Download as pdf or txt
Download as pdf or txt
You are on page 1of 44

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/344448008

Experimental and numerical study of free-surface wave resonance in the gap


between two elongated parallel boxes with square corners

Article in Applied Ocean Research · October 2020


DOI: 10.1016/j.apor.2020.102376

CITATION READS

1 157

5 authors, including:

Hongchao Wang Wenhua Zhao


Technology Centre for Offshore and Marine, Singapore University of Western Australia
8 PUBLICATIONS 17 CITATIONS 75 PUBLICATIONS 582 CITATIONS

SEE PROFILE SEE PROFILE

Scott Draper Paul H. Taylor


University of Western Australia University of Oxford
117 PUBLICATIONS 1,460 CITATIONS 207 PUBLICATIONS 3,485 CITATIONS

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Mooring system View project

system dynamics View project

All content following this page was uploaded by Hongchao Wang on 02 October 2020.

The user has requested enhancement of the downloaded file.


Experimental and numerical study of free-surface wave resonance in the gap
between two elongated parallel boxes with square corners

Hongchao Wanga,b,∗, Wenhua Zhaob , S. Draperb , H. A. Wolgamotb , P. H. Taylorb


a Technology Centre for Offshore and Marine, Singapore (TCOMS), Singapore 118411, Singapore
b Oceans Graduate School, Faculty of Engineering and Mathematical Sciences, The University of Western Australia, Crawley
WA 6009, Australia

Abstract

Experiments with wave groups are carried out to investigate 3D wave resonance in the gap between two fixed
parallel elongated boxes with square corners. The gap responses are re-created in a numerical wave tank
using computational fluid dynamics (CFD) with reasonable accuracy. Based on the CFD results with high
temporal and spatial resolution, the free-surface behaviour, internal flow velocity and vortex structures near
the gap are interrogated in detail to gain insights into the behaviour. It is found that the damping which
affects the oscillatory motion in the gap appears to be linear (or at least dominated by linear damping)
except for the largest excitation, though substantial flow separation is present. A phase-based harmonic
separation method is used to extract the harmonics of the wave field, which is found useful in understanding
the dynamics of wave-structure interactions for gap resonance problems.
Keywords: Gap resonance; harmonics; experiment; CFD; focused wave group; free-surface wave

1 1. Introduction

2 Offloading of liquefied natural gas (LNG) from a floating liquefied natural gas (FLNG) vessel to an LNG
3 carrier (LNGC) in a side-by-side fashion creates a long and relatively narrow gap, in which large resonant
4 free-surface motions (referred to as ‘gap resonance’) may occur under the excitation of incident waves with
5 appropriate frequencies. Gap resonances are essentially standing waves in a gap of length close to an integer
6 number of half-wavelengths for each standing wave. For such highly resonant systems, the magnitude of
7 each gap resonance depends on the system damping. For very narrow gaps, the viscous damping becomes
8 relatively more important as radiation damping is small.
9 Early studies based on linearised potential flow theory have proven to give good estimates of resonant
10 frequencies for gap resonance (Molin, 2001; Molin et al., 2002; Sun et al., 2010), whereas over-prediction
11 of the magnitude of the gap response is a well-known problem because viscous damping is not included.
12 To improve the agreement between numerical simulations and experimental data, various methods have
13 been proposed by integrating additional artificial damping into potential flow models (Huijsmans et al.,

∗ Correspondingauthor
Email address: hongchao.wang2013@gmail.com (Hongchao Wang)

Preprint submitted to Applied Ocean Research September 29, 2020


14 2001; Lee and Newman, 2005; Chen, 2005). Nevertheless, the choice of such artificial damping requires
15 calibration against results from experiments or computational fluid dynamics (CFD) studies – emphasizing
16 the importance of these – and there are no general rules available for the tuning of such artificial damping
17 (Pauw et al., 2007).
18 Model tests in a physical wave basin have been important in investigating gap resonance. A relatively
19 large number of early studies have focused on experiments in 2D (e.g. Saitoh et al. (2006); Faltinsen et al.
20 (2007); Kristiansen and Faltinsen (2009)), with rather fewer in 3D due to the need to use a wave basin. Molin
21 et al. (2009) carried out 3D model tests to investigate gap resonance in between two fixed, rigidly connected
22 side-by-side rectangular barges with round and square bilges under the excitation of irregular waves with
23 different headings. The minimum gap width was 9 m at full-scale (based on scaling by a standard LNGC
24 width of 46 m). Perić and Swan (2015) carried out 3D model tests to study the fluid behaviour in the
25 gap between a full depth box (representing a gravity-based structure) and either a fixed ship-shaped box,
26 or a floating ship, in close proximity. The gap width was 6 m at full-scale. Recently, Zhao et al. (2017)
27 conducted 3D wave basin experiments in transient wave groups to investigate resonant fluid motions in a gap
28 of (full-scale) width 4 m (close to the gap width for LNG offloading operations), for fixed vessels with round
29 and square bilges. In both cases substantial extra damping in addition to radiation damping was clearly
30 present, and for the round bilge cases it was suggested that the damping was almost entirely due to laminar
31 boundary layers. They also found that the gap responses can be strongly driven via frequency-doubling
32 excitation, such that the second harmonic component can be larger than the linear gap response. Chua et al.
33 (2018) conducted 3D model tests to investigate gap responses under different irregular seas conditions and
34 bilge geometries, where the gap width corresponded to 8 m at full-scale.
35 Whilst experiments in wave basins have been important tools for understanding gap resonance problems,
36 measurements of the structure of the flow field below the free-surface have not been attempted. A numerical
37 wave tank (NWT) based on a Navier-Stokes (NS) solver, taking both viscous effects and effects associated
38 with the nonlinear free-surface conditions into account, allows examination of the details of the flow field
39 and may be able to predict both resonant frequencies and amplitudes accurately without any parametrical
40 tuning. A relatively large number of studies have demonstrated that CFD can predict gap responses (excited
41 by regular waves) well in 2D, providing obvious improvement compared to potential flow models (Lu et al.,
42 2008; Sauder et al., 2010; Moradi et al., 2015; Jiang et al., 2019). Despite the extensive application in 2D,
43 CFD has been applied less in 3D where the computational demand is much greater, and the set-up of a
44 3D NWT is much more challenging and requires careful selection of relevant parameters. Feng et al. (2017)
45 simulated regular waves interacting with the square bilge version of the 3D model tested by Molin et al. (2009)
46 and found reasonable agreement with experimental response curves (generated from irregular wave tests).
47 For each frequency considered, it was found necessary to run 20-50 wave periods, to ensure that steady state
48 was reached. Jin et al. (2018) employed a viscous Reynolds-averaged NS solver to investigate hydrodynamic
49 interactions of a 3D conceptual FLNG-LNGC offloading (10 m gap) system with regular waves, which offered

2
(a) Plan view
49.6 7.6

wave paddles

20
0.45
(b) A–A cross section
WG 1
WG 2

Wave absorber
0.833
WG 3 0.067
19.83 0.50 29.77

3.333
40

WG 4

0.185
WG 5
0.767
WG 6
WG 7

A A
20

Figure 1: Definition sketch of the experimental set-up for beam sea conditions (all dimensions in m, note that the gap width is
scaled up by 2.9× for clarity).

50 better predictions for experimental gap responses than a potential flow model, especially when the gap width
51 was small. Whilst most previous CFD studies have focused on regular wave tests, transient wave groups
52 of moderate spectral bandwidth may be a better option for understanding the physics of gap resonance, as
53 multiple gap modes may be simultaneously excited in the gap, requiring much less computational cost than
54 regular wave excitations. Wang et al. (2018) and Wang et al. (2019) employed a NS solver to simulate 3D
55 gap resonance under the excitation of transient wave groups for two elongated boxes with round corners.
56 The numerical results compare favourably with the gap response time series of the free-surface elevation in
57 the gap, including higher harmonics, measured experimentally by Zhao et al. (2017), when both wave and
58 laminar boundary layer processes are resolved in the same simulation.
59 In light of the above, we carried out a series of experiments on 3D gap resonance for two simple boxes
60 with square corners separated by an appropriate gap width for FLNG side-by-side offloading (i.e. 4 m at
61 full-scale), extending the experiments of Zhao et al. (2017). The gap resonance is excited by NewWave-type
62 transient focused wave groups incident from the beam and bow. CFD simulations based on a viscous NS
63 solver were performed to understand free-surface behaviour and internal flow structures taking advantage
64 of the high spatial and temporal resolution of the numerical results. To gain a better understanding of
65 nonlinear wave harmonics interacting with the gap system, the harmonic components of the wave field were
66 separated in the NWT using a phase-based harmonic separation method (Fitzgerald et al., 2014). It should
67 be noted that the aim of this study is to gain fundamental insights into the gap resonance problems, rather

3
Table 1: Incident wave group parameters. A represents the nominal maximum surface elevation and fp represents the spectral
peak frequency of incident wave groups.

Linear excitation Frequency-doubling excitation


Set
A B C D E F
wave direction beam beam beam beam beam head
A (mm) 50 70 35 50 100 100
fp (Hz) 0.953 0.953 0.514 0.514 0.514 0.514

68 than try to achieve the largest responses in the gap (larger gap responses can be achieved using continuous
69 excitations due to e.g. regular waves). The use of transient focused wave groups to excite 3D gap resonance
70 allows the analysis of damping coefficients with a pure decay of gap responses, and better understanding of
71 the modal structures and harmonics in the gap.
72 The organization of the paper is as follows: a brief introduction to the experimental set-up, followed by
73 the damping analysis of experimental gap responses is presented in Section 2. Numerical simulations of the
74 experimental gap responses in a 3D NWT is presented in Section 3. Based on the obtained numerical results,
75 wave fields including higher harmonics and internal flow structures are discussed in Section 4 and Section 5,
76 respectively. Finally, some conclusions are provided.

77 2. Experiments

78 2.1. Experimental set-up

79 A series of 3D gap resonance experiments with square corners were undertaken in the Deepwater Offshore
80 Basin at Shanghai Jiao Tong University. The wave basin has a plan area of 50 m × 40 m and a movable
81 floor allowing testing with flexible water depth from 0 m to 10 m; a water depth of 10 m was used in the
82 experiments. Hinged flap-type wave paddles are installed along two neighbouring sides of the basin and wave
83 absorbers are located opposite to the wave paddles to minimize the effects of reflected waves.
84 Figure 1 shows the definition sketch of the experimental set-up for beam sea conditions. The experimental
85 models have a length-scale of approximately λ = 1 : 60 compared to full-scale vessels. Two identical prismatic
86 boxes were fully restrained from movement in a side-by-side fashion (representative of an FLNG offloading
87 scenario), forming a narrow gap with a width of G = 0.067 m (or 4 m at full-scale). One of the models used
88 in the experiments is shown in Figure 2; bespoke strips were stuck on to the rounded corners of the boxes
89 (which were used in Zhao et al. (2017)) to form square corners. The main dimensions of the boxes are (L,
90 B, H, D)=(3.333, 0.767, 0.425, 0.185) m, where L, B, H and D are the length, width, height and draft of
91 the models, respectively. The midpoint of the gap was 19.83 m away from the short-side wave paddles and
92 20 m away from the long-side walls.

