Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Ocean Engineering 237 (2021) 109600

Contents lists available at ScienceDirect

Ocean Engineering
journal homepage: www.elsevier.com/locate/oceaneng

URANS analysis of a free-running destroyer sailing in irregular


stern-quartering waves at sea state 7
Andrea Serani a ,∗, Matteo Diez a , Frans van Walree b , Frederick Stern c
a
CNR-INM, National Research Council–Institute of Marine Engineering, Rome, Italy
b
MARIN, Maritime Research Institute Netherlands, Wageningen, The Netherlands
c
IIHR–The University of Iowa, Iowa City, USA

ARTICLE INFO ABSTRACT

Keywords: Computational fluid dynamics (CFD) simulations and stochastic validation of free-running 5415M in irregular
URANS stern-quartering sea state 7 and Froude number 0.33 are presented. Unsteady Reynolds-averaged Navier–Stokes
Free-running (URANS) computations are validated against experimental fluid dynamics (EFD) tests. EFD static stability,
Irregular waves
forward speed, wave direction and spectrum were set such that roll-motion resonance were induced, causing
Stochastic validation
large roll angles, significant deck-edge immersion, and static instability. Prerequisite calm-water validation
Dynamic mode decomposition
Data clustering
covers roll decay at zero speed and self-propulsion studies for propeller RPM. Subsequent CFD stochastic
validation is achieved by statistical assessment of EFD data and CFD results (expected value, standard deviation,
and probability distribution) for input waves and ship response: time series are assessed by their wave
energy-spectra, autocovariance (waves only), and block-bootstrap analysis (waves and motions); mean-crossing
amplitudes are studied by bootstrap method. Dynamic mode decomposition is used with EFD and CFD time
histories to unveil the underlying ship dynamics and the motion correlation. Overall, CFD results are in
agreement with EFD. Finally, a clustering approach is used to identify wave sequences causing large roll
angles. As proof of concept, the characteristics of these sequences are used for the deterministic reconstruction
of severe (large roll) and rare (capsize) events via regular wave computations.

1. Introduction Generation Intact Stability Criteria (SGISC, Petacco and Gualeni 2020)
requires the use of validated numerical simulation methods to in-
The prediction of the heavy weather seakeeping performance of vestigate intact stability of ships for their so-called Level 3 criteria.
naval ships is important from both an operational and safety point of Developments for IMO-SGISC purposes may benefit the naval shipbuild-
view. Heavy weather seakeeping of naval ships has traditionally been ing community and vice versa. So far relatively little work has been
investigated by means of scale model testing in large seakeeping basins.
published on the suitability of simulation methods for Level 3 stability
From a safety point of view a large number of conditions needs to be
criteria.
investigated. Furthermore, in order to reduce the statistical uncertainty
in the results a large number of wave encounters has to be met during Computational fluid dynamic (CFD) by URANS computations has
the tests, including severe and rare events. This makes scale model shown the capability of predicting parametric rolling, surf riding,
testing time consuming and expensive. broaching, and course keeping in regular and irregular waves for low
During the last decades low- to high-fidelity simulation methods to high sea states. Sadat-Hosseini et al. (2010) show CFD parametric
have been developed for investigating heavy weather seakeeping of rolling prediction, while Sadat-Hosseini et al. (2011) present CFD
ships. These methods range from blended seakeeping-maneuvering computations of surf riding, periodic motion, and broaching, both for
methods using linear or non-linear strip theory added mass, damp- the ONR Tumblehome (ONRT) in regular waves. Free-running CFD
ing and diffraction data bases via 3D panel methods using various computations in calm water and regular waves are shown for the ONRT
degrees of linearization to high-fidelity unsteady Reynolds-averaged and the 5415M models by Wang et al. (2017) and Toxopeus et al.
Navier–Stokes (URANS) equation solvers. (2018), respectively. Irregular waves free-running computations for
The use of numerical simulation methods has gained considerable
the 5415M at sea state 5 by URANS and potential flow computations
interest also from the civil shipbuilding industry. The IMO Second

∗ Corresponding author.
E-mail address: andrea.serani@cnr.it (A. Serani).
URL: https://www.inm.cnr.it (A. Serani).

https://doi.org/10.1016/j.oceaneng.2021.109600
Received 21 January 2021; Received in revised form 20 June 2021; Accepted 31 July 2021
Available online 10 August 2021
0029-8018/© 2021 Elsevier Ltd. All rights reserved.
A. Serani et al. Ocean Engineering 237 (2021) 109600

