order force. The effects of the inclination models caused by different pontoon boundary conditions on the re-
of the shoreside on the bridge were studied by Deng et al. (2018). sponses of the bridge is examined. These boundary conditions include
The study showed that the displacement of the bridge increases as the fixed, free, and bridge-restricted (i.e., the pontoon is connected to
steepness of the shoreside rises. A software-to-software comparison of the bridge girder) conditions. The hydroelastic responses in inhomoge-
neous waves, including displacements and bending moments, are also
the dynamic characteristics of a floating bridge was presented by Vi-
compared with those in homogeneous waves. In addition, the effect of
uff et al. (2020) using two popular commercial computer programs,
wave short-crestedness is studied.
OrcaFlex and SIMO-RIFLEX.
The studies mentioned above are based on the assumption of ho-
2. Floating bridge concept
mogeneous wave conditions. Fu et al. (2017) first introduced a linear
time-domain method for hydroelastic analysis of a floating bridge
under inhomogeneous waves, by discretizing the inhomogeneous waves The conceptual design of the straight bridge considered in this study
is proposed by COWI (COWI, 2016). As presented in Fig. 1, the length of
into several isolated homogeneous wave fields. Based on the method
this bridge is approximately 4500 m from the abutment in the south end
of Fu et al. (2017), a sensitivity analysis of several homogeneous wave
to the abutment in the north end. The bridge consists of a cable-stayed
fields was conducted by Wei et al. (2019). Based on a one-year field
high bridge part in the south for the required navigation channel, and
measurement of wave conditions in Cheng et al. (2019), Cheng et al.
a floating bridge part which is the primary component of the bridge.
(2018b) defined three kinds of inhomogeneous wave fields. The effect
The floating bridge part consists of bridge girders, columns, pontoons,
of wave inhomogeneity on the responses of the bridge subjected to
and an abutment at the north end. The girders are supported by 18
first- and second-order wave loads was then discussed. Based on the
equal pontoons through columns with a span of 203 m. The bridge is
coherence functions of wave elevations between different locations, Cui
monolithically connected to the end abutment in the south, while the
et al. (2022) proposed a generic method for assessing the effects of in-
bridge is only allowed to move in the longitudinal direction in the north
homogeneous wave loads on a floating bridge. The study revealed that
to account for the thermal expansion. Additionally, the bridge is also
the influence of wave coherence on the axial force is significant, but
side-anchored to the seabed using mooring lines at three locations.
the effect on the bending moment is negligible. The abovementioned
To investigate the effect of wave loads on the bridge, in this study, a
research on inhomogeneous waves was conducted based on a curved
truncated bridge model is employed based on the floating bridge part,
end-anchored bridge without mooring lines. In contrast, in a straight
while the cable-stayed bridge part is neglected. As shown in Fig. 2(a),
side-anchored bridge, the mooring systems are the critical structural
the total length of the bridge model is 3857 m from the truncated south
components. Li et al. (2018) investigated the effect of inhomogeneous end to the north end. The south end of the bridge is artificially fixed,
waves on mooring tensions by applying the method of Fu et al. (2017). while the translational motions of the bridge girder along the 𝑋-axis at
The effect of various wave inhomogeneities on the fatigue damage in the north end are free to move. Since Cheng et al. (2018a) showed that
mooring lines was investigated by Dai et al. (2021) using different water depth variations do not have a significant effect on wave loads,
fatigue analysis methods. water depth is considered constant for all calculations in this study.
When a floating bridge is exposed to waves, only the supported Fig. 2(b) illustrates the wave directions used in this study. The incident
pontoons are subjected to the wave loads. The interaction between the direction of the wave is parallel to the 𝑋-axis and points to the positive
pontoon and the bridge is neglected in the abovementioned studies direction of the 𝑋-axis under an incident wave angle 𝜃 = 0◦ , and the
when calculating the hydrodynamic forces on the pontoons. This sim- positive direction of the incident wave is counterclockwise.
plification is valid for determining the first-order wave forces, because All 18 pontoons have the same geometry, which consists of two
the Froude-Kriloff and diffraction forces on the pontoon are derived semicircle cylinders at each end and a rectangular box in the middle. To
when the body is restrained from moving. However, the second-order increase the added mass in heave, a 5 m width flange is inserted at the
wave loads strongly depend on the first-order solutions, notably the bottom of the pontoon. However, the viscous damping due to the flange
motions of the pontoons, which are governed by the deformation of the is not considered in this study. The main parameters of the pontoon
whole floating bridge. Therefore, to obtain more realistic second-order without ballast are listed in Table 1. Moreover, different ballasts are
forces on a flexible floating bridge, hydroelastic theory, which examines applied to the pontoons to ensure that all pontoons have the same draft
the interaction between the structural deformation and waves, should (10.5 m) in calm water. Note that the amount of ballast is determined
be applied. Based on the assumption of homogeneous wave conditions, by static analysis of the truncated bridge model. The mass properties of
two main hydroelasticity approaches have been employed for second- the ballast on each pontoon are given in Appendix A. Each pontoon is
order hydroelastic analysis: the modal superposition method (Wu et al., oriented with surge along the global 𝑋-axis and sway along the global
1997; Chen et al., 2003, 2006) and direct method (Park et al., 2017). 𝑌 -axis, as shown in Fig. 2(a).

2
S. Li et al. Applied Ocean Research 139 (2023) 103666

Fig. 1. Straight side-anchored bridge concept by COWI (COWI, 2016).

Fig. 2. Truncated floating bridge model: (a) Side view. (b) Top view and wave directions.

Table 1 the same (Girder Section S1) with a length of 25 m, and the length of
Main parameters of pontoons without the ballast (COWI, 2016). the middle section (Girder Section F1) is 153 m. Table 2 lists the main
Designation Value parameters of the bridge girder and column with respect to their local
Length (m) 28 coordinate systems (see Fig. 2(b)).
Width (m) 68 As shown in Fig. 2, the mooring systems of pontoons P9 and P15
Draft (m) 10.5
Freeboard (m) 4
are the same as that of pontoon P3. Each mooring system is composed
Mass (ton) 11 300 of six identical mooring lines. The pretension of each mooring line
Center of gravity in 𝑍-axis directiona (m) −4.2 is 3800 kN. The detailed properties of the mooring lines are not
Roll inertia (ton⋅m2 ) 4 900 000 presented herein, but they are described in the report by COWI (COWI,
Pitch inertia (ton⋅m2 ) 1 360 000
2016). After the mooring lines are modeled by the catenary equation,
Yaw inertia (ton⋅m2 ) 5 700 000
Water depth (m) 500
the mooring restoring forces on the pontoon due to various offsets
are obtained by OrcaFlex (Orcina, 2022). As shown in Fig. 3, the
a Note that 𝑍 = 0 at the free surface. mooring system shows a nonlinear characteristic, especially for large
horizontal offsets (5 to 10 m). However, when the offset is smaller
than 5 m, the force–displacement relationships can be assumed to be
According to the locations of the two ends and pontoons, the bridge linear. Note that the maximum horizontal displacement under wave
is characterized by 20 locations represented by SE, P1, P2, . . . , P18, loads does not exceed 5 m; thus, using the equivalent stiffness matrix to
and NE, as shown in Fig. 2(a). The bridge girder between every two consider the mooring characteristics is reasonable. The linearized stiff-
locations has an equal length of 203 m, including three segments with nesses of sway–sway, roll–sway, roll–roll, and yaw–yaw are 977 kN/m,
different cross sections. The parameters of the first and last sections are –23260 kN⋅m/m, 17558 kN⋅m/deg, and 7929 kN⋅m/deg, respectively.

3
S. Li et al. Applied Ocean Research 139 (2023) 103666

Fig. 3. Mooring characteristics: (a) Sway force on the pontoon due to sway motion. (b) Roll moment on the pontoon due to sway motion. (c) Roll moment on the pontoon due
to roll rotation. (d) Yaw moment on the pontoon due to yaw rotation.