4
Figure 2: One of the simple boxes used in the experiments.

93 The fluid in the gap was set in motion by NewWave-type transient wave groups, which are representative
94 of the average shape of the largest wave in a random sea (Tromans et al., 1991). These wave groups have
95 the following form:

N
A X
η(x, t) = S(fn ) · ∆f · cos[kn (x − x0 ) − ωn (t − t0 ) + θ] (1)
σ 2 n=1

96 where (x0 , t0 ) are the focus position and time, respectively, when the linear free-wave components come
97 into phase, η(x, t) is the free-surface elevation, A corresponds to the maximum surface elevation, kn and
98 ωn = 2πfn are the wave number and frequency of the nth wave component, S(f ) is the power spectral
PN
99 density of a given spectrum and σ 2 = n=1 S(fn ) · ∆f is the variance of the free-surface elevation.
100 Six sets of NewWave-type wave groups, referred to as Set A, B, C, D, E (beam seas, so broadside
101 excitation) and F (head seas, so end-on excitation), were generated both with and without the vessel models
102 in place, based on a Gaussian swell spectrum

Hs2 1 (f − fp )2
S(f ) = √ exp(− ) (2)
16 γ 2π 2γ 2
103 where Hs refers to the significant wave height, and fp is the spectral peak frequency with γ = 0.0775 Hz
104 the (laboratory scale) shape parameter of the Gaussian spectrum. The wave groups were defined to focus
105 linearly at a position 19.83 m away from the wave paddles, i.e. the gap centre for beam sea and mid-ship
106 position for head sea conditions. For each set, four NewWave groups with the same paddle signal envelope
107 but different relative phases θ of 0◦ , 90◦ , 180◦ and 270◦ (referred to as the ‘crest-focused’, ‘down-crossing’,
108 ‘trough-focused’, and ‘up-crossing’ wave groups, respectively) were carried out to facilitate separation of
109 different harmonics using the phase-based harmonic separation method (see Fitzgerald et al. (2014)). As
110 summarised in Table 1, two spectral peak frequencies with combinations of different linear nominal peak
111 wave amplitudes (defined pertaining to the input paddle signal) were used for generating the wave groups.
112 The frequency of the spectral energy peak for Set A and Set B corresponds to the first mode of the gap

5
50
Zhao et al. (2017)

Present

(a) Incident wave (w/o boxes)

-50
60
Zhao et al. (2017)

Present
(mm)

(b) Gap response (with boxes)

-60
-10 -5 0 5 10 15 20 25 30 35 40
t (s)

Figure 3: Comparison of the incident wave group and gap response measured at WG 4 (gap centre) between experimental
results in Zhao et al. (2017) and the present study.

113 resonance (fp = 0.953 Hz, referred to as ‘linear excitation’) whilst that for Set C, D, E and F is approximately
114 half of this value (fp = 0.514 Hz, this will be referred to as ‘frequency-doubling excitation’). It should be
115 noted that the waves created in the experiments were not perfectly focused symmetric NewWave groups due
116 to nonlinear evolution of the wave groups as they propagated down the wave basin, but this is not important
117 for the study here; all that required here is that the relative phasing of the four wave groups with the same
118 envelope is preserved.
119 Seven standard resistance wave gauges (WGs) were used to record the time-history of wave surface
120 elevation with a sampling frequency of 40 Hz, at positions along the gap shown in Figure 1. WG 1 and 7 are
121 1.216 m from the centre of the gap (0.45 m in from the ends), WG 2 and 6 are 0.833 m from the centre of the
122 gap, WG 3 and 5 are 0.5 m from the centre of the gap and WG 4 is central. Hence, for Set F, in head seas,
123 WG 1 to WG 7 are located further and further downward along the gap successively. In the experiments,
124 the wave generation system and data acquisition system used for the wave gauges were synchronised.
125 To check the repeatability of the experiments, the incident wave group and gap response measured at
126 WG 4 for Set A in the present study are compared with those obtained in Zhao et al. (2017)’s square corner
127 tests with the same incident wave parameters in Figure 3. The gap responses are in very good agreement
128 in general, with small discrepancies resulting from slight differences in the incident wave group and possibly
129 other minor changes in the experimental set-up (note that the experiments in Zhao et al. (2017) and the
130 experiments in the present study were conducted in two separate campaigns).

131 2.2. Gap harmonics

132 Using the four-phase harmonic separation method, the total free-surface elevation is separated into har-
133 monic components with their power spectra shown in Figure 4 for Set A and Set B, Figure 5 for Set C, Set

6
4 4
10 10
1st 1st
3 3
10 10
2nd 2nd

10
2 3rd 10
2 3rd
)

)
-1

-1
Hz

Hz
4th 4th
1 1
10 10
2

2
) (mm

) (mm
(a) Set A WG 4 (b) Set B WG 4

0 0
10 10
S( f

S( f
-1 -1
10 10

-2 -2
10 10

-3 -3
10 10
0 1 2 3 4 5 6 0 1 2 3 4 5 6
f (Hz) f (Hz)
Figure 4: Power spectra of free-surface gap harmonics for linear excitation in beam seas with two different amplitudes Set A
(A = 50 mm) and Set B (A = 70 mm). S(f ) is the power spectral density with respect to frequency f .

134 D and Set E, and Figure 6 for Set F. It is found that more harmonic components can be identified with
135 larger incident wave amplitude. Specifically, for Set A (A = 50 mm) the long-wave component, first and
136 second harmonics are separated, but the other harmonics are close to the overall background noise level. The
137 third harmonic can be observed for Set B (A = 70 mm); harmonic components up to the fourth order are
138 separated for Set C (A = 35 mm) and D (A = 50 mm), and those up to the seventh order can be observed
139 for Set E (A = 100 mm). Compared to the gap responses with round corners, less harmonics have been
140 observed here (Zhao et al. (2017) successfully extracted harmonics up to fourteenth order). With increasing
141 incident wave amplitude, the second harmonic responses can become larger than the first harmonic responses
142 (see Figure 5 (b) and (c)). For Set F (frequency-doubling and end-on excitation, A = 100 mm) harmonics
143 up to the eleventh order are found for WG 1 whilst harmonics up to the seventh order are observed for other
144 wave gauges.
145 The time series of the separated harmonic components for Set A and Set D are shown in Figure 7 and
146 Figure 8, respectively. For Set A, the first harmonic components recorded at symmetric positions show
147 good agreement. Similar to the gap responses with round corners (Zhao et al., 2017), the fluid in the gap
148 oscillates with a ‘beating pattern’ after the linear incident wave has passed the gap, which can be regarded
149 as a superposition of standing waves. The second-order sum harmonic and long-wave difference harmonic
150 (which shows the free-surface set-up) are observed whilst higher harmonics are negligible and not shown. It
151 is interesting to note that the local cancellation of the first harmonic (around t = 4 s) lines up in time with
152 the minimum of the long-wave difference harmonic and the maximum of the second-order sum harmonic,
153 which also occurs for Set B (see Figure 11).
154 For Set D the agreement of the second harmonics recorded at symmetric positions is generally good,
155 though some phase difference occurs after t = 25 s. The linear incident wave excites smaller responses
156 localised in time, but this is too low frequency to excite persistent responses in the gap. However, due to the
157 presence of the narrow gap, the response due to local frequency-doubled excitation can produce a vigorous

7
158 gap motion with a similar ‘beating pattern’ with very large amplitude.
159 If the incident wave group has a Stokes form, the second-order sum bound waves can be reconstructed
160 using the linear signal (Walker et al., 2004):

S 2 2
η2 = (η − η1H ) (3)
d 1
161 where S is related to the standard Stokes coefficients, d is the water depth, η1 and η1H are the linear
162 signal and its Hilbert transform, and η2 is the second-order bound wave signal. The top subfigure in Figure
163 8 compares the recorded second harmonic incident wave against the scaled theoretical second-order bound
164 wave component. For the recorded signal the first pulse around the focus time agrees well with the theoretical
165 result. However, additional pulses occur in the experimental data, but are absent from the theoretical result.
166 The first of these pulses represents a set of error waves, which occurs because the control signal applied at the
167 wave paddles is linear whilst the waves generated are inherently nonlinear. The second set of double frequency
168 components at ∼ 30 s presumably arises from reflection of these error waves from the tank boundary. Since
169 the error waves are free waves, not bound to the rest of the group, these interact with the boxes in an
170 essentially linear fashion – and if we construct a response amplitude operator (RAO) and phase response
171 function from the linear wave group Set A, with similar distribution of energy, the responses resulting from
172 the error waves can be then estimated for Set D. Figure 9 shows the maximum predicted gap response caused
173 by the error waves is smaller than 0.7 mm, demonstrating that the effect of error waves is negligible.

174 2.3. Damping analysis

175 As demonstrated for gap resonance with round corners in Zhao et al. (2017) and Wang et al. (2019),
176 the incident wave group excites free-surface modes in the gap during the initial wave excitation, after which
177 the ‘near-trapped’ wave energy carries out purely decaying oscillations due to the transient nature of the
178 incident focused wave group. If the damping affecting the oscillatory gap flow is approximately linear, the
179 ‘beating pattern’ of the decay can be well approximated using a series of decaying sinusoids with frequencies
180 corresponding to the gap resonant modes:

X
η(t) = Am sin(ωm t + θm )e−ωm ξm t (4)
m
ωm
181 where fm = 2π corresponds to the frequency of the mth mode, ξm the normalized modal damping coefficient,
182 and Am and θm the initial modal amplitude and phase of the chosen time window. Assuming the modal
183 frequencies and damping coefficients are independent of the positions along the gap whilst the amplitude
184 varies from position to position, a numerical fitting analysis based on the method of Kumaresan and Tufts
185 (1982) can be used to identify the modal parameters, which performs well in estimating the parameters
186 of exponentially damped sinusoids for short data records. Here the method is applied to analyse a 20 s
187 long time window of gap responses for square corners. The modal parameters for the linear harmonic gap

8
3 3
10 10
1st 1st
2 2
10 10 2nd
2nd

1 3rd 1 3rd
10 10
)

)
-1

-1
4th
Hz

Hz
4th
0 0
10 10 (b) Set D WG 4
2

2
) (mm

) (mm
(a) Set C WG 4

-1 -1
10 10
S( f

-2
10 S( f -2
10

-3 -3
10 10

-4 -4
10 10
0 1 2 3 4 0 1 2 3 4
f (Hz) f (Hz)
4
10
1st
3
10
2nd

10
2 3rd
)
-1
Hz

4th
1
10
2
) (mm

(c) Set E WG 4

0
10
S( f

-1
10

-2
10

-3
10
0 1 2 3 4
f (Hz)
Figure 5: Power spectra of free-surface gap harmonics for frequency-doubling excitation in beam seas with three different
amplitudes Set C (A = 35 mm), Set D (A = 50 mm) and Set E (A = 100 mm).