are presented in Sadat-Hosseini et al. (2015), while the performance


of a trimaran in short-crested head waves are presented by Zhang
et al. (2021). Deterministic validation methods by potential flow are
presented in van Walree et al. (2020). Stochastic validation procedures
are presented in Diez et al. (2018) and applied to CFD simulations of
the Delft catamaran in irregular wave at sea state 5.
The objective of the present work is to assess the capability of a
URANS solver to reproduce accurately the statistics of the incoming
waves and ship response of a free-running destroyer in stern-quartering
irregular waves at sea state 7. Statistics of interest are the standard devi-
ation and probability distribution of time series and the expected value
and probability distribution of mean-crossing amplitudes. Moreover,
the identification of conditions (e.g., wave sequences with their char-
acteristics) causing large roll angles (severe events) and capsize (rare
events) and the capability to recreate deterministically such events are
also addressed.
URANS studies are built on earlier research on the stochastic val-
idation of ship responses in irregular waves (Diez et al., 2018) and
the validation and identification of severe slamming events in irregular
waves (Judge et al., 2020). Namely, free-running CFD simulations of
the 5415M with appendages (skeg, twin split bilge keels, twin rudders
and rudder seats slanted outwards, shafts, and struts) are assessed
for course keeping in irregular stern-quartering waves following a
JONSWAP spectrum at target Froude number (Fr) equal to 0.33. The
nominal significant wave height is equal to 7 m, corresponding to
sea state 7 (high), according to the World Meteorological Organiza-
tion (WMO) definition. The statistics of ship motions, rudder angle,
immersion probes/water on deck are addressed, which was never done
before. Validation studies include roll decay at zero speed with variable
roll gyradius and self-propulsion in calm water to assess revolution
per minute (RPM) values versus EFD data. The stochastic validation of
Fig. 1. MARIN EFD model 7967 (van Walree and Visser, 2010).
free-running CFD is achieved by statistical assessment of EFD data (pro-
vided by the Maritime Research Institute of Netherlands, MARIN, van
Walree and Visser 2010) and CFD results of input waves and ship
response times series. These are achieved by wave energy spectrum, 2.1. Test conditions
autocovariance, and bootstrap analysis following Diez et al. (2018).
The experiments were performed by van Walree and Visser (2010)
Additionally, dynamic mode decomposition (DMD, Schmid 2010) is
and consisted of free-running tests in waves with measurement of
used with both EFD and CFD time histories to unveil the underlying
motions and water levels on the weather deck. It is noted that the
ship dynamics (Diez et al., 2021) and shed light on the differences
loading condition of the model did not satisfy naval stability criteria
between CFD and EFD. Finally, five zero-crossing waves statistics and
in order to achieve large roll amplitude motions in waves. The model
𝑘-means clustering (Lloyd, 1982) are used for the identification of was manufactured at MARIN of wood and appended with skeg, twin
wave sequences causing large roll angles for both EFD and CFD. In split bilge keels, twin rudders and rudder seats slanted outwards, shafts
the remainder of this paper, these sequences are referred to as causal with V-type brackets and struts, and counter-rotating twin propellers.
sequences. Finally, as a proof of concept, the characteristics of causal The rudder was of the spade type. The lateral area of the rudders was
sequences are used to define regular wave conditions, with the aim to 2 × 15.4 m2 , i.e. 2 × 1.8% of the lateral area of the vessel (𝐿𝑝𝑝 × 𝑇 ). The
achieve: (1) severe (large roll angles) and (2) rare events (capsizing) by propellers were fixed pitch type with the direction of rotation inward
increasing the wave amplitude. over the top.
The present study is conducted within the NATO Science and Tech- The model was self-propelled and kept on course by a proportional–
nology Organization, Applied Vehicle Technology (AVT) Research Task derivative (PD) controller actuating the rudder angle (𝛽) at time 𝑡 as
Group AVT-280 on the ‘‘Evaluation of Prediction Methods for Ship follows
Performance in Heavy Weather’’, where several prediction methods, d[𝜓target − 𝜓(𝑡)]
hull forms, and conditions were addressed. The test case in this paper 𝛽(𝑡) =𝐾𝑃 ,𝜓 [𝜓target − 𝜓(𝑡)] + 𝐾𝐷,𝜓 + 𝐾𝑃 ,𝑦 [𝑦target − 𝑦(𝑡)]
d𝑡
addresses the most severe conditions (sea state 7) among those studied
(1)
in AVT-280.
The PD controller coefficients are shown in Table 2. The sway motion
was included with a weak gain to keep the model near the center of
2. Test case
the basin. The free-running tests were performed with propeller RPM
fixed to the self-propulsion point of the model for the envisaged speed.
The hull form under investigation is the MARIN model 7967, which The self-propulsion point was determined with speed trail test in calm
is equivalent to 5415M. This is a geosim replica of the DTMB 5415 water.
model but with different appendages designed by MARIN. The DTMB The experiments were conducted in MARIN’s Seakeeping and Ma-
5415 is an open-to-public naval combatant hull geometry. Fig. 1 (top) neuvering Basin measuring 170 × 40 × 5 m in length, width, and depth,
shows the 1:35.48 scale model and Fig. 1 bottom shows the model respectively. The basin wave makers enabled generation of regular
during a test run. The main particulars of the ship are given in Table 1. and long/short crested irregular waves from any direction relative to

2
A. Serani et al. Ocean Engineering 237 (2021) 109600

Fig. 2. Scheme of EFD setup.