Table 2 condition 𝑆∞ . The details regarding the solution procedure for velocity
Sectional properties of the bridge girder and column (COWI, 2016); 𝐸𝐼𝑦 and 𝐸𝐼𝑧
potential and wave forces on floating bodies can be found in Molin
represent the bending stiffness about the weak and strong girder axis, respectively;
the axial and torsional stiffness are denoted by 𝐸𝐴 and 𝐺𝐽𝑡 , respectively.
(2022).
Properties Girder S1 Girder F1 Column
According to the nonlinear BCDM method (Li et al., 2022a), the
second-order hydroelastic analysis process of the floating bridge is car-
𝐸𝐴 (kN⋅m2 ) 4.64E+08 3.26E+08 3.28E+08
𝐸𝐼𝑦 (kN⋅m2 ) 3.79E+09 2.68E+09 4.06E+09 ried out in two steps, as presented in Fig. 4. Section 3.1 (Step 1) deals
𝐸𝐼𝑧 (kN⋅m2 ) 3.88E+10 2.65E+10 4.06E+09 with the second-order wave force quadratic transfer functions (QTFs)
𝐺𝐽𝑡 (kN⋅m2 ) 4.61E+09 3.20E+09 3.12E+09 on the bridge-restricted pontoons in the frequency-domain. Then, the
Translation mass (ton/m) 24.92 19.68 14.00
time-domain nonlinear hydroelastic equation considering inhomoge-
Rotation mass (ton⋅m2 ) Neglected Neglected Neglected
neous waves is established in Section 3.2 (Step 2). Note that only the
difference-frequency wave force components of the second-order wave
loads are considered in this study.
3. Methodology
3.1. Second-order wave force on the bridge-restricted pontoon
The finite element method is utilized to model the structural com-
ponents of the bridge. As shown in Fig. 4, the girders and columns After coupling the first-order hydrodynamic forces on each pontoon
are modeled by linear Euler–Bernoulli beam elements, while each with the finite element model, the linear hydroelastic equation of the
pontoon is simplified as a rigid floating body with 6 degrees of freedom floating bridge in the frequency-domain can be written as follows (Li
(DOFs) (COWI, 2016). The flexible deformation of each pontoon can be et al., 2022a),
neglected compared to the global response of the floating bridge, hence ( 2( )
the assumption of a rigid body. Moreover, linear external stiffness −𝜔 [𝑚 + 𝐴 (𝜔)]6𝑁×6𝑁 − 𝑖𝜔[𝐶 (𝜔) + 𝑐]6𝑁×6𝑁
matrices are applied to simulate the characteristics of the mooring [ ] ) { }
+ 𝐾 + 𝐾m + 𝑘 6𝑁×6𝑁 {𝑥̄ (𝜔)}6𝑁×1 = 𝐹̄ (1) (𝜔) 6𝑁×1 (1)
system.
Owing to pontoons with large dimensions compared to the wave where the subscript 𝑁 denotes the number of elastic beam element
amplitude and their own motions, the potential flow theory is employed nodes. [𝑚] refers to the mass matrix of the floating bridge, including the
to estimate the wave-induced loads for pontoons in the frequency do- girders, columns, and pontoons. [𝑐] is the structural damping matrix. [𝑘]
main (Molin, 2022). Based on the potential theory, the fluid is assumed is the stiffness matrix of the bridge. {𝑥̄ (𝜔)} is the complex amplitude of
[ ]
to be incompressible and inviscid, and the flow is irrotational. There- the displacement vector. 𝐾m is the linear equivalent stiffness matrix
fore, the flow is governed by the velocity potential 𝛷(𝑋, 𝑌 , 𝑍, 𝑡) which of the mooring system. [𝐴 (𝜔)], [𝐶 (𝜔)], and [𝐾] are the added mass, hy-
satisfies the Laplace equation in the fluid domain 𝛺, and boundary drodynamic damping, and hydrostatic restoring coefficient matrices of
conditions, including the free-surface condition 𝑆F , the body-surface the pontoons, respectively. {𝐹̄ (1) (𝜔)} is the first-order wave excitation
condition 𝑆0 , the sea bottom condition 𝑆B , and the distant radiation force vector.

4
S. Li et al. Applied Ocean Research 139 (2023) 103666

Fig. 4. Flow diagram of the numerical simulation.

Then, the first-order motions of each pontoon are evaluated by where [𝐴(∞)] is the added mass matrix at the infinite wave frequency.
solving Eq. (1). The first-order radiation potential 𝛷R(1) is obtained
𝑡
[𝐻(𝑡)] is the retardation function matrix, and its integral ∫0 [𝐻(𝑡 −
by multiplying the unit radiation velocity potential 𝜙(1)
𝑗𝑚 and pontoon
𝜏)]{𝑥(𝜏)}𝑑𝜏
̇ represents the fluid memory effect. {𝑥(𝑡)}, {𝑥(𝑡)} ̇ and {𝑥(𝑡)}
̈
(1)
motions 𝜉𝑗𝑚 which are governed by the floating bridge. After adding are the displacement, velocity, and acceleration of the element nodes,
respectively. {𝐹 (1) (𝑡)} and {𝐹 (2) (𝑡)} are the first- and second-order wave
the incident wave potential 𝛷I(1) and the diffraction potential 𝛷D (1)
, the
overall first-order velocity potential 𝛷(1) of the flow field considering excitation forces in the time-domain, respectively, which will be de-
the first-order hydroelastic responses of the floating bridge is obtained. fined in Section 3.2.3.
Finally, the second-order wave force QTFs on each pontoon are esti- Compared to the linear time-domain governing equation proposed
mated based on nonlinear potential theory. Note that the hydrodynamic by Fu et al. (2017), the linear equivalent stiffness matrix [𝐾m ] of
interaction among the pontoons is neglected in this paper due to the mooring systems is added to the left side of the equal sign in Eq. (2),
large spacing between adjacent pontoons (Cheng et al., 2018a). and the second-order wave force time history {𝐹 (2) (𝑡)} is added to the
right side, while the remaining terms are unchanged.
3.2. Nonlinear time-domain hydroelastic equation under inhomogeneous
waves 3.2.2. Wave elevation in inhomogeneous wave conditions
For simulating the homogeneous wave field, also known as the open
3.2.1. The dynamic equation of motion in the time-domain sea, only one wave spectrum is required. For long-crested irregular
According to the method presented by Li et al. (2022a), the nonlin- waves, the wave elevation 𝜂 can be written as the sum of a large number
ear time-domain hydroelastic equation of the bridge can be expressed of wave components,
as follows,

𝐿
[ ]

𝑡
⎞ 𝜂(𝑥, 𝑦, 𝑡) = 𝜂̄𝑙(1) cos 𝜔𝑙 𝑡 + 𝜀𝑙 − 𝑘′𝑙 𝑥 cos 𝜃 − 𝑘′𝑙 𝑦 sin 𝜃 (3)
⎜ [𝑚 + 𝐴(∞)]6𝑁×6𝑁 {𝑥(𝑡)}
̈ 6𝑁×1 + ∫ [𝐻(𝑡 − 𝜏)]6𝑁×6𝑁 {𝑥(𝜏)}
̇ 6𝑁×1 𝑑𝜏 ⎟ 𝑙=1
⎜ [ 0 ] ⎟ √ ( )
⎝ + [𝑐]6𝑁×6𝑁 {𝑥(𝑡)}
̇ 6𝑁×1 + 𝐾 + 𝐾m + 𝑘 6𝑁×6𝑁 {𝑥(𝑡)}6𝑁×1 ⎠ where 𝜂̄𝑙(1) = 2𝑆 𝜔𝑙 𝛥𝜔, 𝜔𝑙 , 𝑘′𝑙 , 𝜀𝑙 and 𝜃 are the wave amplitude,
{ (1) (2)
}
= 𝐹 (𝑡) + 𝐹 (𝑡) 6𝑁×1 (2) circular frequency, wavenumber, random phase and wave propagation

5
S. Li et al. Applied Ocean Research 139 (2023) 103666

Fig. 5. Description of the spatially inhomogeneous irregular wave condition.

direction of the wave component number 𝑙, respectively. 𝑆 (𝜔) is the the 𝑚th pontoon within the 𝑞th region under long-crested waves can
unidirectional wave spectrum. be estimated in the time-domain as follows,
For short-crested waves, a directional wave spectrum 𝑆(𝜔, 𝜃) should
(1)

𝐿
(1) ( )√ ( ) (
(1) ( ))
be used to consider the effect of short-crestedness. 𝑆(𝜔, 𝜃) is a two 𝐹𝑚𝑗 (𝑡) = 𝑓̄𝑚𝑗 𝜔𝑙 , 𝜃 𝑞 2𝑆 𝑞 𝜔𝑙 𝛥𝜔 cos 𝜔𝑙 𝑡 + 𝜀𝑙 + 𝛽𝑚𝑗 𝜔𝑙 , 𝜃 𝑞 (8)
dimensional wave spectrum, which in practice is usually expressed as 𝑙=1
follows, The second-order difference-frequency wave forces under long-crested
waves can be expressed by
𝑆 (𝜔, 𝜃) = 𝑆(𝜔)𝐷(𝜃) (4)
𝐿 √
(2)

𝐿 ∑
( ) √ ( )
where 𝐷(𝜃) is a directional distribution function. 𝜃 is the wave direction 𝐹𝑚𝑗 (𝑡) = 2𝑆 𝑞 𝜔𝑙 𝛥𝜔 2𝑆 𝑞 𝜔𝑘 𝛥𝜔
angle of each wave component. The wave elevation 𝜂 can be expressed 𝑙=1 𝑘=1
[ ( ) (( ) ( )) ]
as (Faltinsen, 1993), 𝑄Re 𝑞
𝑚𝑗 (𝜔𝑙 , 𝜔𝑘 , 𝜃 ) cos(( 𝜔𝑙 − 𝜔𝑘) 𝑡 + ( 𝜀𝑙 − 𝜀𝑘)) +
(9)