9
4 4
10 10
1st 1st
3 3
10 10
2nd 2nd

10
2 3rd 10
2 3rd
)

)
-1

-1
Hz

Hz
4th 4th
1 1
10 10
2

2
) (mm

) (mm
(a) Set F WG 1 (b) Set F WG 2

0 0
10 10
S( f

S( f
-1 -1
10 10

-2 -2
10 10

-3 -3
10 10
0 1 2 3 4 5 6 0 1 2 3 4
f (Hz) f (Hz)
4 4
10 10
1st 1st
3 3
10 10
2nd 2nd

10
2 3rd 10
2 3rd
)

)
-1

-1
Hz

Hz
4th 4th
1 1
10 10
2

2
) (mm

) (mm
(c) Set F WG 3 (d) Set F WG 4

0 0
10 10
S( f

S( f

-1 -1
10 10

-2 -2
10 10

-3 -3
10 10
0 1 2 3 4 0 1 2 3 4
f (Hz) f (Hz)
4 4
10 10
1st 1st
3 3
10 10
2nd 2nd

10
2 3rd 10
2 3rd
)

)
-1

-1
Hz

Hz

4th 4th
1 1
10 10
2

2
) (mm

) (mm

(e) Set F WG 5 (f) Set F WG 6

0 0
10 10
S( f

S( f

-1 -1
10 10

-2 -2
10 10

-3 -3
10 10
0 1 2 3 4 0 1 2 3 4
f (Hz) f (Hz)
4
10
1st
3
10
2nd

10
2 3rd
)
-1
Hz

4th
1
10
2
) (mm

(g) Set F WG 7

0
10
S( f

-1
10

-2
10

-3
10
0 1 2 3 4
f (Hz)
Figure 6: Power spectra of free-surface gap harmonics for frequency-doubling excitation in head seas Set F (A = 100 mm).

10
50 Inc 1st

-50
50 WG 1 1st

WG 7 1st
0

-50
50 WG 2 1st

WG 6 1st
0

-50
50 WG 3 1st

WG 5 1st
0

-50
50 WG 4 1st

-50
5 Inc 2nd

-5

4 WG 4 0th

-4
5 WG 4 2nd
(mm)

-5
-10 -5 0 5 10 15 20 25 30 35 40
t (s)

Figure 7: Time series of long-wave difference, first and second harmonic gap responses for Set A (linear excitation, beam sea,
A = 50 mm).

11
2 Inc 2nd

Inc scaled
0

-2
15 WG 1 2nd

WG 7 2nd
0

-15
15 WG 2 2nd

WG 6 2nd
0

-15
15 WG 3 2nd

WG 5 2nd
0

-15
15 WG 4 2nd

-15
0.5 WG 4 0th

0.0

-0.5
50 Inc 1st

-50
30
WG 4 1st

-30
1 WG 4 3rd

-1

0.5 WG 4 4th
(mm)

0.0

-0.5

-10 -5 0 5 10 15 20 25 30 35 40
t (s)

Figure 8: Time series of long-wave difference, first and second harmonic gap responses for Set D (frequency-doubling excitation,
beam sea, A = 50 mm). ‘Inc. scaled’ refers to the scaled theoretical second-order sum wave component computed based on the
linear incident wave component ‘Inc. 1st’.

12
1

(mm)
0

-1
-10 -5 0 5 10 15 20 25 30 35 40
t (s)

Figure 9: Reconstructed responses of error waves for Set D based on the linear transfer function of Set A.

Table 2: Frequencies f (Hz) and damping coefficients ξ × 103 for each mode of the gap resonant response calculated using
experimental results. Set ‘A’, ‘B’, ‘C’, ‘D’, ‘E’ correspond to the beam sea conditions listed in Table 1. ‘M’ stands for results
computed using the method of Molin et al. (2002).

Set f1 ξ1 f3 ξ3 f5 ξ5 f7 ξ7 f2 f4 R.M.S.
A 0.938 11.8 1.045 13.5 1.174 14.1 1.307 15.9 0.986 1.108 1.1%
B 0.938 11.6 1.044 13.6 1.173 14.3 1.309 16.8 0.987 – 1.2%
C 0.943 10.0 1.052 12.3 1.179 17.6 – – – – 3.3%
D 0.939 9.2 1.048 11.6 1.176 18.9 – – – – 3.9%
E 0.934 10.2 1.044 11.4 1.166 14.2 – – – – 4.1%
M 0.933 – 1.059 – 1.192 – 1.335 – 0.998 1.124 –

188 responses of Set A and Set B (starting from t = 10 s), and the second harmonic gap responses of Set C, Set
189 D and Set E (starting from t = 7 s) are listed in Table 2. The root-mean-square (R.M.S.) error is defined as
q
1X
n
(ηk − ηkf )2
R.M.S. = (5)
n max |ηk |
k=1

190 where ηk and ηkf represent the experimental and fitted gap responses from the k th wave gauge, respectively,
191 n is the number of wave gauges and the overbar symbol represents the averaged value over time. The method
192 of Molin et al. (2002) is also used to calculate the resonant frequencies, which gives a good prediction for
193 the first mode, but otherwise slightly over-predicts the other experimentally determined resonant frequencies
194 (see Table 2).
195 For the linear excitation cases (i.e., Set A and Set B), the fitted and experimental results agree well with
196 each other, with R.M.S. error values below 1.2%. In contrast to the round corner results in Zhao et al. (2017),
197 it is interesting to observe that anti-symmetric even modes (mode 2 and 4 for Set A; mode 2 for Set B) are
198 detected although their amplitudes are very small (e.g. the amplitudes of mode 2 and mode 4 are 10.1%
199 and 1.7%, respectively, of the amplitude of mode 1). This could result from anti-symmetric flow caused by
200 flow separation or slight asymmetry of the experimental set-up for square corners (e.g. slight twisting of the
201 model on its central support so the uni-directional wave crests are not exactly parallel to the gap).
202 For the linear excitation cases (Set A and Set B), the resonant frequencies and damping coefficients
203 for the odd modes are in excellent agreement with each other, and the RAO curves for Set A and Set B,

13
4
Set A

Set B
3

1/2
(SRes/SInc)
2

0
0.8 0.9 1.0 1.1 1.2 1.3 1.4 1.5 1.6 1.7

f (Hz)

Figure 10: Response amplitude operator of gap response measured at WG 4 for Set A and Set B.

204 obtained as the ratio of the response and incident wave spectra, match very well as shown in Figure 10. For
205 the frequency-doubling excitation cases (Set C, D and E), the resonant frequencies decrease slightly with
206 increasing incident wave amplitude. The damping coefficients vary slightly for mode 1 and mode 3 whilst
207 those for mode 5 experience larger change which may be attributed to the small amplitude of mode 5.
208 If the damping coefficient is linear and there is a generalised Stokes-type perturbation expansion for the
209 incident wave group, the time series of higher harmonic gap responses can be approximated well for an
210 incident wave group with the same peak spectral frequency but an arbitrary amplitude, by multiplying a
211 scaling factor to the power n of the Stokes expansion. The scaled harmonic structures of Set B are compared
212 with the harmonics of Set A in Figure 11, whilst the comparison for Set C, Set D and Set E is shown in Figure
213 12. It is found that the scaling works very well between Set A and Set B, and Set C and Set D. Generally,
214 all this evidence indicates that the overall effect of changing the incident wave amplitude is minimal for
215 these cases, and the damping affecting the gap flow is linear (or at least dominated by linear damping). For
216 Set E, some nonlinearity can be seen from the Stokes scaling of time series envelopes. This nonlinearity is
217 most obvious at the gap centre, where the response is largest along the gap and presumably with the most
ωηo2
218 significant vortex shedding. Here we note that the Reynolds numbers of these tests Re = ν range from
3 4
219 4.0 × 10 to 3.2 × 10 , where ω is the first resonant frequency, ν the kinematic viscosity and ηo the maximum
220 free-surface elevation at the centre of the gap, which are smaller than Re ≈ 105 corresponding to the first
221 appearance of turbulence for oscillatory flow over a smooth plane wall (Jensen et al., 1989). This makes it
222 plausible that laminar boundary layers dominate the majority of the vertically straight part of the gap away
223 from the corners.
224 Two caveats on obtaining the damping coefficients are discussed here. Firstly, the method of Kumaresan
225 and Tufts (1982) assumes linear superposition of modes, so no nonlinear interaction between the modes.
226 If a significant part of the damping is quadratic in velocity for the square corner case, this analysis may
227 not perform as well as in the round corner (linear) case reported in Zhao et al. (2017). Therefore, results
228 from this method must be treated with some caution, though the trends in frequencies and damping are
229 consistent with the time series. Secondly, if some of the damping is quadratic it means some of the drag

14
50
Set A Set B/1.2

(a) WG 4; Inc. 1st


-50
2
4 Set A Set B/1.2

-4 (b) WG 4; Res. 0th

50
Set A Set B/1.2

(c) WG 1; Res. 1st


-50
50
Set A Set B/1.2

(d) WG 2; Res. 1st


-50
50
Set A Set B/1.2

(e) WG 3; Res. 1st


-50
50
Set A Set B/1.2

(f) WG 4; Res. 1st


-50
5 2
Set A Set B/1.2
(mm)

(g) WG 4; Res. 2nd


-5
-10 -5 0 5 10 15 20 25 30 35 40
t (s)

Figure 11: Time series of harmonics for Set A compared with scaled envelopes (or time series) of harmonics for Set B. ‘Inc.’
and ‘Res.’ stand for ‘Incident wave’ and ‘Response’, respectively.