Table 1 from 2 to 7). The experiments were conducted in irregular long crested
5415M main particulars (full scale). waves, following a JONSWAP spectrum.
Description Symbol Unit Value The spectral shape was obtained from wave height measurements
Length perpendiculars 𝐿𝑝𝑝 m 142.00 with a static wave probe in the center of the basin. The spectrum 𝑆 as
Waterline beam 𝐵 m 19.056 a function of the angular frequency 𝜔 is defined as
Draft 𝑇 m 6.15
Displacement mass ∇ tonnes 8642.6 (𝜔−𝜔𝑝 )2
( 𝜔 )4 −
Longitudinal center of gravitya LCG m 72.00 𝛼𝑔 2 − 54 𝑝
𝑒
2𝜎 2 𝜔2
𝑝
Vertical center of gravity VCG m 8.453
𝑆(𝜔) = 𝑒 𝜔 𝛾 (2)
𝜔5
Metacentric height GM m 1.010
Inertia radius about x-axis 𝑘𝑥𝑥 m 7.78 with
Inertia radius about y-axis 𝑘𝑦𝑦 m 35.50 2𝜋
𝜔𝑝 = (3)
Inertia radius about z-axis 𝑘𝑧𝑧 m 35.50 𝑇𝑝
Natural roll period at zero speed 𝑇𝜙 s 16.61
Propeller diameter 𝐷 m 6.15 and
Pitch ratio at 0.7 R – – 0.865 {
𝜎 = 0.07 for 𝜔 ≤ 𝜔𝑝
Expanded blade area ratio – – 0.580 (4)
𝜎 = 0.09 for 𝜔 > 𝜔𝑝
a From the aft perpendicular.
where 𝑔 is gravity acceleration, 𝜔𝑝 is the peak angular frequency
associated with the spectral component with maximum wave energy,
Table 2
and 𝛼 = 0.0168 (in full scale), 𝛾 = 3.3, and 𝜎 are the spectral shape
5415M rudder control details (full scale).
parameters.
Description Symbol Unit Value
The measured signals include motions, velocities, and accelerations
Yaw proportional constant 𝐾𝑃 ,𝜓 deg/deg 3.000
for the ship six degrees of freedom (DoF) and the wave elevation along
Yaw derivative constant 𝐾𝐷,𝜓 deg/(deg/s) 17.870
Sway proportional constant 𝐾𝑃 ,𝑦 deg/m 0.282 the waterline and on the weather deck at nine locations. Two locations
Target yaw 𝜓target deg 0.000 are used here as shown in Fig. 2. In addition, four acoustic wave
Target sway 𝑦target m 0.000 probes (fixed to the main carriage, 𝑥-direction) were used to record the
Rudder maximum rate RR deg/s 16.800 incident wave elevation at locations surrounding the model. Details of
Rudder maximum angle 𝛽max deg 35.000
the location of wave elevation probes (for the model at rest) relative
to the longitudinal midship location (station 10, st 10) are shown in
Fig. 2.
the free-sailing model. Beaches on the opposite sides absorbed the The stochastic validation of CFD results is presented for test case
incoming waves. A main carriage (𝑥-direction) and a sub-carriage (𝑦- 217 (van Walree and Visser, 2010), which is selected as benchmark
direction) followed the model and provided the required power and for current study since (i) presents large roll angles, including one rare
data storage. No wind makers were included in the experiments. event (capsizing) and (ii) has the largest number of runs. The test case
The experimental data were made available by the Cooperative collects 7 EFD runs at nominal Fr = 0.33, with nominal peak period
Research Navies (CRN) group, see van Walree and Visser (2010). 𝑇𝑝 = 9.2 s and wave heading of 300 deg. It is close to the resonance
condition for the roll, as shown in Fig. 3. The nominal significant wave
2.2. EFD validation data set height 𝐻1∕3 is close to 6 m and lays in between sea states 6 (very rough)
and 7 (high), according to the WMO definition. Nevertheless, the actual
EFD data set includes a total of 16 test cases. Specifically, EFD test significant wave height observed in the experiments equals 7 m, which
cases differ in nominal significant wave height (7.5, 10.0, and 12.5 m), identifies a sea state equal to 7 and is used to select a desired set of CFD
modal period (9 and 10.5 s), wave heading (300 and 330 deg), nominal runs. A total of 130 encounter waves were recorded, with an average
speed (12, 18, and 24 kn), run length, and number of repetitions (runs, of 20 encounter waves per run. The average run length per run is close

3
A. Serani et al. Ocean Engineering 237 (2021) 109600

Fig. 3. Comparison of nominal, EFD, and CFD wave encounter frequency to roll natural Fig. 4. Propeller open water curve.
frequencies for VC 1 (U indicates velocity magnitude).

Table 3
Propeller body forces details (nondimensional values) for CFDShip-Iowa.
Symbol Starboard Portside
𝐽 0.75344 0.75344
𝑐𝑡0 0.41249 0.41249
𝑐𝑡1 −0.31135 −0.31135
𝑐𝑡2 −0.13264 −0.13264
𝑐𝑞0 0.05274 −0.05274
𝑐𝑞1 −0.02917 0.02917
𝑐𝑞2 −0.02251 0.02251
𝑟𝑝 ∕𝐿𝑝𝑝 0.02166 0.02166
𝑟ℎ ∕𝑟𝑝 0.16005 0.16005
𝑝1 ∕𝐿𝑝𝑝 𝑥 =-0.93223 𝑥 =-0.93223
𝑦 =-0.03275 𝑦 = 0.03275 Fig. 5. Validation case 1 EFD and CFD versus nominal JONSWAP spectrum (encounter
𝑧 =-0.03601 𝑧 =-0.03601 wave). Note that wave probe 2 is used for EFD and CFD analyses.
𝑝2 ∕𝐿𝑝𝑝 𝑥 =-0.94227 𝑥 =-0.94227
𝑦 =-0.03275 𝑦 = 0.03275
𝑧 =-0.03653 𝑧 =-0.03653
value of the level-set function, positive in the water and negative in the
air. The 6 DoF rigid body equations of motion are solved to calculate
linear and angular motions of the ship.
to 300 s (in full scale). The model scale data rate is equal to 198.8 Hz. A simplified non-interactive body-force model is used for the pro-
Severe roll motions of more than 50 deg were recorded. peller, which prescribes axisymmetric body force with axial and tangen-
tial components. The body force field is applied at the position occupied
3. CFD method by the propellers, described by a disk volume extended radially from
𝑟ℎ to 𝑟𝑝 and axially from 𝑝1 (𝑥, 𝑦, 𝑧) to 𝑝2 (𝑥, 𝑦, 𝑧) (which also define the
The code CFDShip-Iowa V4.5 (Huang et al., 2008) is used for the propeller axis), whose values are listed in Table 3. The propeller model
CFD computations. CFDShip-Iowa is an overset, block structured CFD requires the open water curves and advance coefficient (𝐽 ) as input
solver designed for ship applications using either an absolute or a and provides the torque and thrust forces. The open water curves (see
relative inertial non-orthogonal curvilinear coordinate system for arbi- Fig. 4) are defined as a second order polynomial fit of the experimental
trary moving control volumes. Turbulence models include blended 𝑘 − 𝐾𝑇 (𝐽 ) and 𝐾𝑄 (𝐽 ) curves, as follow
𝜀∕𝑘−𝜔 based isotropic and anisotropic RANS and DES approaches with
either integration to the wall or wall functions. A single-phase level-set 𝐾𝑇 (𝐽 ) = 𝑐𝑡0 + 𝑐𝑡1 𝐽 + 𝑐𝑡2 𝐽 2 (5)
method is used for free-surface capturing. Captive, semi-captive, and
𝐾𝑄 (𝐽 ) = 𝑐𝑞0 + 𝑐𝑞1 𝐽 + 𝑐𝑞2 𝐽 2 (6)
full 6 DoF capabilities for multi-objects with parent/child hierarchy are
available. The actual propeller or body-force propeller model can be The trust and torque coefficient for the polynomial fit are summarized
employed for propulsion. Proportional–integral–derivative (PID) con- in Table 3. The advance coefficient 𝐽 is based on the ship instantaneous
trollers are available to correct the error between a measured process velocity at the center of gravity and does not consider the actual
variable and a desired set-point by calculating and then outputting propeller inflow.
a corrective action that can adjust the process accordingly. Numer- The governing equations are discretized using finite difference
ical methods include advanced iterative solvers, higher order finite schemes on body-fitted curvilinear grids. The time derivatives in the
differences with the conservative formulation, PISO (pressure-implicit turbulence and momentum equations are discretized using second order
with splitting of operators) or projection methods for pressure–velocity finite Euler backward difference. Convection terms in the turbulence
coupling, and parallelization with MPI-based domain decomposition. and momentum equations are discretized with higher order upwind for-
Dynamic overset grids use SUGGAR (the Structured, Unstructured, and mula. The viscous terms in the momentum and turbulent equations are
Generalized overset Grid AssembleR code developed by Noack 2005) computed with similar considerations using a second order difference
to compute the domain connectivity. scheme. The PISO algorithm is solved using the PETSc toolkit. SUGGAR
is used to obtain the overset interpolation information. An MPI-based
3.1. Computational setup domain decomposition approach is used, where each decomposed block
is mapped to one processor.
For the current study, absolute inertial earth-fixed coordinates For boundary conditions, the no-slip condition is applied on the
are employed. The turbulence is computed by the isotropic Menter’s solid surfaces. The far field boundary conditions are imposed on the top
blended 𝑘−𝜀∕𝑘−𝜔 (BKW) model with shear stress transport (SST) using and bottom of the background. The irregular wave boundary conditions
no wall function. The location of the free surface is given by the ‘‘zero’’ are applied for inlet, outlet, and side surfaces of the domain. The