𝐿 ∑
𝐺
( ) 𝑄Im 𝑞
𝑚𝑗 𝜔𝑙 , 𝜔𝑘 , 𝜃 sin 𝜔𝑙 − 𝜔𝑘 𝑡 + 𝜀𝑙 − 𝜀𝑘
(1)
𝜂(𝑥, 𝑦, 𝑡) = 𝜂̄𝑙𝑔 cos 𝜔𝑙 𝑡 + 𝜀𝑙𝑔 − 𝑘′𝑙 𝑥 cos 𝜃𝑔 − 𝑘′𝑙 𝑦 sin 𝜃𝑔 (5)
𝑙 𝑔 For short-crested waves, the time series of the first- and second-order
( √ ) forces are expressed as,
(1)
where 𝜂̄𝑙𝑔 = 2𝑆 𝜔𝑙 , 𝜃𝑔 𝛥𝜔𝛥𝜃 and 𝜀𝑙𝑔 are the wave amplitude and
random phase of the wave component for the frequency of 𝜔𝑙 and the (1)

𝐿 ∑
𝐺
(1) ( )√ ( )
𝐹𝑚𝑗 (𝑡) = 𝑓̄𝑚𝑗 𝜔𝑙 , 𝜃𝑔 2𝑆 𝑞 𝜔𝑙 , 𝜃𝑔 𝛥𝜔𝛥𝜃
direction of 𝜃𝑔 , respectively. 𝑙 𝑔
( ))
Thus far, we have introduced the description of a homogeneous (1) (
cos 𝜔𝑙 𝑡 + 𝜀𝑙𝑔 + 𝛽𝑚𝑗 𝜔𝑙 , 𝜃𝑔 (10)
wave field, which means that the significant wave height, wave period,
and wave direction are assumed to be constant in the region where

𝐿 ∑
𝐿 ∑ 𝐺 √
𝐺 ∑
( ) √ ( )
the floating bridge is located. Based on the measured wave data at (2)
𝐹𝑚𝑗 (𝑡) = 2𝑆 𝑞 𝜔𝑙 , 𝜃𝑔 𝛥𝜔𝛥𝜃 2𝑆 𝑞 𝜔𝑘 , 𝜃𝑟 𝛥𝜔𝛥𝜃
three locations in the Bjørnafjord crossing site (Cheng et al., 2019), the 𝑙=1 𝑘=1 𝑔=1 𝑟=1
inhomogeneity in a wave field can be represented by the spatially vary- [ ( ) (( ) ( )) ]
𝑄Re
𝑚𝑗 (𝜔𝑙 , 𝜔𝑘 , 𝜃𝑔 , 𝜃𝑟 ) cos(( 𝜔𝑙 − 𝜔𝑘) 𝑡 + ( 𝜀𝑙𝑔 − 𝜀𝑘𝑟)) +
ing wave spectral parameters (wave heights, periods, and directions) (11)
and the coherence of wave elevations (random phase angles) at these 𝑄Im
𝑚𝑗 𝜔𝑙 , 𝜔𝑘 , 𝜃𝑔 , 𝜃𝑟 sin 𝜔𝑙 − 𝜔𝑘 𝑡 + 𝜀𝑙𝑔 − 𝜀𝑘𝑟
locations. According to Fu et al. (2017), one reasonable assumption to (1) (1)
where 𝑓̄𝑚𝑗 and 𝛽𝑚𝑗 are the amplitude and phase of the linear wave
consider the effect of inhomogeneous wave conditions is to discretize
force transfer functions {𝐹̄ (1) (𝜔)}, respectively. 𝑄Re Im
𝑚𝑗 and 𝑄𝑚𝑗 are the
the inhomogeneous wave field into different regions based on the
real and imaginary components of the difference-frequency QTFs {𝑄},
spatial inhomogeneity characteristics of the wave field, as shown in
respectively.
Fig. 5.
Consequently, under inhomogeneous waves, the time histories of
For long-crested irregular waves, the wave elevation 𝜂 in each
the first- and second-order wave excitation forces on the floating bridge
region can be rewritten as,
in Eq. (2) can be expressed in the form of the following array:
𝐿 √
∑ ( ) ( ) { (1) } { {
(1)
} {
(1)
} {
(1)
} }𝑇
𝜂 𝑞 (𝑥, 𝑦, 𝑡) = 2𝑆 𝑞 𝜔𝑙 𝛥𝜔 cos 𝜔𝑙 𝑡 + 𝜀𝑙 − 𝑘′𝑙 𝑥 cos 𝜃 𝑞 − 𝑘′𝑙 𝑦 sin 𝜃 𝑞 (6) 𝐹 (𝑡) = ⋯ , 𝐹1,1 (𝑡) , … , 𝐹𝑚,6 (𝑡) , … , 𝐹𝑀,6 (𝑡) …
(12)
{ (2) } { { } { } { } }𝑇
𝑙=1
(2) (2) (2)
where 𝑆 𝑞 (𝜔) and 𝜃 𝑞 are the unidirectional wave spectrum and the 𝐹 (𝑡) = … , 𝐹1,1 (𝑡) , … , 𝐹𝑚,6 (𝑡) , … , 𝐹𝑀,6 (𝑡) …
incident wave angle in the 𝑞th region, respectively.
For short-crested waves, the wave elevation 𝜂 in the 𝑞th region can 3.2.4. Newman’s approximation
be expressed as, This paper focuses on the second-order difference-frequency forces,
𝐺 √
also known as slowly-varying drift forces. Although it is preferable

𝐿 ∑
( ) to solve the whole difference-frequency QTFs by near- or middle-field
𝜂 𝑞 (𝑥, 𝑦, 𝑡) = 2𝑆 𝑞 𝜔𝑙 , 𝜃𝑔 𝛥𝜔𝛥𝜃
𝑙 𝑔 methods, it is time-consuming to solve the second-order potential and
( ) requires considerable computational effort. Newman (1974) proposed
cos 𝜔𝑙 𝑡 + 𝜀𝑙𝑔 − 𝑘′𝑙 𝑥 cos 𝜃𝑔 − 𝑘′𝑙 𝑦 sin 𝜃𝑔 (7)
that the QTFs can be approximated by the mean drift force. Different
where 𝑆 𝑞 (𝜔, 𝜃)
= 𝑆 𝑞 (𝜔)𝐷𝑞 (𝜃)
is the directional wave spectrum in the 𝑞th variants for the Newman approximation (de Hauteclocque et al., 2012)
region. Moreover, the random phase angle of the same wave component have been established, and the formula used in this paper is described
is identical in different regions. as (Orcina, 2022),

⎧sgn (𝑄 (𝜔 , 𝜃 )) 𝑄 (𝜔 , 𝜃 ) 𝑄 (𝜔 , 𝜃 )
3.2.3. Wave excitation forces in inhomogeneous wave conditions ⎪ d 𝑙 𝑔 d 𝑙 𝑔 d 𝑘 𝑟
( ( )) ( ( ))
After obtaining the wave elevations, the corresponding time series ( ) ⎪ if sgn 𝑄d 𝜔𝑙 , 𝜃𝑔 = sgn 𝑄d 𝜔𝑘 , 𝜃𝑟
𝑄 𝜔𝑙 , 𝜔𝑘 , 𝜃𝑔 , 𝜃𝑟 ≈ ⎨
of wave excitation forces can be generated based on the first- and ⎪0
second-order wave force transfer functions which have been introduced ⎪ ( ( )) ( ( ))
⎩ if sgn 𝑄d 𝜔𝑙 , 𝜃𝑔 ≠ sgn 𝑄d 𝜔𝑘 , 𝜃𝑟
in Section 3.1. According to the first-order force {𝐹̄ (1) (𝜔)} in the
frequency-domain, the 𝑗th DOF first-order wave excitation forces on (13)

6
S. Li et al. Applied Ocean Research 139 (2023) 103666

Fig. 7. Hydrodynamic meshes of each pontoon.