15
50
Set C Set D/1.4 Set E/2.7

(a) WG 4; Inc. 1st


-50
0.5 2 2
Set C Set D/1.4 Set E/2.7

0.0

(b) WG 4; Res. 0th


-0.5
10 Set C Set D/1.4
2
Set E/2.7
2

(c) WG 1; Res. 2nd


-10
10 Set C Set D/1.4
2
Set E/2.7
2

(d) WG 2; Res. 2nd


-10
10 Set C Set D/1.4
2
Set E/2.7
2

(e) WG 3; Res. 2nd


-10
2 2
10 Set C Set D/1.4 Set E/2.7
(mm)

(f) WG 4; Res. 2nd


-10

-10 -5 0 5 10 15 20 25 30 35 40
t (s)

Figure 12: Time series of harmonics for Set C compared with scaled envelopes (or time series) of harmonics for Set D and Set
E. ‘Inc.’ and ‘Res.’ stand for ‘Incident wave’ and ‘Response’, respectively.

230 force is proportional to the square of the amplitude of gap oscillation, so even in powers of amplitude, but
231 contributing odd harmonics in frequency when expanded into harmonic components (see Fitzgerald et al.
232 (2014)). This would damage the Stokes’ structure assumed to exist for the phase-based harmonic separation
233 method to work. Whilst this appears to affect the largest excitation case Set E, we see no evidence as such
234 for Set A, Set B, Set C and Set D.

235 3. Numerical simulations

236 3.1. Numerical methods

237 To reproduce the experiments, numerical simulations were carried out in a 3D NWT established using
238 OpenFOAM to solve the incompressible two-phase NS equations for the beam sea conditions Set A and Set
239 D and the head sea condition Set F. Wave generation of incident waves and absorption of reflected waves
240 were realised using waves2Foam (Jacobsen et al., 2012) with the free-surface captured using the volume-of-
241 fluid method. Second-order schemes were used for spatial discretization whilst a backward Euler scheme
242 was used for temporal discretization. The time step of the simulations was runtime adaptive to keep the
243 Courant-Friedrichs-Lewy number less than 0.5 for all the cells in the computational domain. The largest

16
50 Expt. crest Num. crest

Expt. trough Num. trough

(a) Set A
-50
50 Expt. crest Num. crest

Expt. trough Num. trough

(b) Set D
-50
100 Expt. crest Num. crest

Expt. trough Num. trough


(mm)

(c) Set F
-100
-10 -5 0 5 10 15 20
t (s)

Figure 13: Numerical wave re-creation of experimental crest- and trough-focused total incident wave groups for Set A, Set D
and Set F. ‘Expt.’ and ‘Num.’ refer to ‘Experimental’ and ‘Numerical’, respectively.

244 time step in the simulations is smaller than 1/100 of the peak wave period during wave propagation and
245 much smaller during wave-structure interactions as a result of the small near-wall mesh size.

246 3.2. Set-up of numerical wave tank

247 The origin of the Cartesian coordinate system in the NWT is positioned at the still water level at the
248 centre of the gap in the mid-ship plane, with x axis aligned with the wave propagation direction, z axis
249 pointing upward and y axis satisfying the right-hand rule. The lengths of the NWT were chosen as 8.0λp ,
250 5.2λp and 5.6λp for Set A, Set D and Set F, respectively (where λp of Set A is approximately a quarter of
251 that for Set D and Set F). The widths of the NWT were chosen to be sufficiently large, so 4L for Set A
252 and 12L (equal to the width of wave tank in the physical experiments) for Set D and Set F, to minimize
253 the influence of reflected waves from side-walls, whilst the water depths were chosen as λp for Set A and
254 0.58λp for Set D and Set F (which fulfil the deep-water conditions in the experiments). 200 cells per peak
255 wavelength and 50 cells per nominal incident wave height were used to generate the incident wave groups for
256 Set A and Set D, whilst 100 cells per nominal wave height were used for Set F. The relaxation zones used
257 for absorbing reflected waves at both ends of the NWT have lengths of 2.0λp , 1.5λp and 1.5λp for Set A, Set
258 D and Set F, respectively. Further details about the numerical set-up can be found in Wang et al. (2018)
259 and Wang et al. (2019).
260 The experimental crest- and trough-focused wave groups were re-created with excellent accuracy in the

17
50
Entire domain

Half domain

(mm)
0

-50
-5 0 5 10 15 20 25
t (s)

Figure 14: Comparison of gap responses measured at the gap centre using an entire domain and a half domain for Set A.

261 NWT using a linear iterative method (see Wang et al. (2018)) and the results are shown in Figure 13. The
262 re-created wave groups were then used to excite gap responses with fixed parallel boxes in the NWT. To
263 save computational cost, symmetry boundary conditions were applied in the mid-ship plane for beam sea
264 conditions and in the mid-gap plane for the head sea condition. To check the effect of this, a simulation using
265 an entire domain without symmetry boundary conditions was carried out for Set A and the gap responses
266 measured at the gap centre (WG 4) are compared to the half domain results shown in Figure 14. The results
267 are in good agreement demonstrating that the symmetry boundary condition works well.

268 3.3. Re-creating experiments

269 As demonstrated in Wang et al. (2019), the size of the near-wall mesh is key for accurate prediction of
270 gap responses for 3D gap problems with round corners. The numerical results show better agreement with
271 experimental data as the near-wall mesh is refined. Here we follow the same procedure for the square corner
272 cases. For Set A two meshes with near-wall mesh sizes of 2 mm and 0.4 mm, referred to as A1 and A2,
273 were used to carry out the simulations. Meshes A1 and A2 have 43.9 and 49.2 million cells, respectively, and
274 the simulations requires approximately 0.06 and 0.2 million CPU hours, respectively, computed using 720
275 cores on a Cray XC40 supercomputer with 2.6 GHz Intel Xeon E5-2690 v3 ‘Haswell’ CPUs. The obtained
276 numerical time series of gap responses are compared to the experimental data in Figure 15. It is found
277 that the time series of A1 and A2 are in good agreement, indicating that for these range of mesh sizes the
278 near-wall mesh size has insignificant effect on gap responses, in contrast to the round corner cases (Wang
279 et al., 2019). To further investigate the effect of near-wall mesh size on the decay of gap responses whilst
280 saving computational cost, a series of 2D decay tests were carried out by releasing a water column in a gap
281 formed by two fixed boxes. Details are given in Appendix A. Surprisingly, the near-wall mesh size plays a
282 minor role in determining the gap responses – a coarse mesh with first near-wall cell height of 7.5 mm agrees
283 well with a fine mesh with first near-wall cell height of 0.1 mm. Here note should be made that Kristiansen
284 and Faltinsen (2012) demonstrated convergence of gap responses by varying the number of uniform cells
285 across the gap and concluded that gap responses were not sensitive to mesh refinement in the gap for their
286 2D geometry with square corners. The horizontal mesh size across the gap in their study ranges from 6
287 mm to 22 mm for a gap with width of 90 mm. The mesh used in the present study is much finer. In light

18
Table 3: Frequencies f (Hz) and damping coefficients ξ × 103 for each mode of the gap resonant response calculated using
numerical results. ‘CFD’ and ‘DIFF’ refer to ‘Numerical results computed using CFD’ and ‘Numerical results computed using
DIFFRACT (Eatock Taylor and Chau, 1992; Sun et al., 2015)’, respectively. Set ‘A’ and ‘D’ correspond to the beam sea
conditions listed in Table 1.

Set f1 ξ1 f3 ξ3 f5 ξ5 f7 ξ7 R.M.S.
CFD A 0.957 11.8 1.057 12.2 1.180 10.1 1.138 10.3 1.9%
CFD D 0.960 9.1 1.058 9.7 1.183 11.6 – – 1.0%
DIFF A 0.965 3.4 1.064 4.2 1.187 5.8 1.323 6.8 0.1%

288 of the above, to efficiently simulate the 3D gap resonance problem and at least capture large-scale vortex
289 structures, mesh A1 is chosen as the benchmark mesh for the following study.
290 From Figure 15 the CFD results agree reasonably well with the experimental data for the first pulse
291 in the excitation stage (for t < 5 s); after that the CFD results start to lead in time with the difference
292 accumulating over time. At the end of the time window in Figure 15, the experimental and CFD results
293 are approximately 180 degree out of phase. For comparison, the gap response predicted using the linear
294 potential flow model DIFFRACT (Eatock Taylor and Chau, 1992; Sun et al., 2015) is included in Figure
295 15(d). It is found that the CFD results give a much better prediction compared to the potential flow results.
296 In terms of phase lead, ‘Potential flow model’ > ‘CFD model’ > ‘Experiment’. Likewise, the CFD results
297 of second harmonics for Set D start to lead the experimental results in phase after the initial excitation
298 (see Figure 16). Applying the method of Kumaresan and Tufts (1982) to the time series of numerical gap
299 responses, the resonant frequencies and damping coefficients for each mode are listed in Table 3. Here the
300 even modes are absent since the numerical simulations force the solution to be symmetric along the gap.
301 It is found that the resonant frequencies are over-predicted by CFD simulations for both Set A and Set D,
302 with a trend consistent with the phase difference. Compared to the CFD results for the round corner set-up
303 in Wang et al. (2019), the frequency difference in the first mode for the square corner set-up (2%) is clearly
304 larger than that for the round corner set-up (0.1%) whilst the frequency differences for the third and fifth
305 modes for both set-ups (approximately 1% and 0.5%, respectively) are similar. Given that the round corner
306 set-up has a much better agreement in phase, it is clear that the accumulating phase difference is mainly
307 caused by the frequency difference in the first mode.
308 The over-prediction of resonant frequencies might be attributed to the coupling of gap modes with the
309 unexpected 5 Hz component in the spectra of gap responses (see Figure 4) possibly caused by vertical
310 structural vibration of the bridge which supports the models in the experiments. This would have a bigger
311 effect on the first mode than the other higher modes, which is consistent with the mismatch in frequencies
312 for different modes. With such structural resonance the relative stiffness of the gap system will decrease by
313 (f /fr )2 , where f is the first modal frequency of the gap system and fr is the structural resonant frequency.
p
314 As a result, the first modal frequency of the gap system will decrease by 1 − 1 − (f /fr )2 ≈ 2%, which is

19
50
Expt. A1 A2

(a) WG 1

-50
50
Expt. A1 A2

(b) WG 2

-50
50
Expt. A1 A2

(c) WG 3

-50
70
DIFF Expt. A1 A2
(mm)

(d) WG 4

-70
-5 0 5 10 15 20 25
t (s)

Figure 15: Comparison of experimental and numerical crest-focused responses at various locations along the gap for Set A.
‘Expt.’ stands for experimental result; ‘A1’ and ‘A2’ are CFD results with different meshes; ‘DIFF’ represents potential flow
results calculated using DIFFRACT (Eatock Taylor and Chau, 1992; Sun et al., 2015).