4
A. Serani et al. Ocean Engineering 237 (2021) 109600

Table 4
CFDShip-Iowa computational grids details (S/P: Starboard/Portside).
Description Parent/Child Grid size I ×J ×K
Boundary Layer S/P Parent 953,258 (×2) 173 × 51 × 106
Skeg Parent 95,038 61 × 38 × 41
Rudder Top S/P Parent 95,040 (×2) 99 × 32 × 30
Rudder Top-Collar S/P Parent 95,175 (×2) 45 × 47 × 45
Rudder In S/P Child 94,392 (×2) 57 × 36 × 46
Rudder Out S/P Child 94,392 (×2) 57 × 36 × 46
Rudder Refinement S/P Parent 95,040 (×2) 99 × 30 × 32
Bilge keels S/P Parent 282,744 (×2) 154 × 36 × 51
Shaft Root S/P Parent 95,220 (×2) 45 × 46 × 46
Shaft S/P Parent 188,370 (×2) 91 × 46 × 45
Shaft Collar S/P Parent 94,962 (×2) 38 × 51 × 49
Shaft Refinement S/P Parent 95,120 (×2) 41 × 58 × 40
Strut In S/P Parent 94,640 (×2) 52 × 35 × 52
Strut Out S/P Parent 94,640 (×2) 52 × 35 × 52
Wave Ref Stern Parent 866,880 43 × 180 × 112
Wave background Parent 19,600,000 (×2) 700 × 160× 175
Fig. 6. Detail of the boundary-layer computational grid. Total 44,871,864

irregular wave boundary is based on a linear superposition of 𝑁 = 80 grid verification for the bare hull was performed in earlier work (Serani
elemental wave components. The amplitude 𝜁𝑖 for each component is
et al., 2021), where grid convergence index uncertainties were found
derived from the spectrum as
smaller than 2% and 4% for resistance and the motions RMS (on