periods and the wave spectrum. Therefore, the damping coefficients


used are 𝜇 =0.0025 and 𝜆 =0.02 in this study. As a result, the modal
damping ratios for the first two natural frequencies are 1.63% and
1.45%.
The time-domain hydroelastic equation (2) is solved by a step-by-
step numerical integration based on the WBZ-𝛼 (Chung and Hulbert,
1993) integration method in this study. According to Cook et al. (2001),
the time step 𝛥𝑡 is suggested to satisfy the following criterion,
Fig. 6. The workflow of numerical implementation. 𝑇min 𝑇
< 𝛥𝑡 < min (15)
30 10
where 𝑇min denotes the cut-off period, which is defined by the smallest
( )
where 𝑠𝑔𝑛 denotes the sign function. 𝑄d 𝜔𝑙 , 𝜃𝑔 is the mean drift natural period of interest. In this study, only the first-order and second-
force on each pontoon. Note that Eq. (13) is applied to determine the order slowly-varying drift wave loads are considered. 𝑇min =2 s is
slowly-varying drift forces in six DOFs of the pontoons in this study. sufficient during the analysis; thus, 𝛥𝑡 is set to 0.1 s.
For the load cases in Table 4, each time simulation contains an
3.3. Numerical implementation and modeling initial 5000 s of the transient response and subsequent 3-hour steady-
state response. Because the structural and wave radiation damping
Both the boundary element method (BEM) and finite element are relatively small for the low frequency of oscillation, the transient
method (FEM) are employed to calculate the response of the bridge. In responses of the bridge decay very slowly. Therefore, the initial 5000
this study, the BEM is applied to obtain the wave excitation forces and s are removed from the response time series in the post-analysis.
hydrodynamic coefficients of the pontoons in the frequency-domain
by using WAMIT (Lee and Newman, 2014). The FEM is used to 4. Results and discussion
model the floating bridge, which is implemented in an in-house code
called HAISE (Hydroelasticity Analysis under Inhomogeneous Sea En- 4.1. Determination of wave conditions and load cases
vironment), developed by the authors. HAISE is available to estimate
the static and dynamic characteristics of the structure. The dynamic According to Cheng et al. (2019), the waves in the Bjørnafjord are
analysis comprises ‘‘dry’’ and ‘‘wet’’ modal analysis, frequency- and mainly generated by local winds, which are generally short-crested.
time-domain analysis considering wave inhomogeneity. This code has Therefore, the directional wave spectrum 𝑆 (𝜔, 𝜃) = 𝑆(𝜔)𝐷(𝜃) is re-
been fully validated in the frequency- and time-domains (Wei et al., quired to simulate the wind seas. The unidirectional wave spectrum
2017, 2018; Li et al., 2022a,b, 2023), and is thus not given here. 𝑆(𝜔) considered in this study is a 3-parameter JONSWAP spectrum that
The workflow of the present method is a combination of WAMIT is denoted by the significant wave height 𝐻𝑠 , the spectral peak period
and HAISE, as shown in Fig. 6. After meshing the wetted surface of 𝑇𝑝 and the spectral peak parameter 𝛾. The spectral peak parameter 𝛾
pontoons, WAMIT is first employed to calculate the hydrodynamic co- used in all regions is set to 3.3. The directional distribution function
efficients and first-order wave excitation force transfer functions. Then, 𝐷(𝜃) follows the cos𝑛 distribution (DNV, 2010).
the frequency-domain (FD) solver of HAISE is used to solve the first- ( )
𝛤 (1 + 𝑛∕2) | | 𝜋
order coupled RAOs of the bridge and pontoon. The second-order wave 𝐷(𝜃) = √ cos𝑛 𝜃 − 𝜃𝑝 , |𝜃 − 𝜃𝑝 | ≤ (16)
𝜋𝛤 (1∕2 + 𝑛∕2) | | 2
excitation force QTFs for bridge-restricted pontoons are calculated, by
incorporating RAOs into WAMIT. Finally, the time-domain (TD) solver where 𝑛 is the spreading coefficient, which is set to 4 for modeling
of HAISE is employed to estimate the hydroelastic responses of the short-crested waves in this study. 𝛤 denotes the Gamma function, and
floating bridge under inhomogeneous waves. 𝜃𝑝 is the main wave direction.
Fig. 7 shows the hydrodynamic panel model of the submerged part In this study, the 100-year wind wave environment in Bjørnafjord
of each pontoon with 9200 panels. In the finite element modeling of for the main wave direction of 225◦ is chosen, based on hindcast data
the bridge girders and columns, the element lengths are 5–15 m for from 2002 to 2017 (Cheng et al., 2021). Moreover, the wind-generated
the girders, and 3–5 m for the columns, depending on the locations and wave conditions in Bjørnafjord are inhomogeneous due to the complex
girder sections. The sum of the number of elements for the bridge model surrounding landscape. The inhomogeneous wave field for deploying
is 470. Rayleigh damping is used to model the structural damping, and the floating bridge is discretized into 18 regions corresponding to
the resulting modal damping ratio can be expressed as, the 18 locations of the pontoons, as illustrated in Fig. 5. The wave
spectral parameters of each region are then determined based on the
𝜉𝑖 = 0.5(𝜇∕𝜔𝑖 + 𝜆𝜔𝑖 ) (14)
interpolation of a long-term environmental contour map proposed by
where 𝜇 and 𝜆 are the mass and stiffness proportional damping co- Cheng et al. (2021), to consider the spatial variation of the inhomoge-
efficients, respectively. 𝜔𝑖 is the circular frequency corresponding to neous waves, as summarized in Table 3. Note that the contour map is
the 𝑖th wet vibration mode. In this study, a recommendation proposed normalized, and homogeneous wave spectral parameters with 𝐻𝑠 =2.4
by Viuff et al. (2020) is applied to define 𝜇 and 𝜆, which requires a m, 𝑇𝑝 = 5.9 s, and 𝜃𝑝 = 225◦ are required to obtain the actual wave
target damping ratio less than 2% in the frequency range of the natural parameters in an inhomogeneous wave field.

7
S. Li et al. Applied Ocean Research 139 (2023) 103666

Table 3 4.3. Natural periods and mode shapes


Wave spectral parameters of different regions.
Pontoon No 𝐻𝑠 (m) 𝑇𝑝 (s) 𝜃𝑝 (◦ ) Pontoon No 𝐻𝑠 (m) 𝑇𝑝 (s) 𝜃𝑝 (◦ ) Modal analysis is primarily conducted to identify the critical natural
P1 2.40 6.02 220 P10 2.33 5.87 228 periods and mode shapes that might be excited by environmental loads.
P2 2.40 6.01 220 P11 2.30 5.85 229
Table 5 lists the first 20 natural periods for the complete straight
P3 2.40 5.99 221 P12 2.26 5.83 230
P4 2.40 5.98 222 P13 2.23 5.81 232
bridge concept (see Fig. 1) by COWI (COWI, 2016) and the truncated
P5 2.40 5.96 223 P14 2.20 5.78 233 floating bridge model (see Fig. 2) presented in this study. The natural
P6 2.40 5.94 224 P15 2.17 5.75 234 periods of the truncated floating bridge are estimated by using an
P7 2.40 5.93 225 P16 2.10 5.72 236 iterative procedure to consider the influence of frequency-dependent
P8 2.39 5.91 226 P17 2.01 5.69 238
added mass. The mode shapes of the complete and truncated bridges
P9 2.36 5.89 227 P18 1.93 5.65 239
are also compared according to the dominating motions, as shown in
Fig. 10.
In general, the natural periods and mode shapes of the complete
bridge and the truncated floating bridge are reasonably similar. With
regard to the truncated floating bridge model, the first seven modes
are horizontal modes with some contribution from torsion, in which
the primary natural periods are 77.63 s and 68.47 s, respectively. This
indicates that the second-order slowly-varying drift forces dominate
the horizontal displacement. As a result, it is necessary to investigate
the effect of drift force models caused by various pontoon boundary
conditions on the responses of the floating bridge. Due to a large
number of pontoons, the combinations of motions from different pon-
toons result in various mode shapes, but have a slight effect on the
natural periods. From mode 9 and upward, the difference in the natural
periods between the two adjacent modes is smaller than approximately
0.4 s. The lower natural periods overlapping with the main wave
Fig. 8. Wave power spectra at several pontoon positions for the inhomogeneous wave excitation periods in this study, are summarized in Appendix B. The
field. heave motions of pontoons cause eighteen vertical modes with natural
periods ranging from 7.73 to 11.49 s. However, the vertical modes with
lower periods are dominated by pendulum motions of the pontoons and
Under inhomogeneous wave conditions, the pontoons located in the the deformation of girders, as shown in Fig. B.1. Similar to the modal
northern part (P8 to P18) of the bridge are subjected to less 𝐻𝑠 than properties of the curved floating bridge (Cheng et al., 2018a), 15 modes
the pontoons in the southern part (P1 to P7), as presented in Table 3. are dominated by torsional motions with natural periods in the range
Additionally, 𝑇𝑝 and 𝜃𝑝 for each pontoon are obviously different, with of 2.8 to 10.2 s.
Note that the south high bridge part of the complete bridge concept
𝑇𝑝 ranging from 6.02 to 5.65 s and 𝜃𝑝 from 200◦ to 239◦ . The spatial
is simplified as a fixed boundary at the south end of the truncated
variation of the wave parameters, i.e., 𝐻𝑠 , 𝑇𝑝 , and 𝜃𝑝 , implying that
floating bridge model. As a result, differences in the natural periods
the inhomogeneous wave loads on the floating bridge. In addition,
of the two types of bridges are observed, especially for the modes of
the wave parameters of the homogeneous wave field, especially for
the complete bridge with considerable motions at the south end of
𝐻𝑠 , are specified at a higher level than that of the inhomogeneous the floating bridge part, such as modes 3, 4, 5, and 9 (see Fig. 10).
wave field. This also indicates that the ‘‘sum’’ of sea states at each Moreover, this simplification further changes the dominant motions
pontoon for the homogeneous wave is more severe than those for the of the same numbered modes for the two types of bridges. The first
inhomogeneous wave field. Fig. 8 illustrates the wave power spectra at torsional motion mode of the complete bridge is mode 9, while the
several pontoons in inhomogeneous waves. Generally, the wave spectra first torsional mode of the truncated bridge occurs in mode 10, as
in inhomogeneous waves are lower than those in homogeneous waves. shown in Fig. 10(b). Similarly, for the first horizontal motion in the 𝑋-
As listed in Table 4, a series of load cases are defined to investigate axis direction and the first vertical motion in the 𝑍-axis direction, the
the effects of different modeling aspects of wave loads on the dynamic corresponding mode numbers are different for the two types of bridges
responses of the floating bridge, including the wave short-crestedness, (see Fig. 10(c), (d)). Although the local motions at the south end of the
wave inhomogeneity, and second-order slowly-varying drift force mod- two types of bridges are different due to the simplification of the high
els caused by various pontoon boundary conditions. Notably, when bridge part of the complete bridge, the modes all show similar general
determining the drift forces of the pontoons, the most realistic pontoon shapes. This indicates that the modes included in the truncated bridge
boundary condition is bridge-restricted, which considers the coupled model very well represent the dynamic characteristics of the complete
bridge.
motion of the bridge and pontoon.