20
315 consistent with the difference between the experimental and numerical results for the first modal frequency.
316 We note that the mounting arrangement for the boxes in these tests is slightly different to and less stiff than
317 that in the previous round-corner box tests Zhao et al. (2017) where this 5 Hz response on the wave gauges
318 was not observed. However, two things are worth noting here – (i) a similiar artefact of the 5 Hz component
319 was observed in the CFD results albeit with much smaller energy; (ii) a similar disagreement in frequency
320 was observed by Sauder et al. (2010) in their comparison of CFD results with the 2D experimental results
321 of Maisondieu et al. (2001) for decay of oscillation in a gap.
322 Despite the discrepancies in resonant frequencies, it is found the numerical damping coefficients in Table
323 3 generally agree well with the experimental results, especially for the first mode. Here we try to demonstrate
324 this by reconstructing the experimental time series by shifting the modal resonant frequencies and initial
325 phases to match those of the CFD results. The reconstructed results are shown in Figure 17 with starting
326 time chosen as t = 10 s and t = 7 s for Set A and Set D, respectively. It is found the agreement between
327 the CFD and reconstructed experimental results become much better when the small differences in resonant
328 frequencies are eliminated.
329 The CFD gap responses for the end-on frequency-doubling case Set F are shown in Figure 18. Generally,
330 the agreement between the numerical and experimental results is good. This enables us to examine the field
331 information in detail, by virtue of the high spatial and temporal resolution of the numerical simulations.

332 4. Free-surface behaviour

333 In this section, the simulation results of free-surface motions for broadside excitation, Set A (linear) and
334 Set D (frequency-doubling), and end-on excitation, Set F (frequency-doubling), are discussed. The harmonics
335 of the full wave field are successfully extracted utilising the two-phase harmonic separation method (or phase-
336 inversion method) of Fitzgerald et al. (2014), which is very difficult to achieve in a physical wave basin,
337 for better understanding the effects associated with different harmonic components during wave-structure
338 interaction.

339 4.1. Wave field

340 Based on the obtained numerical results, the full interaction of the wave groups with the boxes is shown in
341 supplementary Movies 1 and 2 for Set A and Set D, respectively. For Set A the incident wave group interacts
342 with the gap flow from both the wave propagating direction and the direction along the gap. Multiple modes
343 with different spatial structures are instantaneously excited when the incident waves pass the gap; these
344 resonant modes interfere and oscillate in the gap after the incident wave group passes the gap, slowly leaking
345 energy to the outside flow. Set D exhibits similar interaction patterns as Set A, but the excitation of gap
346 modes lags behind the incident wave group (which can also be identified by comparing Figure 13 to Figure
347 15 or Figure 16).

21
20
Expt. Num.

(a) WG 1

-20
20
Expt. Num.

(b) WG 2

-20
20
Expt. Num.

(c) WG 3

-20
20
Expt. Num.
(mm)

(d) WG 4

-20
-5 0 5 10 15 20 25
t (s)

Figure 16: Time series of second harmonic gap responses for Set D.

22
50
CFD

Reconstructed expt.

(a) Set A

-50
20
CFD

Reconstructed expt.
(mm)

(b) Set D

-20
-5 0 5 10 15 20 25 30
t (s)

Figure 17: Reconstructed experimental gap responses at the gap centre using CFD resonant frequencies and initial phases for
Set A and Set D.

348 For Set F the evolution of gap flow is shown in supplementary Movie 3 and snapshots captured at
349 different times are shown in Figure 19. In contrast to the beam sea conditions, the incident wave group
350 mainly interacts with the gap flow in the direction of wave propagation; the gap responses are gradually
351 excited as the incident waves travel downstream along the gap (for the beam conditions the gap responses are
352 almost simultaneously excited along the gap). Figure 19(a), (b) and (c) show the moments when the largest
353 incident wave event collides with the up-wave end of the box, arrives at the central focus position and passes
354 the down-wave end, respectively. Large free-surface motions are observed near the up-wave entry of the gap
355 (arising from the large run-up at the up-wave edge of the box as shown in Figure 19(a)), and decrease as
356 the incident wave group propagates down the gap (note how distinct the free-surface motions on either side
357 of the box are at the focus time). The trend here is similar to that for the head sea condition with round
358 corners, which was found to be consistent with potential flow theory (see Zhao et al. (2020)). Nonetheless,
359 the decrease of free-surface motion downstream along the gap is much more dramatic for the square corner
360 case considered here, which is caused by additional damping arising from vortex shedding from the up-wave
361 entry as well as the bottom of the box (see Section 5.2 for details). From Figure 19(c) and (d) it is found
362 that modes of different wavelengths (including obvious higher harmonics) are excited and superposed with
363 the incident wave group in the gap.

364 4.2. Harmonics of free-surface motion

365 For the frequency-doubling excitation cases Set D and Set F, both crest- and trough-focused wave groups
366 were used to excite gap resonance in the NWT. The phase-based harmonic separation method due to Fitzger-
367 ald et al. (2014) is employed to separate odd and even harmonics for each sampling point in the wave field,
368 which are then used to extract the harmonic component by digital filtering. Despite the difference in head-

23
100
WG 1 Expt.

WG 1 Num.

-100
100
WG 2 Expt.

WG 2 Num.

-100
100
WG 3 Expt.

WG 3 Num.

-100
100
WG 4 Expt.

WG 4 Num.

-100
100
WG 5 Expt.

WG 5 Num.

-100
100
WG 6 Expt.

WG 6 Num.

-100
100
WG 7 Expt.

WG 7 Num.
(mm)

-100
-10 -5 0 5 10 15 20 25 30 35
t (s)

Figure 18: Crest-focused responses measured at various locations along the gap for Set F.

24
(a) t = −0.6 s (b) t = 0 s

(c) t = 0.48 s (d) t = 4 s

-130 -65 0 65 130

Figure 19: Snapshots of numerically computed response wave field for the frequency-doubling end-on case Set F (in the region
−0.8L < x < 0.8L, −2B < y < 0, unit: mm). Note that the vertical axis is stretched, such that 1 unit in the vertical equals 15
units in the horizontal. The box is rendered as transparent to show more clearly the free-surface motions in the gap (indicated
using thin lines) compared with those outside the gap (indicated using thick lines). The incident waves propagate from left to
right. The thin and thick lines represent the free-surface motions in and outside of the gap, respectively.

25
369 ing, the patterns of different harmonics interacting with the gap system for the two numerically simulated
370 frequency-doubling excitation cases (Set D and Set F) have many aspects in common. Here the interaction
371 of long-wave component, first, second and third wave harmonics with the boxes are shown in supplementary
372 Movies 4-7 from a global view for Set D, whilst a local view is used in supplementary Movies 8-11 for Set F,
373 where all the harmonics are suitably scaled for better visualisation.
374 For the second-order difference long-wave component of Set D, a significant local mound of water is
375 observed near the leading end of the up-wave box, which slowly builds up and persists as long as the
376 incoming wave envelope. The mound arrives at its crest around the focus time (see Figure 20(a)) and
377 thereafter decreases slowly to the mean free-surface level, whilst there is no associated trough.
378 For the linear harmonic, since the linear frequencies are well below the resonant frequencies no modal
379 free-surface motions are excited in the gap. The diffracted wave field outside the gap is similar to that for
380 a single large box of width (2B + G). The linear incoming waves within the box region (approximately
381 −L/2 < y < 0) excite the gap faster than the waves outside the box region (approximately y < −L/2). The
382 resultant phase difference of the waves allows interaction of the incident linear waves outside the box with
383 the gap flow from the direction along the gap, which is apparent near the gap ends (see Figure 20(b)).
384 Large second-order sum harmonic waves are induced when the incident waves collide with the up-wave
385 box, which are then reflected back up-wave. Due to the local wave-wave-structure interaction, significant
386 second-order sum dynamics are excited in the gap (see Figure 20(c)). It is noteworthy that the gap motion is
387 initially driven from the gap end (around t = 0 s). Over the next few cycles the free-surface motion near the
388 gap centre becomes increasingly large as waves propagate towards the gap centre as if there is a wave maker
389 near the gap end (roughly from t = 0 s to t = 5 s). Then the waves in the gap seem to cease propagating,
390 and the free-surface motion near the gap centre becomes smaller over the next few cycles (roughly from t = 5
391 s to t = 8 s), after which the gap motion is driven at the gap centre and starts propagating towards the gap
392 end. Later on, this process repeats itself leading to waves bouncing back and forth along the gap. Zhao et al.
393 (2020) demonstrated that in the gap the group velocity of waves around the lowest modal frequencies are
394 decreased compared to the free field, whilst the phase velocities increase, giving rise to the patterns observed
395 here.
396 The snapshot of the third harmonic captured at the same time as the second harmonic is shown in
397 Figure 20(d), where higher modes of shorter wavelengths are clearly excited in the gap. Similar to the
398 second harmonic, the gap motion for the third harmonic is initially driven from the gap end, and waves
399 bouncing along the gap are present, which are visually easier to identify due to the shorter wavelengths of
400 the gap structures.
401 For all the harmonics, surface disturbances (wiggles) are present near the leading end of the up-wave box
402 presumably due to the presence of vortex shedding beneath.
403 The snapshots of the long-wave difference component, first, second and third harmonics for Set F are
404 shown in Figure 21. The free-surface displacements are generally smooth except at the gap ends. Similar

26
1.3

0.7

-0.7

-1.3

(a) Long-wave component, t = −0.2 s

54.2

27.5

-27.5

-54.2

(b) First harmonic, t = 0 s

10.0

5.1

-5.1

-10.0

(c) Second harmonic, t = 1.32 s

2.0

1.0

-1.0

-2.0

(d) Third harmonic, t = 1.32 s

Figure 20: Snapshots of numerically computed harmonics for the beam sea frequency-doubling case Set D (in the region
−4.5B < x < 3.2B, −0.9L < y < 0, unit: mm). Note that the vertical axis is stretched, such that 1 unit in the vertical equals
250, 7.5, 30 and 135 units in the horizontal for the long-wave component, first, second and third harmonics, respectively. The
incident waves propagate from left to right.