𝜁𝑖 = 2𝑆(𝜔𝑖 )𝛥𝜔𝑖 (7) average) in regular head waves. For the hull appended with bilge keels
only, Wilson et al. (2006) showed insensitivity to the grid (using grid
where 𝜔𝑖 is the wave angular frequency and 𝛥𝜔𝑖 is the frequency
refinement ratio of 20.25 and finest boundary layer and bilge keels grids
band around 𝜔𝑖 . The 𝜔𝑖 values for the components are selected evenly
of 2.3M points) with an uncertainty smaller than 1% associated to the
distributed within 0.41 and 1.47 rad/s for the current study as in the
EFD campaign, with a constant 𝛥𝜔 = 0.0135 rad/s. The phase for each roll angle in roll-decay analysis.
component is selected randomly. Nominal JONSWAP spectrum for the
encounter wave is shown in Fig. 5. 4. Stochastic validation methods
Free-running CFD simulations are conducted for the 5415M model
(model scale 𝐿𝑝𝑝 = 4 m) at Fr = 0.33 and Re = 6.62E+06, with constant
propeller RPM to mimic the experimental procedures. Self-propulsion The current work follows the approach proposed by Diez et al.
computations in calm water with speed controller are used to find the (2018) for irregular wave validation studies. Statistical assessment and
RPM providing the nominal Froude number. validation are studied for time series values of wave elevation 𝜁, ship
motions, rudder angle, ship 𝑥- and 𝑦-velocities, and immersion probes.
3.2. Domain, grids, and time step Assessment and validation methods include Fourier analysis for the
assessment of the wave energy spectrum moments, analysis of the
The computational domain is a rectangular cuboid defined by autocovariance (AC), and moving block bootstrap (MBB) for the statis-
−2.0 < 𝑥∕𝐿𝑝𝑝 < 1.1, −1.21 < 𝑦∕𝐿𝑝𝑝 < 1.21, −1.00 < 𝑧∕𝐿𝑝𝑝 < 0.25. tical uncertainty of expected value (EV), standard deviation (SD), and
The ship axis is aligned with the 𝑥-axis, with the bow at 𝑥∕𝐿𝑝𝑝 = 0 and quantiles of the wave elevation (primary variable). The statistical un-
the stern at 𝑥∕𝐿𝑝𝑝 = −1. The free surface at rest lies at 𝑧∕𝐿𝑝𝑝 = 0. certainty of wave amplitude (secondary variable) EV, SD, and quantile
The model is appended with the skeg, bilge keels, rudders, rudder is evaluated by standard bootstrap (Efron, 1981) method, assuming in-
seats, shafts, and shaft brackets as in the experimental model but not
dependent and identically distributed random variable. The combined
appended with actual propellers. The computational grids are over-
use of the kernel density estimator (Miecznikowski et al., 2010) with
set, with independent grids for the hull, appendages, refinement and
bootstrap methods provides validation values and uncertainties for the
background, and then assembled to generate the total grid. The hull
probability density function (PDF). The same methodologies are used to
(boundary layer) grids are generated with a hyperbolic grid generator
using a double O-topology, one each for the starboard and the port evaluate the statistical uncertainties associated with the ship motions,
sides. The bilge keels and skeg use H-topology grids and overset the rudder angle, ship 𝑥- and 𝑦-velocities, and immersion probes, where AC
boundary layer grids. Two double O-grids are used for each rudder such and MBB methods are used for primary variables and bootstrap is used
that inboard and outboard sides are conformal. O-topology grids are for secondary variables. Errors and confidence intervals of the statistical
used for the rudder seats, shafts, and shaft brackets. A Cartesian grid estimators are used to define validation criteria of CFD outputs versus
is used to impose the far-field boundary conditions and to resolve the EFD benchmark values.
flow far from the hull, with a refinement Cartesian block close to the
stern. The total number of grid points is about 45M. Details of the grids
4.1. Spectral moments
are shown in Table 4 and Fig. 6.
Simulation (nondimensional) time step is set equal to 0.004, which
corresponds to 0.008 s in model scale and a data rate equal to 129.2 Hz. The wave energy spectrum moments, 𝑚0 , 𝑚1 , and 𝑚2 are evaluated
At nominal speed with no motions, the ship travels one ship length in as
250 time steps and one wavelength corresponding to the modal period ∞
in 350 time steps. For the current test case, CFD simulation time was 𝑚𝑘 = 𝜔𝑘 𝑆(𝜔)d𝜔, 𝑘 = 0, 1, 2 (8)
∫0
approximately 1M CPU hours for a full-scale duration of 1 hour, on a
HPC system with 2.7 GHz Intel Xeon Platinum cores. where 𝑆(𝜔) = 𝑎2 (𝜔)∕2 and 𝑎 is the elemental wave component ampli-
Grid sensitivity analysis for the fully appended model is very time tude. Numerically, Eq. (8) is evaluated using discrete Fourier transform
consuming and beyond the scope of the present work. Nevertheless, of the time series.

5
A. Serani et al. Ocean Engineering 237 (2021) 109600

4.2. Statistics of interest and 𝑐 = (1−𝜑)(1+𝜑). From the original set of 𝐶 blocks, a number of 𝐶 ′ =
𝑀∕𝑙 blocks are drawn at random with replacement and concatenated
Statistics of interest are the expected value (EV), standard deviation in the order they are picked, forming a new bootstrapped series of size
(SD), cumulative distribution function (CDF), and probability density 𝑀. 𝑄 = 100 bootstrapped series are used here.
function (PDF). They are evaluated numerically using a sample of 𝑀 EV and SD are evaluated as per Eqs. (9) and (10). The CDF is
items. Items are extracted from time series as 𝜉𝑖 = 𝜉(𝑡𝑖 ), 𝑖 = 1, … , 𝑀. assessed by quantile function 𝑞, evaluated at probabilities 𝑝 = 0.025,
For secondary variables (amplitudes), 𝑀 mean-crossing waves are 0.1, 0.2, … , 0.8, 0.9, 0.975. Sorting all the 𝜉(𝑡𝑖 ) (𝑖 = 1, … , 𝑀) within
identified and the associated 𝜉𝑖 defined (note that the sample size is each bootstrapped series, such that 𝜉(𝑡𝑖−1 ) ≤ 𝜉(𝑡𝑖 ) ≤ 𝜉(𝑡𝑖+1 ), then
different for primary and secondary variables).
The statistics of interest are evaluated as follows: 𝑞(𝑝) = 𝜉⌈𝑝𝑀⌉ (19)

1 ∑
𝑀
The validation value for EV and its 95% confidence lower and upper
EV(𝜉) = 𝜉 (9)
𝑀 𝑖=1 𝑖 bounds are evaluated as


√ 1 ∑ 𝑀
[ ]2 EV = Median(EV𝑏 ) = EV⌈0.5𝑄⌉ (20)
SD(𝜉) = √ 𝜉 − EV(𝜉) (10)
𝑀 − 1 𝑖=1 𝑖 EV𝑙 = EV⌈0.025𝑄⌉ and EV𝑢 = EV⌈0.975𝑄⌉ (21)

1 ∑ [
𝑀
] where EV𝑏 represents the EV value of the 𝑞th bootstrapped series,
CDF(𝜉, 𝑦) = 𝛿 𝜉𝑖 ≤ 𝑦 (11)
𝑀 𝑖=1 ordered such as EV𝑏−1 ≤ EV𝑏 ≤ EV𝑏+1 . Finally