4.4. First-order RAOs and mean drift forces of each pontoon


4.2. Static responses of the floating bridge in calm water
The mean drift forces are greatly influenced by the motions. Thus,
before determining the drift forces, the first-order RAOs of each bridge-
Fig. 9 depicts the static responses under permanent loads due to restricted pontoon for a wave direction of 225◦ are analyzed in Fig. 11.
the self-weight of the bridge girders, columns, pontoons with ballast, The RAOs vary dramatically with the incident wave periods, and a
and mooring lines, as well as the buoyancy of pontoons in calm water. series of peaks are observed at the natural periods of the floating
As shown in Fig. 9(a), the vertical displacements at each pontoon bridge, especially for sway, roll, and yaw. This indicates that the RAOs
are positive due to the hydrostatic buoyancy. The maximum value of of the pontoons have a strong dependence on the natural vibration
the vertical displacements does not exceed 0.1 m, indicating that the characteristics of the floating bridge. Moreover, due to the various
pontoons are appropriately ballasted. The bending moment and shear positions of the pontoons on the bridge, the RAOs of each pontoon are
force distributions along each span are generally similar. obviously different. For instance, large sway motions mainly occur on

8
S. Li et al. Applied Ocean Research 139 (2023) 103666

Table 4
Load cases and descriptions.
No Wave crest Wave condition Wave force Pontoon boundary conditions
1 Long-crested 𝐻𝑠 = 2.4 m, 𝑇𝑝 = 5.9 s, 𝜃𝑝 = 225◦ 1st –
2 Long-crested 𝐻𝑠 = 2.4 m, 𝑇𝑝 = 5.9 s, 𝜃𝑝 = 225◦ 1st + 2nd Free
3 Long-crested 𝐻𝑠 = 2.4 m, 𝑇𝑝 = 5.9 s, 𝜃𝑝 = 225◦ 1st + 2nd Fixed
4 Long-crested 𝐻𝑠 = 2.4 m, 𝑇𝑝 = 5.9 s, 𝜃𝑝 = 225◦ 1st + 2nd Bridge-restricted
5 Long-crested Inhomogeneous wave presented in Table 3 1st –
6 Long-crested Inhomogeneous wave presented in Table 3 1st + 2nd Free
7 Long-crested Inhomogeneous wave presented in Table 3 1st + 2nd Fixed
8 Long-crested Inhomogeneous wave presented in Table 3 1st + 2nd Bridge-restricted
9 Short-crested 𝐻𝑠 = 2.4 m, 𝑇𝑝 = 5.9 s, 𝜃𝑝 = 225◦ 1st –
10 Short-crested 𝐻𝑠 = 2.4 m, 𝑇𝑝 = 5.9 s, 𝜃𝑝 = 225◦ 1st + 2nd Free
11 Short-crested 𝐻𝑠 = 2.4 m, 𝑇𝑝 = 5.9 s, 𝜃𝑝 = 225◦ 1st + 2nd Fixed
12 Short-crested 𝐻𝑠 = 2.4 m, 𝑇𝑝 = 5.9 s, 𝜃𝑝 = 225◦ 1st + 2nd Bridge-restricted
13 Short-crested Inhomogeneous wave presented in Table 3 1st –
14 Short-crested Inhomogeneous wave presented in Table 3 1st + 2nd Free
15 Short-crested Inhomogeneous wave presented in Table 3 1st + 2nd Fixed
16 Short-crested Inhomogeneous wave presented in Table 3 1st + 2nd Bridge-restricted

Fig. 9. Static structural responses along the floating bridge in calm water: (a) Vertical displacement (VD). (b) Weak axis bending moment (𝑀𝑌 ). (c) Shear force in the 𝑍-axis
(𝐹𝑍 ).

the pontoons near the middle of the bridge, which is consistent with the the pontoons near the two ends of the bridge are larger than those on
mode shapes dominated by horizontal motions, as shown in Fig. 10(a). the pontoons located in the middle region of the bridge.
Fig. 12 shows the surge and sway mean drift forces on the pontoons. To further investigate the effect of various pontoon boundary con-
As the RAOs vary dramatically with the wave periods, so do the drift ditions on the mean drift forces, the comparison of the drift forces on
forces on the pontoons. The drift forces are mainly distributed in the fixed, free, and bridge-restricted pontoons is shown in Fig. 13. The drift
period region corresponding to the vertical modes, because the drift forces at P2 (near the south end), P7 (near the center), and P15 (near
forces are strongly related to the pontoon’s capacity to cause waves. the north end) are used to study the effect of various pontoon boundary
Although the sway motions are significant, with a maximum value of conditions. It should be noted that the drift forces on all fixed pontoons
50 m, their effect on the drift forces is negligible. The drift forces on are the same, but those on free pontoons vary slightly. This is because

9
S. Li et al. Applied Ocean Research 139 (2023) 103666

Fig. 10. Typical mode shapes of the bridge with dominating motions: (a) Y. (b) T. (c) X. (d) V.

Table 5
Natural periods of the complete straight bridge concept and the truncated floating bridge model; X denotes horizontal motion in the 𝑋-axis direction, Y denotes horizontal motion
in the 𝑌 -axis direction, T denotes torsional motion, and V denotes vertical motion in the 𝑍-axis direction.
Mode number COWI (COWI, 2016) (Complete straight bridge) Present (Truncated floating bridge model) DiffL(|𝑇𝐹 − 𝑇𝐶 |)∕𝑇𝐶
Natural period 𝑇𝐶 Dominating motion Natural period 𝑇𝐹 Frequency Dominating motion
(–) (s) Primary Secondary (s) (rad/s) Primary Secondary (%)
1 78.25 Y – 77.63 0.081 Y – 0.80
2 71.21 Y – 68.47 0.092 Y – 3.84
3 40.78 Y T 37.23 0.169 Y T 8.70
4 32.48 Y T 27.95 0.225 Y T 13.96
5 23.44 Y T 20.56 0.306 Y T 12.30
6 17.69 Y T 16.14 0.389 Y T 8.75
7 14.09 X V 12.74 0.493 Y T 9.55
8 14.09 Y T 12.69 0.495 X V 9.91
9 13.06 T Y 11.49 0.547 V X 12.00
10 11.45 V X 11.17 0.563 T Y 2.46
11 11.32 T Y 11.11 0.565 V X 1.84
12 10.92 V – 11.09 0.566 V X 1.59
13 10.89 V X 11.05 0.568 V X 1.50
14 10.88 Y T 11.04 0.569 V X 1.46
15 10.87 V X 11.02 0.570 V X 1.39
16 10.87 V – 10.99 0.572 V X 1.12
17 10.84 V X 10.84 0.579 V X 0.03
18 10.76 V X 10.75 0.585 V X 0.13
19 10.66 V X 10.60 0.593 Y T 0.59
20 10.50 V X 10.54 0.596 V X 0.39

10
S. Li et al. Applied Ocean Research 139 (2023) 103666

Fig. 11. First-order RAOs of the bridge-restricted pontoons: (a) Surge. (b) Sway. (c) Heave. (d) Roll. (e) Pitch. (f) Yaw.

11
S. Li et al. Applied Ocean Research 139 (2023) 103666

Fig. 11. (continued).