27
-3.8 -1.9 0 1.9 3.8 -88.4 -44.8 0 44.8 88.4
(a) Long-wave component, t = −0.2 s (b) First harmonic, t = 0 s

-18.4 -9.3 0 9.3 18.4 -8.6 -4.4 0 4.4 8.6


(c) Second harmonic, t = 2.68 s (d) Third harmonic, t = 2.68 s

Figure 21: Snapshots of numerically computed harmonics for the frequency-doubling end-on case Set F (in the region −0.8L <
x < 0.8L, −2B < y < 0, unit: mm). Note that the vertical axis is stretched, such that 1 unit in the vertical equals 300, 15, 60
and 150 units in the horizontal for the long-wave component, first, second and third harmonics, respectively. Other details as
Figure 19.

28
405 to the beam sea condition, a significant local mound of water slowly climbs up the front flat end of the two
406 boxes with no associated trough for the long-wave component. There is a smooth variation along the gap
407 with little vertical displacement at the down-wave end. The linear wave field is very smooth in space and
408 the gap motion follows the linear part of the incoming wave group though with smaller amplitudes (e.g. see
409 Figure 21(b) at the focus time). The second harmonic is the dominant response of the gap where significant
410 gap dynamics are present. The wavelength of motion within the gap is approximately twice that of the
411 waves outside (which are mostly second-order sum bound waves), and the phase speed of waves in the gap
412 is visually faster than the bound waves outside. Trapped waves are present in the gap – initially the gap
413 motion is clearly driven from the up-wave end; only at later times after the incident wave group has passed
414 the gap do the gap motions move upstream. This is consistent with the observation from the experiments
415 of Zhao et al. (2020). The wave propagation direction alternates in the gap and this process perpetuates
416 until the gap motion dies out. For the third harmonic the gap structure is reasonably smooth (note that
417 the free-surface is scaled up by a factor of 150). Persistent standing waves shorter in wavelength than the
418 second gap harmonic are clearest at later times, and waves are visually easier to identify.

419 4.3. Maximum free-surface motion


420 The maximum amplitude of free-surface elevation of each wave field point near the box region is plotted
421 in Figure 22, which clearly shows the pattern of wave-structure interaction. Much wave energy reflects off
422 the up-wave side when the incident wave group interacts with the structures whilst much less wave energy
423 is transmitted beyond the down-wave box. The maximum wave elevation occurs near the up-wave side
424 of the box whilst at different locations. Specifically, for Set A the maximum free-surface elevation takes
425 place at (x, y) = (−1.04B, −0.70B), i.e. at the up-wave edge of the box and 1.16 m away from the mid-
426 ship position; for the frequency-doubling case Set D at (x, y) = (−1.17B, 0), i.e. at the mid-ship plane
427 and 0.1 m away from the up-wave edge of the box; for the frequency-doubling and end-on case Set F at
428 (x, y) = (−0.55L, 0.54B), i.e. at the longitudinal centreline and 0.016 m away from the up-wave side of the
429 box. It is found that the wave-structure interaction is largest for Set A due to its smallest wavelength among
430 the three cases, whilst smallest for Set F due to its largest wavelength and smaller cross section perpendicular
431 to the wave propagation direction. The ratio of maximum free-surface elevation to the linear nominal peak
432 wave amplitude A in the simulations is largest for Set A with a value of 2.26, whilst those for Set D and Set
433 F have smaller values of 1.67 and 1.40, respectively.
434 The maximum responses along the gap are plotted in Figure 23, where the gap responses are normalised by
435 the linear incident wave amplitude A in the numerical simulations, and X/L = 0 and X/L = ±0.5 represent
436 the gap centre (mid-ship position) and gap ends, respectively; for Set F, X/L = −0.5 and X/L = 0.5
437 represent the up-wave and down-wave gap ends, respectively. During the excitation stage, the shape of free-
438 surface motions in the gap are essentially a superposition of the linear waves and excited modal motions. The
439 linear incoming waves are flat along the majority of the gap, with diffracted waves affecting the motions of the
440 free-surface close to the gap ends; whilst the resonant modes have shapes spanning different numbers of half

29
105

70

35

0
(a) Set A

84

56

28

0
(b) Set B

141

94

47

0
(c) Set C

Figure 22: Plan view of maximum of the total response wave field over all times (in the region −0.95L < x < 0.80L, −6.5B <
y < 0, unit: mm). The incident waves propagate from left to right.

30
441 wavelengths along the gap. For the linear excitation beam sea case Set A, the incident wave group rapidly
442 excites the first mode (which is dominant) as it passes the gap. As a result, the maximum gap responses
443 are determined by the linear travelling wave group and the shape of the first gap mode (which spans half a
444 wavelength along the gap, with one antinode at the gap centre and two nodes near the gap ends). Hence, the
445 maximum gap response decreases from the largest point at the mid-ship position to around X/L = 0.25 (see
446 Figure 23(a)). For the beam sea frequency-doubling case Set D, the second harmonic excitation excites the
447 first gap mode after the linear incident wave group passes the gap. Consequently, the maximum gap responses
448 mainly follow the linear waves which level out from the mid-ship position to around X/L = 0.25 (see Figure
449 23(b)). For both cases linear waves interact with the gap flow near the gap ends (see e.g. Figure 20(b)),
450 inducing larger maximum gap responses near the gap ends close to the amplitude of linear incident waves.
451 Based on the discussion above and assuming the linear waves in the gap are approximately the same for Set
452 A and Set D, the difference of maximum response along the gap between these cases approximately reflects
453 the shape of the first gap mode (also affected by the third and fifth mode shapes), which is demonstrated in
454 Figure 24.
455 Similar to Set D, the higher resonant harmonics for Set F are gradually excited after the main incident
456 group passes the gap. As a result, the maximum response along the gap shown in Figure 23(c) is consistent
457 with the linear travelling wave group, where the maximum gap response is largest at the up-wave entry
458 and gradually decreases towards the gap end in general. The slight increase of maximum response near the
459 down-wave exit is caused by the wave group touching the downstream end of the gap from outside. Of course,
460 the incident wave groups were not designed to maximise the response in the gap, so maximum responses in
461 random wave fields would be different.

462 5. Internal flow structures

463 In this section, the velocity field in the gap is investigated first for Set A and Set F. For comparison the
464 results for a counterpart of Set A but with round corners (here referred to as ‘Set AR’, which corresponds to
465 ‘Case A’ in Wang et al. (2019)) are shown as well. The bulk velocity field in the gap leads to the formation
466 of boundary layers and vortex shedding from sharp corners, which are then discussed via interrogation of
467 internal flow field information.

468 5.1. Velocity field

469 The vertical and transverse velocities below the free-surface in the gap are recorded along the gap centre-
470 line at z = −0.43D for Set A and Set AR, and z = −0.65D for Set F. The trace of velocity vectors, obtained
471 from the vertical and transverse velocities with respect to each sampling point are plotted during a nominal
472 cycle (note that these are not particle paths for passive tracer particles on the free-surface) in Figure 25 to
473 Figure 27.

31
1.0

0.5

(a) Set A
0.0
1.0

0.5

(b) Set D
0.0
1.5

1.0
'

0.5

(c) Set F
0.0
-0.50 -0.25 0.00 0.25 0.50
End Centre End

X / L

Figure 23: Variation of maximum response along the gap for (a) Set A, linear broadside excitation, (b) Set D, frequency-doubling
broadside excitation and (c) Set F, frequency-doubling end-on excitation. η 0 is the maximum response at each location along
the gap on the gap centreline over all times normalised by the nominal incident wave amplitude A; X/L is the normalised
distance away from the gap centre.

0.6

0.3
'

0.0

-0.50 -0.25 0.00 0.25 0.50


End Centre End
X / L

Figure 24: The difference of maximum response along the gap between Set A and Set D. Other details as Figure 23.

32
0.3
(a)
0.0

-0.3
-0.50 -0.25 0 0.25 0.50

0.3
(m/s)

(b)

0.0
U

-0.3
-0.50 -0.25 0 0.25 0.50

End Centre End

X/L
Figure 25: Trace of velocity vectors along the vertical centre plane of the gap over a nominal cycle for Set A (a) t ∈ [1.1, 2.3] s
and (b) t ∈ [9.7, 10.8] s. X/L as Figure 23. The circular symbols represent the positions of sampling points along the gap and the
horizontal and vertical displacements away from the sampling positions represent horizontal and vertical velocity components,
respectively. To avoid misinterpretation caused by figure scaling, the same length scale is used for units of velocity along each
axis.

0.5
(a)

0.0

-0.5
-0.50 -0.25 0 0.25 0.50
0.5
(b)
(m/s)

0.0
U

-0.5
-0.50 -0.25 0 0.25 0.50

End Centre End

X/L
Figure 26: Trace of velocity vectors along the vertical centre plane of the gap over a nominal cycle for the round corner
counterpart of Set A (a) t ∈ [1.4, 2.4] s and (b) t ∈ [11.2, 12.2] s. Other details as Figure 25.

474 For the beam sea condition Set A, it is found the vertical velocity generally dominates the transverse
475 velocity in the gap in both the excitation and decay stages (see Figure 25(a) and (b), respectively). The
476 vertical velocity is largest near the mid-ship position in the decay stage since all mode shapes for vertical
477 velocity have an antinode at the mid-ship position. For Set AR with round corners shown in Figure 26, the
478 vertical velocities in the gap are generally larger than those for Set A. The velocity vectors in the excitation
479 and decay stages show similar patterns with Set A, whilst the spatial variation is smoother since the effect
480 of vortex shedding is negligible.
481 For the head sea condition Set F, it is found the amplitude of velocity is largest near the up-wave gap
482 end in the excitation stage (see Figure 27(a)), as a result of the strong wave-structure interactions. In
483 contrast to the beam sea scenarios, the vertical velocity decreases downstream along the gap and becomes
484 more comparable in amplitude to the horizontal velocity (so compatible with travelling wave packets rather
485 than standing modal patterns in the gap). In addition, it is noteworthy that the velocity in the gap drops
486 dramatically when the incident wave group has passed the gap (see Figure 27(b)).

33
0.8
(a)

0.0

-0.8
-0.50 -0.25 0 0.25 0.50
0.8
(b)
(m/s)

0.0
U

-0.8
-0.50 -0.25 0 0.25 0.50

Up-wave end Centre Down-wave end

X/L
Figure 27: Trace of velocity vectors along the vertical centre plane of the gap over a nominal cycle for Set F (a) t ∈ [−1.6, 0.6]
s and (b) t ∈ [6, 8] s. Other details as Figure 25.