The PDF is here evaluated using kernel density estimation 𝑈EV = 0.5(EV𝑢 − EV𝑙 ) (22)
(Miecznikowski et al., 2010) as
[ ] The validation values and confidence interval for SD are similarly
1 ∑ 1
𝑀
(𝑦 − 𝜉𝑖 )2 evaluated.
PDF(𝜉, 𝑦) = √ exp − (12)
𝑀ℎ 𝑖=1 2𝜋 2ℎ2

where ℎ = SD(𝜉)𝑀 −1∕5 is a bandwidth (Silverman, 2018). 4.5. Bootstrap method

4.3. Autocovariance analysis The validation and confidence intervals of secondary variables are
evaluated by the bootstrap method (Efron, 1981). The method con-
The analysis of the AC is used to provide confidence intervals for structs an empirical probability function from the sample (assumed
EV and SD of primary variables (time series values). Validation values independent and identically distributed) assigning a probability of 1∕𝑁
are directly provided by Eqs. (9) and (10), whereas confidence intervals to each item. This represents the nonparametric maximum likelihood
are evaluated using the variance of the mean Var(EV) and the variance estimate of the actual distribution. The associated empirical CDF is
of the variance, Var(SD2 ) (Belenky et al., 2015), respectively defined as given by Eq. (11).
𝑀−𝑖 (
From the empirical CDF, it is calculated a number 𝐵 of new sets,
)
SD2 2 ∑ 𝑖 each containing new random items 𝜉𝑖⋆ (𝑖 = 1, … , 𝑁) defined as
Var(EV) = + 1− 𝑅(𝜏𝑖 ) (13)
𝑀 𝑀 𝑖=1 𝑀
𝑀−𝑖 ( ) 𝜉𝑖⋆ = CDF−1 (𝑝𝑖 ) (23)
2 2SD4 4 ∑ 𝑖 [ ]2
Var(SD ) = + 1− 𝑅(𝜏𝑖 ) (14)
𝑀 𝑀 𝑖=1 𝑀 where 𝑝𝑖 are 𝑁 uniformly distributed random items in [0, 1]. Ordering
the original items such that 𝜉𝑖−1 ≤ 𝜉𝑖 ≤ 𝜉𝑖+1 , the new items 𝜉𝑖⋆ are
where 𝑅(𝜏𝑖 ) is the autocovariance function, here used in the following
weighted form 𝜉𝑖⋆ = 𝐽[𝑝𝑖 𝑁] (24)

1 ∑
𝑀−𝑖
[ ][ ] Validation values and 95% confidence intervals for EV, SD, and
𝑅(𝜏𝑖 ) = 𝜉𝑗 − EV(𝜉) 𝜉𝑗+1 − EV(𝜉) (15)
𝑀 𝑗=1 quantiles are evaluated similarly to MBB method.
The statistical uncertainties associated to EV and SD are evaluated
at the 95% confidence interval assuming a normal distribution for the 4.6. Validation metrics
EV estimated as
√ Input irregular wave from both EFD and CFD is assessed and val-
𝑈EV = 2 Var(EV) (16)
idated versus nominal benchmark values from the spectrum. Specifi-
and by the central-limit theorem for the SD as cally, the error associated with EFD and CFD wave energy spectrum
√ moments is evaluated as
√ √

𝑈SD √ √ Var(SD2 ) 𝐸 = 𝑚(EFD or CFD)
− 𝑚(𝑛) (25)
= 1+2 −1 (17) 𝑘 𝑘
SD SD2
where superscripts 𝑛 indicates nominal values.
4.4. Moving block-bootstrap method For primary and secondary variables, the validation error for EV is
evaluated as
To define the validation values and confidence intervals of statistical 𝐸 = EV(CFD) − EV(EFD) (26)
estimators of Eqs. (9)–(12) a MBB method is applied to time series,
using a number 𝐶 = 𝑀 − 𝑙 + 1 of moving blocks, each formed by 𝜉𝑖 , and associated validation uncertainty is assessed as following
𝑖 = 𝑗, … , 𝑗 + 𝑙 − 1, where 𝑗 is the block index and 𝑙 = (2𝜑∕𝑐)2∕3 𝑀 1∕3 is √
[ ]2 [ ]2
(CFD) (EFD)
an optimal block length (Carlstein et al., 1986) with 𝑈𝑣 = 𝑈EV + 𝑈EV (27)
∑ [ ][ ]
𝑀 𝑀−1 𝑖=1 𝜉𝑖+1 − EV(𝜉) 𝜉𝑖 − EV(𝜉) Validation is achieved if |𝐸| ≤ 𝑈𝑣 . SD validation error and uncertainty
𝜑= ∑ [ ] (18)
2
(𝑀 − 1) 𝑀 𝑖=1 𝜉𝑖 − EV(𝜉)
are similarly assessed.

6
A. Serani et al. Ocean Engineering 237 (2021) 109600

Fig. 8. EFD versus CFD roll decay (first three periods).


Fig. 7. Wave sequence causing severe event concept.
Table 5
Roll decay CFD results in calm water at zero speed.