Fig. 12. Magnitudes of the mean drift force on the bridge-restricted pontoons: (a) Surge. (b) Sway.

the ballast added to the pontoon has an impact on the gravity part of bending moment 𝑀𝑍 is mainly contributed by the first-order wave
the hydrostatic stiffness matrix. Then, the first-order RAOs and induced forces. The difference in the horizontal displacement due to various
drift forces are not the same, due to the difference in the ballast for slowly-varying drift force models is considerable, while the difference
each pontoon. However, the discrepancies in drift forces between the in 𝑀𝑍 is negligible. Compared with the horizontal displacement of the
free pontoons are not significant in this study, so only the drift forces bridge-restricted condition, also known as the realistic case, the fixed
on the P7 pontoon are presented in Fig. 13. The surge drift forces on and free simplifications underestimate and overestimate the responses,
the bridge-restricted pontoons are larger than those on the fixed and respectively. In addition, the horizontal displacements along the overall
free pontoons. However, a considerable peak is found in the sway drift bridge girder under the homogeneous wave field are all larger than
force on the free pontoon because of the roll resonance of the freely those under the inhomogeneous wave condition. The discrepancy for
moving pontoon. 𝑀𝑍 along the bridge girder ranging from P9 to NE is increased between
homogeneous and inhomogeneous wave conditions.
4.5. Long-crested wave condition To identify the reason for such effects of the different drift force
models and wave inhomogeneity, power spectral analyses of the hor-
The significant difference among the three mean drift force models izontal displacement and strong axis bending moment 𝑀𝑍 at P11 are
caused by different pontoon boundary conditions (see Fig. 13) also shown in Figs. 16 and 17, respectively. The horizontal displacement is
results in an apparent discrepancy in the slowly-varying drift forces. dominated by the first mode resonant response with a natural frequency
This effect on the responses of the floating bridge under long-crested of 0.081 rad/s, and hence is primarily related to the slowly-varying
waves is investigated in this section. drift forces. Therefore, various drift force models result in noticeably
Figs. 14 and 15 show the standard deviation of the horizontal dis- different horizontal displacements. Moan and Eidem (2020) suggested
placement and strong axis bending moment 𝑀𝑍 along the bridge girder that the dynamic features of the bridge can be judged based on the
under different slowly-varying drift force models, respectively. The natural frequencies and mode shapes as well as the spatial distribution
results of only the first-order wave forces are also presented to identify of the excitation force. The effect of the generalized excitation force for
the effect of second-order wave forces. The horizontal displacement the first mode which has a mode shape consisting of approximately one
is dominated by the second-order drift forces, while the strong axis ‘‘half-wave’’ along the overall bridge (see Fig. 10), can be represented

12
S. Li et al. Applied Ocean Research 139 (2023) 103666

Fig. 13. Effect of pontoon boundary conditions on the mean drift force for several pontoons: (a) Surge. (b) Sway.

Fig. 14. Effect of different drift force models on the standard deviation of the horizontal displacement (HD) in the 𝑌 -axis direction along the bridge girder for long-crested waves
under different wave conditions.

by the effect of the sum of wave loads on all pontoons. Therefore, contributed by the mode shape curvature which has a very large value
the ‘‘sum’’ of sea states at each pontoon has a significant effect on for higher order modes with more half-waves.
the horizontal displacement. In this study, the ‘‘sum’’ of sea states The vertical displacement and weak axis bending moment 𝑀𝑌 along
for the homogeneous wave field is more severe than that for the the bridge girder under different slowly-varying drift force models are
inhomogeneous condition, as presented in Table 3; thus, homogeneous analyzed in Fig. 18. The power spectral density (PSD) of the vertical
waves result in a larger horizontal displacement than inhomogeneous displacement and weak axis bending moment 𝑀𝑌 of the bridge girder
waves. at two positions (P2 and P15) are analyzed in Fig. 19. The vertical
With regard to the strong axis bending moment 𝑀𝑍 , numerous displacement and weak axis bending moment 𝑀𝑌 with or without drift
modes, especially those with natural frequencies overlapping the wave forces are almost the same, for both homogeneous and inhomogeneous
frequencies, are mainly involved in the contribution to 𝑀𝑍 . Although waves. This is because the primary vertical modes, with natural periods
different drift force models also have a significant effect on 𝑀𝑍 at ranging from 7.73 to 11.49 s, are not excited by the slowly-varying drift
frequencies below 0.3 rad/s, these low-frequency responses are not forces, as shown in Fig. 19. In addition, compared with the vertical
dominant for the complete 𝑀𝑍 . This is because the bending moment is responses at P18, a larger vertical displacement and 𝑀𝑌 at P1 are

13
S. Li et al. Applied Ocean Research 139 (2023) 103666

Fig. 15. Effect of different drift force models on the standard deviation of the strong axis bending moment (𝑀𝑍 ) along the bridge girder for long-crested waves under different
wave conditions.

Fig. 16. Power spectral density of the horizontal displacement (HD) in the 𝑌 -axis direction at P11 for long-crested waves under different wave conditions: (a) Homogeneous waves.
(b) Inhomogeneous waves.

Fig. 17. Power spectral density of the strong axis bending moment (𝑀𝑍 ) at P11 for long-crested waves under different wave conditions: (a) Homogeneous waves. (b) Inhomogeneous
waves.

found. The main factor for this discrepancy might be the difference in region between P1 and P7. The difference in responses between homo-
extra vertical bending moments which is induced by the surge force on geneous and inhomogeneous waves could also be explained through
pontoons and the distance, i.e., the column length between pontoons generalized excitation forces for vertical modes. Fig. 18 suggests that
and girders. The column length at P1, which is located in the high part the vertical displacement and 𝑀𝑌 are contributed by several modes
of the floating bridge, is larger than that at P18, as shown in Fig. 2. with numerous half-waves, such as mode 29 and mode 32, since the
Under homogeneous waves, the magnitude of the surge force for all peak values are all observed at the position of each pontoon. This
pontoons is the same, whereas under inhomogeneous waves, the surge indicates that the vertical responses of the bridge at different positions
force is larger at P1 than at P18 due to the more severe sea state at P1. are mainly related to the wave loads of the pontoons located at the
In the southern part of the floating bridge between SE and P7, the corresponding positions. Therefore, the local sea state for the pontoon
vertical displacement and weak axis bending moment 𝑀𝑌 under inho- and the wave excitation force in the surge, heave, and pitch DOFs,
mogeneous waves are slightly larger than those under homogeneous could be the major factors, dominating the vertical responses. Under
waves, although 𝐻𝑠 of the two wave conditions are the same in the the inhomogeneous wave field, 𝑇𝑝 is larger than 5.9 s from P1 to P7,

14
S. Li et al. Applied Ocean Research 139 (2023) 103666

Fig. 18. Effect of different drift force models on the standard deviation of the vertical displacement and weak axis bending moment along the bridge girder under long-crested
waves: (a) Vertical displacement (VD). (b) Weak axis bending moment (𝑀𝑌 ).

Fig. 19. Power spectral density of vertical displacement (VD) and weak axis bending moment (𝑀𝑌 ) at P2 and P15 for long-crested waves: (a) VD at P2. (b) 𝑀𝑌 at P2. (c) VD at
P15. (d) 𝑀𝑌 at P15.

15
S. Li et al. Applied Ocean Research 139 (2023) 103666

Fig. 20. Effect of different drift force models on the standard deviation of the horizontal displacement and strong axis bending moment along the bridge girder under short-crested
waves: (a) Horizontal displacement (HD). (b) Strong axis bending moment (𝑀𝑍 ).

Fig. 21. Effect of different drift force models on the standard deviation of the vertical displacement and weak axis bending moment along the bridge girder under short-crested
waves: (a) Vertical displacement (VD). (b) Weak axis bending moment (𝑀𝑌 ).

which is closer to the natural periods of vertical modes (see Table B.1). To summarize, the horizontal displacement is more sensitive to
Another factor is that the 𝜃𝑝 from P1 to P7 is smaller than 225◦ , various drift force models, while the vertical displacement, as well as
resulting in increased excitation forces of the pontoons. the strong and weak axis bending moments, are not. The assumption

16
S. Li et al. Applied Ocean Research 139 (2023) 103666

Table A.1
Mass of ballast in pontoons.
Pontoon No Mass (ton) Center of gravity Roll inertiab (ton m2 ) Pitch inertiab (ton m2 ) Yaw inertiab (ton m2 )
in 𝑍-axis directiona (m)
P1 2 735 −10.17 901 980 163 589 1 065 623
P2 2 785 −10.17 918 418 166 570 1 085 044
P3 1 050 −10.37 346 187 62 787 408 995
P4 2 988 −10.14 985 466 178 730 1 164 256
P5 3 101 −10.13 1 022 961 185 531 1 208554
P6 3 201 −10.12 1 055 931 191 510 1 247 505
P7 3 216 −10.11 1 060 733 192 381 1 253 179
P8 3 216 −10.11 1 060 733 192 381 1 253 179
P9 1 391 −10.33 458 949 83 238 542 215
P10 3 216 −10.11 1 060 733 192 381 1 253 179
P11 3 216 −10.11 1 060 733 192 381 1 253 179
P12 3 216 −10.11 1 060 733 192 381 1 253 179
P13 3 216 −10.11 1 060 733 192 381 1 253 179
P14 3 216 −10.11 1 060 733 192 381 1 253 179
P15 1 391 −10.33 458 949 83 238 542 215
P16 3 216 −10.11 1 060 733 192 381 1 253 179
P17 3 216 −10.11 1 060 733 192 381 1 253 179
P18 3 216 −10.11 1 060 733 192 381 1 253 179
a
Relative to the water line.
b
Relative center of gravity of the ballast.