487 5.2. Vortex structures

488 Here the vortex structures are examined through numerical flow visualization of the vorticity fields with
489 axes of rotation in the horizontal plane, with X−vorticity ωx and Y −vorticity ωy defined as:

∂uz ∂uy

ωx =
∂y ∂z
(6)
∂ux ∂uz
ωy = −
∂z ∂x
490 For the beam sea condition Set A, vortex structures rotating about the y axis (the direction along the
491 gap) are extracted based on an iso-value of 9.4 s−1 from the Y −vorticity field shown in supplementary Movie
492 12, where only half of the gap is displayed with the gap centre shown at the top of the movie picture. The
493 snapshots for the evolution of vortex structures in the gap are shown in Figure 28. Initially, pairs of vortices
494 are generated at the entry along the gap, forming long vortex tubes extending from the mid-ship plane to
495 the region close to the gap end. With broadside incident waves, the roll-up of the vortex sheets is initially
496 smooth and close to symmetric with respect to the gap centre, whilst only at the gap end is the field more
497 disordered, since it is affected by local wave diffraction. The system is symmetric at very early times, and
498 even at approximately t = 1 s (so one period after the main wave group arrives) the vortex tubes are still
499 close to symmetric (see Figure 28(a)). Thereafter, the vortex tubes start to develop structures along their
500 axes, after which the flow becomes increasingly disordered. The vortex tubes oscillate into and out of the
501 gap periodically following the vertical gap velocity, which squeezes them in the direction across the gap (see
502 Figure 28(b)). At the peak of the oscillation, multiple tubes are present in the gap. Some pairs of cores of
503 opposite sign are ejected from the gap and move away from the gap under their mutual interaction (these
504 disappear from picture once their vorticity falls below the limit iso-value of 9.4 s−1 due to diffusion). With
505 multiple modes of different spatial structures excited in the gap, the vertical velocity varies transversely

34
(a) (b)

(c) (d)
Figure 28: Vortex structures rotating about y axis near the gap region at various phases during a nominal cycle in the excitation
stage for Set A. The iso-values for the blue and red colours are +9.4 s−1 and −9.4 s−1 , respectively, where the signs represent
opposite directions of rotation. Incident waves propagate from left to right.

506 leading to different vortex structures along the gap shown in Figure 29. As a result, the vortex tubes start
507 to twist and finally break down into small chunks of vortices (see Figure 28(c) where the time corresponds
508 to that in Figure 29). The vortex structures shed from the bottom of the boxes become smaller in size as
509 the gap motions decay over time (see Figure 28(d)).
510 For the head sea condition Set F, the vortex structures rotating about the x axis (so now along the gap)
511 are extracted based on an iso-value of 9.4 s−1 from the X−vorticity field shown in supplementary Movie
512 13 from the down-wave view and Movie 14 from the up-wave view. Here the down-wave view is defined as
513 looking from the wave-maker in the direction of the box with the gap displayed on the right of picture, and
514 vice versa for the up-wave view. Both the main boundary vortex layers and the shed vortex sheets which roll
515 up into vortex tubes are clearly visible along the sides of the box. At very early times, the vortex structures
516 inside and outside the gap are visually almost the same but with opposite directions of rotation (see Figure
517 30(a) and (b)), meaning the gap effects are negligible. The tubes are initially smooth and long, with their
518 origin close to the front bottom corner of the box and shapes following the motion of incoming waves. As
519 expected, the incident wave troughs stretch the cores whilst the wave crests compress the cores. The vortex
520 tubes are unstable under axial compression, so twists occur in both the red and blue cores. By around
521 t = 2.5 s, these cores are disrupted and break up into disordered patches of vorticity (see Figure 30(c) and
522 (d)) and the vortex structures inside and outside of the gap become very different. There is considerable
523 vortex shedding and roll-up on the front face of the box as well.
524 The vortex structures rotating about the y axis (so with rotation about the direction of the undisturbed

35
(a) (b)

(c) (d)

Figure 29: Vorticity contour in cross sections extracted from various slices (a) y = 0 m – gap centre; (b) y = 0.5 m; (c) y = 0.833
m; (d) y = 1.2165 m captured at t = 7.7 s for Set A. The free-surface is indicated using a black solid line. Incident waves
propagate from left to right.

525 incident crest line direction) are extracted based on an iso-value of 9.4 s−1 from the Y −vorticity field shown
526 in supplementary Movies 15 and 16 for down-wave view and up-wave view, respectively. Initially, a smooth
527 red vortex tube of a flat shape is formed via roll-up of vortex sheet and shed from the leading end of the
528 box off the base. It is raised towards the free-surface (see Figure 31(a)) before a blue vortex tube of an ‘ω’
529 shape (see t = −1.96 s), which moves downward following the incident wave trough. As the main wave group
530 arrives at the leading end of the box, vortex lumps develop and complex vortical interactions are visible.
531 Initially, the end effects are the same inside and outside the gap at the same longitudinal position. But this
532 rapidly changes so that the vortex structures are very different between the inside and outside of the gap.
533 The majority of vortex structures are located near the up-wave end of the gap, which are mostly advected
534 from the front face of the box, whilst obvious vortex structures develop along the whole length of the outside
535 edge of the box.

536 6. Conclusions

537 The purpose of this work has been to investigate three-dimensional (3D) gap resonance problems between
538 two fixed parallel elongated boxes with square corners at laboratory scale through physical experiments and
539 numerical simulations. Transient focused wave groups incident from both beam sea and head sea directions

36
(a) Down-wave view (b) Up-wave view

(c) Down-wave view (d) Up-wave view

Figure 30: Vortex structures rotating about x axis at different times for the head sea condition Set F. The colours are the same
as Figure 28. The free-surface is indicated using a transparent grey plane. ‘Down-wave view’ is looking from the wave-maker
towards the boxes and ‘Up-wave view’ is looking from the wave absorber towards the boxes.

540 are used in physical experiments to excite 3D gap resonance, both linearly and nonlinearly. The damping
541 coefficients are found to be almost the same for the most significant first and third modes of response, despite
542 different incident wave amplitudes. In addition, scaling consistent with a Stokes expansion works very well
543 for cases with different incident wave amplitudes except for the largest excitation case (with incident wave
544 amplitude of 100 mm and maximum gap oscillation of 63 mm). This evidence implies that the damping
545 affecting the oscillatory gap flow is dominated by linear damping for most of the cases studied here even
546 though significant vortex shedding is expected.
547 The physical experiments are re-created in a 3D numerical wave tank (NWT) with reasonable agreement,
548 both in terms of free-surface time histories and damping, which enables further investigation of numerical
549 field information. The only clear discrepancy between the NWT and experiments is an over-prediction of
550 the first modal resonant frequency, which might be caused by a vertical structural resonance of the model
551 and support bridge interacting with the hydrodynamic gap mode. In contrast to the 3D gap problem with
552 round corners in Wang et al. (2019), the near-wall mesh size for square corners has little influence on the gap
553 responses. This appears to imply that the dissipation due to vortex shedding dominates that of wall friction,
554 though wave radiation damping is always present. Possibly the effect of the vortex shedding could have a
555 linear form within the relatively low Reynolds number flow regime studied in this work. Here note should be
556 made of the experimental observation from Huse and Utnes (1994), who carried out a series of heave decay
557 tests for a column with a sharp edge at the lower end (representative of a tension-leg platform). They found
558 that the heave damping was purely linear up to 50 mm of initial displacement away from equilibrium, and

37
(a) Down-wave view (b) Down-wave view

(c) Down-wave view (d) Up-wave view

Figure 31: Vortex structures rotating about y axis at different times for the head sea condition Set F. Other details as Figure
30.

559 the linear damping still dominated the nonlinear damping when the initial displacement was increased to 100
560 mm (the ratio of contribution between linear and nonlinear damping is approximately 2). The column used
561 in their study has a diameter of 625 mm, which is not too dissimilar to the width of the boxes considered
562 here.
563 The patterns of wave-structure interactions are investigated for both beam and head sea conditions under
564 the excitation of transient wave groups. The gap motions can be regarded as a superposition of the linear
565 travelling wave group and the excited modal motions. For linear excitation the incident wave group rapidly
566 excites the modal motions as it passes the gap, and the maximum response in the gap is determined by both
567 the linear travelling wave group and the resonant modal motions excited in the gap. For frequency-doubling
568 excitation the modal motions in the gap are excited after the incident wave group passes the gap, and the
569 maximum response in the gap is mainly determined by the linear part of the incoming wave group.
570 The harmonics are separated for the whole wave field, which is not feasible in a physical wave basin.
571 The phase-based harmonic separation method works very well in space – the free-surface elevation structure
572 captured matches well the idea of a Stokes-like expansion in frequency harmonics, even for an incident
573 compact (so broad-banded) group. Each harmonic component has its own interaction with the gap system.
574 Some robust features are found, including local mounds of water near the leading end of the up-wave box
575 for the long-wave component, modal structures of different wavelengths and near-trapped waves bouncing
576 along the gap for the second and third harmonics.
577 The bulk velocity field in the gap is examined, which is associated with the free-surface motions and

38
578 vortex structures in the gap. The vertical velocity is found to dominate the longitudinal horizontal velocity
579 for beam sea conditions, meaning that some aspects of the full 3D gap resonance model can be explored
580 using its 2D counterpart. For the head sea condition, the gap velocity is largest near the up-wave end and
581 decreases downstream within the gap, and it decays much faster than the beam sea conditions under the
582 excitation of transient wave groups.
583 The nature of 3D vortex processes for the sharp-edged boxes is investigated by interrogation of the
584 vorticity field, and the vortex structures for the head sea condition are shown for the first time. The axial
585 compression of the vortex tubes leads to instability and twisting followed by complete breakdown. For the
586 head sea condition, the vortex structures inside and outside the gap become very different soon after the
587 focus time.
588 Overall, numerical re-creation of high-quality experiments in high temporal and spatial resolution enables
589 exploration of wave field and internal flow structures in detail. The phase-based harmonic separation method
590 is useful in understanding the dynamics of wave-structure interactions.

591 Acknowledgement

592 This work was supported by the ARC Industrial Transformation Research Hub for Offshore Floating
593 Facilities which is funded by the Australian Research Council, Woodside Energy, Shell, Bureau Veritas and
594 Lloyd’s Register (Grant No. IH140100012). The support in resources provided by the Pawsey Supercomput-
595 ing Centre with funding from the Australian Government and the Government of Western Australia, and the
596 help from Shanghai Jiao Tong University with experiments are acknowledged. The first author is grateful
597 for the support from Australian Government RTP, APA and Shell-UWA scholarships, and the support from
598 A∗STAR Engineering Research Council under the Marine & Offshore Strategic Research Programme (Grant
599 No. 17219 00089). The authors acknowledge useful discussions with Professor B. Molin of Aix–Marseille
600 Université and Professor R. Eatock Taylor of the University of Oxford.