4.7. Dynamic mode decomposition Data 𝑘𝑥𝑥 ∕𝐵 [−] 𝐸% 𝑇 [𝑠] 𝐸% 𝛿 [−] 𝐸%


EFD 0.413 16.30 0.142
CFD 0.385 −6.8 16.27 0.3 0.128 10.1
DMD (Schmid, 2010) is applied to motion time histories as an
0.400 −3.1 16.81 3.1 0.121 15.0
additional approach to assess the underlying ship dynamics (Diez et al., 0.413 0.0 17.27 6.0 0.114 20.0
2021) and shed light onto similarities and differences between EFD and 0.428 3.6 17.87 9.6 0.109 23.5
CFD results. DMD is a data-driven and equation-free dimensionality- Note: 𝐸 is % of EFD.
reduction/reduced-order modeling method, which provides a linear
representation of a possibly nonlinear system dynamics by means of
a set of complex conjugate modes 𝝓𝑘 and frequencies 𝜔𝑘 . The resulting the elbow method (Ketchen and Shook, 1996) is used with the WCSS
reduced order model can be used for diagnostic, state estimation, metrics.
future-state prediction, and control (Kutz et al., 2016). Here, DMD is As additional clustering metrics, the silhouette is used. The silhou-
used to unveil motion correlation and compare EFD and CFD modes, ette method provides a metrics of consistency of data within clus-
frequencies, and modal participation. Details of assumptions and equa- ters (Rousseeuw, 1987). Assume 𝑎𝑖 as the average Euclidean distance
tions for DMD in general may be found in Kutz et al. (2016). Example between 𝐝𝑖 and any other data point within the cluster 𝐝𝑖 belongs to.
applications to ship maneuvering in waves are given in Diez et al. Assume then 𝑐𝑖 as the smallest average Euclidean distance of 𝐝𝑖 to all
(2021). data points in any other cluster 𝐝𝑖 does not belong to. The silhouette
associated to 𝐝𝑖 is defined as
5. Causal-sequence identification approach 𝑎𝑖 − 𝑐𝑖
𝑠𝑖 = (31)
max[𝑎𝑖 , 𝑐𝑖 ]
Severe events (SEs) are defined using an arbitrary threshold value
and is a measure of how similar the data point is to points in its own
𝜎, based on the roll amplitude SD, whereas capsizings are assumed
cluster as opposed to other clusters. It may be noted that 𝑠𝑖 ranges from
rare events. By definition, SEs occur when the roll amplitude is greater
−1 to 1, where 1 indicates maximum similarity. The average silhouette
than EV + 𝜎SD. Their analysis includes the assessment of the wave of all data points is used as a metrics for proper data clustering:
sequence causing the event. Specifically, 𝑍 = 5 zero-crossing waves
are here considered, as shown in Fig. 7. The assessment is based on ∑
𝐻
𝑠avg = 𝑠𝑖 (32)
the within the sequence EV and SD of wave height, amplitude, and 𝑖=1
encounter period, as well as on their weighted counterpart (EVw and
Note that for 𝑘 = 1 the silhouette is not defined. By convention, for
SDw ), respectively defined as
𝑘 = 1 it is 𝑠avg = 0.
∑𝑍
𝑤𝑖 𝜉𝑖
EVw (𝜉) = ∑𝑖=1
𝑍
(28) 6. Numerical results
𝑖=1 𝑤𝑖

√ 1 ∑𝑍 [ ]2
√ In this section numerical results are presented for the wave elevation
√ 𝑍−1 𝑖=1 𝜉𝑖 − EVw (𝜉)
SDw (𝜉) = √ (29)
1 ∑𝑍
at probe 2 (see Fig. 2), the ship motions (surge, sway, heave, roll,
𝑍 𝑖=1 𝑤𝑖 pitch, yaw), the rudder angle, 𝑥- and 𝑦-velocities, and immersion probes
where 𝑤𝑖 are the weights. R3 and R5, including the statistics of primary (time series values)
Additionally, the 𝑘-means clustering method (Lloyd, 1982) is used and secondary (amplitudes) variables. Errors and uncertainties for the
to identify the data cluster associated to SEs or at least large roll statistical estimators provided by bootstrap methods are normalized
angles. It allows to build partitions of the data in 𝑘 different sets with 2SD. Ship motions are observed in the carriage coordinate system
(clusters), defined by representative points (centroids). Here, wave and projected onto the ship axes. Unless differently specified, values
sequence parameters are used to form the desired data points 𝐝𝑗 . The are expressed for the full scale. Results are also presented for the DMD
of relevant variables and the analysis of wave sequences causing severe
assignment of data points to 𝑘 clusters is achieved by minimization of
events.
the within-cluster sum of squares (WCSS)

𝑘 ∑
6.1. Roll decay validation in calm water
WCSS = ‖𝐝𝑗 − 𝝁𝑖 ‖2 (30)
𝑖=1 𝐝𝑗 ∈𝐾𝑖
Fig. 8 shows 5415M roll decay at zero speed in calm water, com-
where 𝝁𝑖 are the cluster centroids, evaluated by averaging all data paring EFD and CFD results. In van Walree and Visser (2010), a 𝑘𝑥𝑥 =
points within the 𝑖th cluster 𝐾𝑖 . The WCSS in Eq. (30) is used as evalu- 0.413𝐵 value was determined by means of roll decay tests in the MARIN
ation metrics to identify the optimal number of clusters 𝑘. Specifically, basin at zero speed. A natural roll period equal to 𝑇 = 16.30 s (in full

7
A. Serani et al. Ocean Engineering 237 (2021) 109600

Table 6
CFD self-propulsion validation in calm water.
Data RPM 𝐸%EFD Fr 𝐸%EFD
EFD 950 0.33
CFD 922 2.9 0.33 0.0

Fig. 10. Cumulative (top) wave elevation RMS and (bottom) number of roll periods,
as a function of the number of CFD runs.