Table B.1
Natural periods of mode 21 to mode 60 for the truncated floating bridge model; X denotes horizontal motion in the 𝑋-axis direction, Y denotes horizontal motion in the 𝑌 -axis
direction, T denotes torsional motion, and V denotes vertical motion in the 𝑍-axis direction.
Modenumber Natural period Frequency Dominating motion Mode number Natural period Frequency Dominating motion
(–) (s) (rad/s) Primary Secondary (–) (s) (rad/s) Primary Secondary
21 10.27 0.612 V X 41 4.82 1.304 X V
22 10.13 0.621 T Y 42 4.80 1.308 T Y
23 9.92 0.634 V X 43 4.55 1.380 T Y
24 9.41 0.668 V X 44 4.47 1.404 T Y
25 9.02 0.696 Y T 45 4.34 1.449 X V
26 8.97 0.700 V X 46 4.30 1.463 T Y
27 8.70 0.722 T Y 47 4.11 1.528 T Y
28 8.54 0.736 V X 48 4.05 1.552 T Y
29 8.14 0.771 V X 49 4.03 1.561 T Y
30 7.90 0.796 V X 50 3.85 1.630 X V
31 7.85 0.801 T Y 51 3.59 1.748 X V
32 7.73 0.812 V X 52 3.57 1.759 T Y
33 7.54 0.833 Y T 53 3.48 1.805 X V
34 6.88 0.913 Y T 54 3.38 1.858 X V
35 6.56 0.958 T Y 55 3.20 1.962 X V
36 6.04 1.040 Y T 56 3.17 1.983 T Y
37 5.76 1.090 T Y 57 3.06 2.055 X V
38 5.75 1.093 X V 58 2.89 2.172 X V
39 5.36 1.173 T Y 59 2.81 2.239 Y T
40 5.09 1.236 Y T 60 2.79 2.248 X V

of free pontoons overestimates the horizontal displacement, but fixed In general, the effects of wave inhomogeneity and different drift
pontoons underestimate it. However, this obvious discrepancy in the force models on both the vertical and horizontal responses in short-
horizontal displacement might not result in a significant difference crested waves are similar to those observed in long-crested wave con-
in the acceleration, due to the relatively low frequency. Compared ditions. The causes of such effects have been discussed in Section 4.5.
with homogeneous waves, inhomogeneous waves cause smaller hori- On the other hand, under short-crested waves, both the horizontal
zontal displacements and strong axis bending moments. However, the displacement and strong axis bending moment 𝑀𝑍 are smaller than
maximum values of both the vertical displacement and the weak axis those under long-crested waves, as shown in Fig. C.1. However, short-
bending moment are observed in inhomogeneous waves, because the crested waves result in larger vertical displacement and weak axis
vertical responses depend more on the local sea state for each pontoon.
bending moment 𝑀𝑌 (see Fig. C.2). In the present case study, the wave
short-crestedness seems to increase in the maximum 𝑀𝑌 by 46% for ho-
4.6. Short-crested wave condition
mogeneous waves and 35% for inhomogeneous waves. This implies that
Wind-induced waves in fjords are usually short-crested. The sen- the vertical responses induced by wave components in other directions,
sitivity of the slowly-varying drift response to the different pontoon such as 200◦ to 250◦ , are larger than those induced by the main wave
boundary conditions of the pontoon under short-crested waves, is direction of 225◦ . Note that the validity of Newman’s approximation
studied in this section. The standard deviation of the horizontal dis- for short-crested waves is not certain SINTEF (2017), Kim and Yue
placement and strong axis bending moment 𝑀𝑍 along the bridge girder (1989), because Eq. (13) neglects the additional interaction due to the
under different slowly-varying drift force models in short-crested waves presence of two waves with the same frequency but different directions
are shown in Fig. 20. Fig. 21 illustrates the standard deviation of the of propagation. This indicates the uncertainty in discussing the effect
vertical displacement and weak axis bending moment 𝑀𝑌 . of short crestedness on second-order responses.

17
S. Li et al. Applied Ocean Research 139 (2023) 103666

Fig. B.1. Mode shapes of mode 21 to mode 60 for the truncated floating bridge model.

5. Summary and conclusions displacement (VD) and the bending moments (𝑀𝑌 and 𝑀𝑍 ) are mainly
contributed by first-order wave forces.
In this paper, a numerical study is carried out to determine the first- (2) The horizontal displacement is significantly affected by different
and second-order hydroelastic responses of a straight side-anchored drift force models. Compared with the horizontal displacement of the
pontoon-supported floating bridge in inhomogeneous waves, based on bridge-restricted condition which is the real boundary condition in the
the beam-connected-discrete-modules (BCDM) hydroelasticity method. floating bridge, the fixed and free simplifications underestimate and
The first-order RAOs of the pontoons, which are restricted by the overestimate the responses, respectively.
deformation of the bridge girder, are first solved in the frequency- (3) Short-crested waves result in larger vertical responses (VD and
domain. Comparing the mean drift forces caused by various pontoon 𝑀𝑌 ) than long-crested waves but cause smaller horizontal responses
boundary conditions, i.e., fixed, free, and bridge-restricted, it is found (HD and 𝑀𝑍 ) in the present study. This indicates that wave short-
that drift forces are mainly dominated by the heave motions of pon- crestedness should be considered when investigating the wave load
toons. Note that the bridge-restricted boundary condition considers effect of floating bridges.
the coupled motion of the bridge and pontoon when determining the (4) The wave inhomogeneity associated with the main wave direc-
drift forces of the pontoons. Then, Newman’s approximation is applied tion 𝜃𝑝 , and the wave peak period 𝑇𝑝 , causes larger vertical displace-
to obtain the second-order slowly-varying drift forces in six DOFs ment and weak axis bending moment, because the vertical responses
of the pontoons. Finally, the effects of the second-order wave force, depend more on the local wave conditions for each pontoon. However,
wave short-crestedness, and wave inhomogeneity on the displacements the horizontal displacement is more related to the ‘‘sum’’ of sea states
and bending moments of the bridge girder are studied, and several at each pontoon, and a smaller horizontal displacement is therefore
conclusions can be drawn: observed under inhomogeneous waves.
(1) Compared with the first-order responses, second-order forces In conclusion, this study proposes an approach to obtain the wave
cause a considerable horizontal displacement (HD), because the first drift forces considering the coupled motion of the bridge and pontoon,
mode is excited by low-frequency wave forces. However, the vertical and to determine the induced second-order hydroelastic responses.

18
S. Li et al. Applied Ocean Research 139 (2023) 103666

Fig. C.1. Effect of short-crested waves and wave inhomogeneity on the standard deviation of the horizontal displacement and strong axis bending moment along the bridge girder:
(a) Horizontal displacement (HD). (b) Strong axis bending moment (𝑀𝑍 ).

Fig. C.2. Effect of short-crested waves and wave inhomogeneity on the standard deviation of the vertical displacement and weak axis bending moment along the bridge girder:
(a) Vertical displacement (VD). (b) Weak axis bending moment (𝑀𝑌 ).

However, since only one main wave direction (225◦ ) is considered The uncertainty of Newman’s approximation for short-crested waves
in this study, more wave conditions are required to enhance the and forces in the vertical plane, as well as its applicability for pontoon-
conclusions. In addition, the floating bridge response is also affected by supported floating bridges, will be studied by complete bi-frequency
wind and current loads, which should be considered in future studies. and bi-directional QTFs in the future.