601 References

602 Chen, X., 2005. Hydrodynamic analysis for offshore LNG terminals, in: Proceedings of the 2nd International
603 Workshop on Applied Offshore Hydrodynamics, Rio de Janeiro, Brazil.

604 Chua, K.H., de Mello, P., Malta, E., Vieira, D., Watai, R., Ruggeri, F., Eatock Taylor, R., Nishimoto, K.,
605 Choo, Y.S., 2018. Irregular seas model experiments on side-by-side barges, in: The 28th International
606 Ocean and Polar Engineering Conference, June 10-15, Sapporo, Japan, ISOPE-I-18-420.

607 Eatock Taylor, R., Chau, F., 1992. Wave diffraction theory—some developments in linear and nonlinear
608 theory. Journal of Offshore Mechanics and Arctic Engineering 114, 185–194.

609 Faltinsen, O., Rognebakke, O., Timokha, A., 2007. Two-dimensional resonant piston-like sloshing in a
610 moonpool. Journal of Fluid Mechanics 575, 359–397.

39
611 Feng, X., Bai, W., Chen, X., Qian, L., Ma, Z., 2017. Numerical investigation of viscous effects on the gap
612 resonance between side-by-side barges. Ocean Engineering 145, 44–58.

613 Fitzgerald, C., Taylor, P.H., Eatock Taylor, R., Grice, J., Zang, J., 2014. Phase manipulation and the
614 harmonic components of ringing forces on a surface-piercing column. Proc. R. Soc. Lond. A 470 (2168),
615 20130847.

616 Huijsmans, R.H.M., Pinkster, J.A., De Wilde, J.J., 2001. Diffraction and radiation of waves around side-
617 by-side moored vessels, in: The Eleventh International Offshore and Polar Engineering Conference, June
618 17-22, Stavanger, Norway, ISOPE-I-01-061.

619 Huse, E., Utnes, T., 1994. Springing damping of tension leg platforms, in: Offshore Technology Conference,
620 May 2-5, Houston, Texas, OTC-7446-MS.

621 Jacobsen, N., Fuhrman, D., Fredsøe, J., 2012. A wave generation toolbox for the open-source CFD library:
622 OpenFoam R . Intl. J. Numer. Meth. Fluids 70, 1073–1088.

623 Jensen, B., Sumer, B., Fredsøe, J., 1989. Turbulent oscillatory boundary layers at high Reynolds numbers.
624 Journal of Fluid Mechanics 206, 265–297.

625 Jiang, S.C., Bai, W., Cong, P.W., Yan, B., 2019. Numerical investigation of wave forces on two side-by-side
626 non-identical boxes in close proximity under wave actions. Marine Structures 63, 16–44.

627 Jin, Y., Chai, S., Duffy, J., Chin, C., Bose, N., 2018. URANS predictions on the hydrodynamic interaction
628 of a conceptual FLNG-LNG offloading system in regular waves. Ocean Engineering 153, 363–386.

629 Kristiansen, T., Faltinsen, O., 2012. Gap resonance analyzed by a new domain-decomposition method
630 combining potential and viscous flow draft. Applied Ocean Research 34, 198–208.

631 Kristiansen, T., Faltinsen, O.M., 2009. Studies on resonant water motion between a ship and a fixed terminal
632 in shallow water. Journal of Offshore Mechanics and Arctic Engineering 131, 11.

633 Kumaresan, R., Tufts, D., 1982. Estimating the parameters of exponentially damped sinusoids and pole-zero
634 modeling in noise. IEEE Trans. Acoust., Speech Signal Process. 30, 833–840.

635 Lee, C.H., Newman, J., 2005. Computation of wave effects using the panel method. Chakrabarti, S.K. (Ed.),
636 Numerical Models in Fluid-Structure Interaction. WIT Press, Southampton, 211-251.

637 Lu, L., Teng, B., Cheng, L., Li, Y., 2008. Numerical simulation of hydrodynamic resonance in a narrow
638 gap between twin bodies subject to water waves, in: The Eighteenth International Offshore and Polar
639 Engineering Conference, July 6-11, Vancouver, Canada, ISOPE-I-08-344.

640 Maisondieu, C., Molin, B., Kimmoun, O., Gentaz, L., 2001. Simulation bidimensionnelles des écoulements
641 dans une baie de forage. etude des modes de résonance et des amortissements.

40
642 Molin, B., 2001. On the piston and sloshing modes in moonpools. Journal of Fluid Mechanics 430, 27–50.

643 Molin, B., Remy, F., Camhi, A., Ledoux, A., 2009. Experimental and numerical study of the gap resonances
644 in-between two rectangular barges, in: 13th Congress of the International Maritime Association of the
645 Mediterranean, Istanbul, Turkey. IMAM.

646 Molin, B., Remy, F., Kimmoun, O., Stassen, Y., 2002. Experimental study of the wave propagation and
647 decay in a channel through a rigid ice-sheet. Applied Ocean Research 24, 247–260.

648 Moradi, N., Zhou, T., Cheng, L., 2015. Effect of inlet configuration on wave resonance in the narrow gap of
649 two fixed bodies in close proximity. Ocean Engineering 103, 88–102.

650 Pauw, W.H., Huijsmans, R.H.M., Voogt, A., 2007. Advances in the hydrodynamics of side-by-side moored
651 vessels, in: Proceedings of the 26th International Conference on Offshore Mechanics and Arctic Engineer-
652 ing, June 10-15, San Diego, California, USA. pp. 597–603.

653 Perić, M., Swan, C., 2015. An experimental study of the wave excitation in the gap between two closely
654 spaced bodies, with implications for LNG offloading. Applied Ocean Research 51, 320–330.

655 Saitoh, T., Miao, G., Ishida, H., 2006. Theoretical analysis on appearance condition of fluid resonance in a
656 narrow gap between two modules of very large floating structure, in: Proceedings of the Third Asia-Pacific
657 Workshop on Marine Hydrodynamics, Shanghai, China, pp. 170–175.

658 Sauder, T., Kristiansen, T., Ostman, A., 2010. Validation of a numerical method for the study of piston-like
659 oscillations between a ship and a terminal, in: The Twentieth International Offshore and Polar Engineering
660 Conference, June 20-25, Beijing, China, ISOPE-I-10-312.

661 Sun, L., Eatock Taylor, R., Taylor, P.H., 2010. First-and second-order analysis of resonant waves between
662 adjacent barges. Journal of Fluids and Structures 26, 954–978.

663 Sun, L., Eatock Taylor, R., Taylor, P.H., 2015. Wave driven free surface motion in the gap between a tanker
664 and an FLNG barge. Applied Ocean Research 51, 331–349.

665 Tromans, P.S., Anaturk, A.R., Hagemeijer, P., 1991. A new model for the kinematics of large ocean waves-
666 application as a design wave, in: The First International Offshore and Polar Engineering Conference,
667 August 11-16, Edinburgh, The United Kingdom.

668 Walker, D., Taylor, P.H., Eatock Taylor, R., 2004. The shape of large surface waves on the open sea and the
669 Draupner new year wave. Applied Ocean Research 26, 73–83.

670 Wang, H., Draper, S., Zhao, W., Wolgamot, H.A., Cheng, L., 2018. Development of a computational fluid
671 dynamics model to simulate three-dimensional gap resonance driven by surface waves. Journal of Offshore
672 Mechanics and Arctic Engineering 140 (6), 061803.

41
673 Wang, H., Wolgamot, H.A., Draper, S., Zhao, W., Taylor, P.H., Cheng, L., 2019. Resolving wave and laminar
674 boundary layer scales for gap resonance problems. Journal of Fluid Mechanics 866, 759–775.

675 Zhao, W., Taylor, P.H., Wolgamot, H.A., Molin, B., Eatock Taylor, R., 2020. Group dynamics and wave
676 resonances in a narrow gap: modes and reduced group velocity. Journal of Fluid Mechanics 883, A22.
677 doi:10.1017/jfm.2019.879.

678 Zhao, W., Wolgamot, H.A., Taylor, P.H., Eatock Taylor, R., 2017. Gap resonance and higher harmonics
679 driven by focused transient wave groups. Journal of Fluid Mechanics 812, 905–939.

680 Appendix A

681 The pure oscillatory decay stage of gap free-surface motion in the mid-ship plane of the 3D gap resonance
682 problem excited by an incident transient focused wave group is similar to the decay of a water column with
683 an initial displacement away from its equilibrium position in two dimensions, though the former oscillates
684 with multiple resonant frequencies. To efficiently investigate the effect of near-wall mesh size on the free-
685 surface motions in the gap from CFD, 2D decay tests were carried out, in which the configuration is the
686 same as the cross-section in the mid-ship plane for the 3D gap problem with a predefined water column
687 height of 50 mm (chosen to approximately match the gap responses for the 3D problem). Four meshes were
688 used to carry out simulations with the first near-wall cell height chosen as 7.5 mm, 2mm, 0.4 mm and 0.1
689 mm (referred to as G1, G2, G3 and G4, respectively, with mesh details shown in Figure 32) and the time
690 series of free-surface elevation at the gap centre were compared in Figure 33(a). It is noteworthy that the
691 time series of free-surface elevation for all cases agree well, though significantly different resolution is used
692 for different meshes. An additional case with slip boundary conditions applied to the boxes was carried
693 out using mesh G1, whilst the time history of free-surface elevation shows no difference compared with the
694 non-slip counterpart (see Figure 33(b)). The contours of vorticity for both cases are shown in Figure 34.
695 With slip boundary conditions the boundary layers are not modelled, although vortex shedding still occurs
696 due to the presence of sharp corners. This may indicate that wall friction plays a much less significant role
697 than vortex shedding in determining gap responses for square corners.

42
(a) G1 (b) G2

(c) G3 (d) G4

Figure 32: Illustration of mesh details near the gap entrance for 2D decay tests.

50
G1 G2 G3 G4

(a)
-50
50
G1 non-slip G1 slip
(mm)

(b)
-50
0 5 10 15 20
t (s)

Figure 33: Time series of free-surface elevation measured at gap centre for 2D decay tests with different (a) meshes and (b)
boundary conditions.

(a) Mesh G1 non-slip (b) Mesh G1 slip

Figure 34: Comparison of vorticity contour between cases with non-slip and slip boundary conditions. The vorticity is normalised
by the ratio of maximum vertical velocity in the gap to gap width.

43

View publication stats

You might also like