19
S. Li et al. Applied Ocean Research 139 (2023) 103666

CRediT authorship contribution statement Chung, J., Hulbert, G.M., 1993. A time integration algorithm for structural dynamics
with improved numerical dissipation: the generalized-𝛼 method. J. Appl. Mech. 60
(2), 371–375.
Shuai Li: Conceptualization, Methodology, Investigation, Writing – Cook, R.D., Malkus, D.S., Plesha, M.E., Witt, R.J., 2001. Concepts and Applications of
original draft, Visualization. Torgeir Moan: Conceptualization, Writing Finite Element Analysis, fourth ed. Wiley.
– review & editing. Shixiao Fu: Conceptualization, Writing – review & COWI, 2016. Straight bridge navigation channel in south summary of analyses. Techni-
editing, Supervision, Funding acquisition. Shiyuan Zhang: Writing – cal Report NOT-KTEKA-020, Report for the Norwegian Public Road Administration,
COWI AS, Oslo, Norway.
review & editing. Yuwang Xu: Writing – review & editing. Cui, M., Cheng, Z., Moan, T., 2022. A generic method for assessment of inhomogeneous
wave load effects of very long floating bridges. Mar. Struct. 83, 103186.
Declaration of competing interest Dai, J., Leira, B.J., Moan, T., Alsos, H.S., 2021. Effect of wave inhomogeneity on
fatigue damage of mooring lines of a side-anchored floating bridge. Ocean Eng.
219, 108304.
The authors declare that they have no known competing finan- de Hauteclocque, G., Rezende, F., Waals, O., Chen, X., 2012. Review of approximations
cial interests or personal relationships that could have appeared to to evaluate second-order low-frequency load. In: International Conference on
influence the work reported in this paper. Offshore Mechanics and Arctic Engineering. Vol. 44885, American Society of
Mechanical Engineers, pp. 363–371.
Deng, S., Fu, S., Moan, T., Wei, W., Gao, Z., 2018. Hydro-elastic analysis of a floating
Data availability bridge in waves considering the effect of the hydrodynamic coupling and the shore
sides. In: International Conference on Offshore Mechanics and Arctic Engineering.
Vol. 51203, American Society of Mechanical Engineers, V001T01A039.
The authors are unable or have chosen not to specify which data
DNV, 2010. Recommended practice DNV-RP-C205: environmental conditions and
has been used. environmental loads. Technical Report, DNV, Oslo, Norway.
Faltinsen, O.M., 1993. Sea Loads on Ships and Offshore Structures. Vol. 1, Cambridge
Acknowledgments University Press.
Fu, S., Wei, W., Ou, S., Moan, T., Deng, S., Lie, H., 2017. A time-domain method for
hydroelastic analysis of floating bridges in inhomogeneous waves. In: International
The authors gratefully acknowledge the National Natural Science Conference on Offshore Mechanics and Arctic Engineering. Vol. 57779, American
Foundation of China (Grant No. 52088102, 52271283, 52111530135), Society of Mechanical Engineers, V009T12A010.
Kim, M.H., Yue, D.K.P., 1989. Slowly-varying wave drift forces in short-crested irregular
National Science Fund for Distinguished Young Scholars, China (Grant
seas. Appl. Ocean Res. 11 (1), 2–18.
No. 51825903), Joint Funds of the National Natural Science Founda- Lee, C.H., Newman, J.N., 2014. Wamit User Manual Version 6.4S. WAMIT Inc,
tion of China (Grant No. U19B2013), Shanghai Science and Technol- Massachusetts Institute of Technology, Massachusetts, USA.
ogy Program, China (Grant No. 19XD1402000 & 19JC1412801), Key Li, S., Fu, S., Wei, W., Moan, T., 2018. A comparison study on the hydroelasticity of
two types of floating bridges in inhomogeneous wave conditions. In: International
Projects for Intergovernmental Cooperation in International Science,
Conference on Offshore Mechanics and Arctic Engineering. Vol. 51265, American
Technology and Innovation, China (Grant No. 2018YFE0125100), Nat- Society of Mechanical Engineers, V07AT06A045.
ural Science Foundation of Shanghai, China (Grant No. 22ZR1434100), Li, S., Fu, S., Zhang, S., Moan, T., 2022a. Second-order hydroelastic analysis of a flexible
and Chenguang Program of Shanghai Education Development Founda- floating structure under spatially inhomogeneous waves. Mar. Struct. 86, 103306.
Li, S., Fu, S., Zhang, S., Xu, Y., 2023. Second-order sum-frequency hydroelastic analysis
tion and Shanghai Municipal Education Commission, China (Grant No.
of a flexible structure in bichromatic waves. Ocean Eng. 269, 113485.
19CG10). Li, S., Fu, S., Zhang, S., Xu, Y., Song, B., 2022b. Numerical study of mean drift
force on flexible structures. In: Proceedings of the 9th International Conference
Appendix A. The ballast of each pontoon Hydroelasticity in Marine Technology.
Moan, T., Eidem, M.E., 2020. Floating bridges and submerged tunnels in Norway—The
history and future outlook. In: Wang, C.M., Lim, S.H., Tay, Z.Y. (Eds.), WCFS2019.
Table A.1 lists the mass properties of the ballast on each pontoon. Springer Singapore, Singapore, pp. 81–111.
Molin, B., 2022. Offshore Structure Hydrodynamics. Cambridge University Press.
Newman, J.N., 1974. Second order slowly varying forces on vessels in irregular waves.
Appendix B. Lower natural periods and mode shapes (mode 21 to In: Symposium on Dynamics of Marine Vehicles and Structures in Waves. London,
mode 60) UK, pp. 182–186.
Orcina, 2022. OrcaFlex user manual: OrcaFlex version 11.2b. Orcina Ltd, Cumbria, UK.
Park, D.M., Kim, J.H., Kim, Y., 2017. Numerical study of mean drift force on stationary
The natural periods and corresponding mode shapes for mode 21 to
flexible barge. J. Fluids Struct. 74, 445–468.
mode 60 are presented in Table B.1 and Fig. B.1, respectively. SINTEF, 2017. Simo Theory Manual Version 4.10. SINTEF Ocean, Trondheim, Norway.
Viuff, T., Xiang, X., Leira, B.J., Øiseth, O., 2020. Software-to-software comparison of
end-anchored floating bridge global analysis. J. Bridge Eng. 25 (5), 04020022.
Appendix C. Effects of short-crested waves and wave inhomogene-
Watanabe, E., 2003. Floating bridges: past and present. Struct. Eng. Int. 13 (2),
ity 128–132.
Wei, W., Fu, S., Moan, T., Lu, Z., Deng, S., 2017. A discrete-modules-based frequency
The second-order hydroelastic responses, as illustrated in Figs. C.1 domain hydroelasticity method for floating structures in inhomogeneous sea
conditions. J. Fluids Struct. 74, 321–339.
and C.2, are obtained by bridge-restricted drift force models. Wei, W., Fu, S., Moan, T., Song, C., Deng, S., Lie, H., 2019. A time-domain method for
hydroelasticity of a curved floating bridge in inhomogeneous waves. J. Offshore
References Mech. Arct. Eng. 141 (1).
Wei, W., Fu, S., Moan, T., Song, C., Ren, T., 2018. A time-domain method for
hydroelasticity of very large floating structures in inhomogeneous sea conditions.
Chen, X., Moan, T., Fu, S., Cui, W., 2006. Second-order hydroelastic analysis of a
Mar. Struct. 57, 180–192.
floating plate in multidirectional irregular waves. Int. J. Non-Linear Mech. 41 (10),
Wu, Y., Maeda, H., Kinoshita, T., 1997. The second order hydrodynamic actions on a
1206–1218.
flexible body. J. Inst. Ind. Sci., Univ. Tokyo 49 (4), 8–19.
Chen, X., Wu, Y., Cui, W., Tang, X., 2003. Nonlinear hydroelastic analysis of a moored
Xiang, X., Svangstu, E., Nedrebø, Ø., Jakobsen, B., Eidem, M.E., Larsen, P.N., Sørby, B.,
floating body. Ocean Eng. 30 (8), 965–1003.
2017. Viscous damping modelling of floating bridge pontoons with heaving skirt
Cheng, Z., Gao, Z., Moan, T., 2018a. Hydrodynamic load modeling and analysis of a and its impact on bridge girder bending moments. In: International Conference
floating bridge in homogeneous wave conditions. Mar. Struct. 59, 122–141. on Offshore Mechanics and Arctic Engineering. Vol. 57748, American Society of
Cheng, Z., Gao, Z., Moan, T., 2018b. Wave load effect analysis of a floating bridge in Mechanical Engineers, V07BT06A001.
a fjord considering inhomogeneous wave conditions. Eng. Struct. 163, 197–214. Xiang, X., Viuff, T., Leira, B., Øiseth, O., 2018. Impact of hydrodynamic interaction
Cheng, Z., Svangstu, E., Gao, Z., Moan, T., 2019. Field measurements of inhomogeneous between pontoons on global responses of a long floating bridge under wind waves.
wave conditions in Bjørnafjorden. J. Waterw. Port Coast. Ocean Eng. 145 (1), In: International Conference on Offshore Mechanics and Arctic Engineering. Vol.
05018008. 51265, American Society of Mechanical Engineers, V07AT06A049.
Cheng, Z., Svangstu, E., Moan, T., Gao, Z., 2021. Assessment of inhomogeneity in Xu, Y., 2020. Floating bridge. In: Cui, W., Fu, S., Hu, Z. (Eds.), Encyclopedia of Ocean
environmental conditions in a Norwegian fjord for design of floating bridges. Ocean Engineering. Springer Singapore, Singapore, pp. 1–10.
Eng. 220, 108474.

20

You might also like