Fig. 9. CFD self-propulsion validation results in calm water.

scale) was then determined from the recorded roll motion time series.
With an assumed added mass in water, the 𝑘𝑥𝑥 was determined and
ballast was adjusted to meet the required value. The weakness in this
approach is the assumed added mass for roll which is normally in-
between 10 and 20% of the mass moment of inertia in air. The actual
value used depends on the appendage configuration. For destroyer type
ships with relatively large twin rudders, propeller axes, brackets and
bilge keels it is difficult to establish an accurate value for the added Fig. 11. Autocovariance for wave probe 2 of EFD data and CFD selected simulations.
mass.
For this reason, CFD preliminary tests and validation of 𝑘𝑥𝑥 ∕𝐵 value
are conducted, using the following four values: 0.385, 0.400, 0.413, and 6.3. Input wave validation
0.428. CFD replicates EFD tests by releasing the ship from an initial
roll angle of 21 deg using 6 DoF simulations. Validation results are Eight CFD runs are computed to achieve a total of 215 encounter
summarized in Table 5, including average decay period (𝑇 , based on waves. Input wave validation is based on wave probe 2 since it is
first 3 periods), logarithmic decrement (𝛿), and percent errors (𝐸%). the least affected by the ship wake and/or motion-generated waves.
Validation is presented in terms of spectral moments 𝑚0 , 𝑚1 , and 𝑚2 in
A 𝑘𝑥𝑥 = 0.385𝐵 is found to match EFD results with an error of 0.3%
Table 7. CFD errors versus nominal values are smaller than EFD and
for the average decay period and about 10% error for the logarithmic
the average absolute error of CFD vs. EFD is equal to 24%.
decrement. This value is selected for the CFD simulations that follow.
To achieve a closer agreement between EFD and CFD input waves,
It may be noted that the error between EFD and CFD estimation of a subset of CFD runs is selected. To this aim, CFD runs are sorted based
the 𝑘𝑥𝑥 is close to 7%, that is lower than the typical uncertainty of EFD on their wave-probe-2 root mean square (RMS). A subset of 5 CFD
results (about 8%) for twin propeller and rudder configurations. The runs is then selected, considering a trade-off between the cumulative
𝑘𝑥𝑥 assessment is still an open issue for validation studies and different RMS (see Fig. 10, top) and the cumulative number of roll periods
CFD solvers have shown similar discrepancy with EFD (see, e.g., Vison- (see Fig. 10, bottom), compared to those provided by EFD data. RMS
neau et al. 2020, Sanada et al. 2020). Furthermore, it is worth noting values in Fig. 10 (top) are shown along with the associated uncer-
that nowadays MARIN uses an oscillating table to determine the full tainty, evaluated by AC analysis. Selecting 5 CFD runs, the average
inertia matrix of the models. The measurement accuracy for the 𝑘𝑥𝑥 is absolute error of CFD vs. EFD for the input wave reduces to 6.4%
much better than that obtained from a swing table and the problem of (see Table 7). EFD and CFD encounter spectra are compared to the
assuming an added mass in water does not exist anymore. nominal JONSWAP in Fig. 5. It may be noted that the wave probe can
be to some extent affected by the wave produced by the ship motion.
The energy shift towards higher frequencies of EFD and CFD compared
6.2. Self-propulsion RPM validation in calm water to the nominal distribution is quite noticeable, indicating a nonlinear
wave development and possibly effects of interaction with the ship.
Finally, the AC of EFD and selected CFD runs for wave probe 2 is shown
Fig. 9 and Table 6 show the RPM validation using self propul-
in Fig. 11. The curves are quite reasonable, also considering that the
sion computations in calm water. The test is conducted with the self-
spectrum is quite narrow banded.
propelled ship aiming at constant speed corresponding to Fr = 0.33. The Table 8 shows the encounter wave validation for primary variables
CFD simulation achieves the target Fr with an RPM lower than EFD of (time series values), using AC and MBB methods, and secondary vari-
about 3% and equal to 922 in model scale (see the corresponding nom- ables (amplitudes) using the bootstrap method. Validation is achieved
inal 𝐽 in Fig. 4). Since the aim of the present work is to perform CFD for wave elevation SD (about 5.2% error by MBB method) and wave
simulations at the same EFD speed (at least considering the nominal amplitude EV (about 5.8% error by bootstrap method). Errors are
speed), this value is used for CFD simulations in irregular waves. normalized with 2SD, as a metric of data range. Finally, Fig. 12 shows

8
A. Serani et al. Ocean Engineering 237 (2021) 109600

Table 7
Input wave validation.
Analysis N. N. wave Tot. Tot. 𝑚0 𝑚1 𝑚2 Avg.
runs comp. run N .enc. Value 𝐸% Value 𝐸% Value 𝐸% abs.
per run time/𝑇𝑝 waves [m2 ] [m2 rad/s] [m2 rad2 /s2 ] 𝐸%
EFD vs Nominal 7 80 221 130 3.05 41.0 1.20 48.4 0.83 168.4 86.0
CFD (all) vs Nominal 8 80 362 215 2.33 7.89 0.91 11.4 0.64 105.2 41.5
CFD (sel.) vs Nominal 5 80 221 132 2.86 32.4 1.12 37.2 0.79 153.8 74.5
CFD (all) vs EFD −23.5 −25.0 −23.6 24.0
CFD (sel.) vs EFD −6.14 −7.55 −5.45 6.38

Table 8
Encounter wave validation using selected CFD runs.
Variable Method EV SD
Value Uncertainty (𝑈 ) Value Uncertainty (𝑈 )
EFD CFD |𝐸|% EFD CFD 𝑈𝑣 EFD CFD |𝐸|% EFD CFD 𝑈𝑣
Elevation AC 0.20 0.04 4.72 0.73 3.44 3.41 1.75 1.69 1.54 10.79 9.64 14.27
MBB 0.21 0.00 5.74 1.12 2.54 2.54 1.83 1.64 5.19 4.38 9.06 9.23
Amplitude Bootstrap 2.29 2.16 5.75 5.56 6.24 7.76 1.13 0.98 6.64 4.26 5.38 6.32

Note: 𝐸 and 𝑈 are %2SD.

Fig. 12. Quantile function and PDF of wave elevation (a,b) and amplitude (c,d): comparison of EFD and CFD values.

the quantiles and PDF of wave elevation and amplitude, using MBB (surge, sway, and yaw) present the higher errors/discrepancies with
and bootstrap methods respectively. Both EFD and CFD values are respect to EFD. This can be related to the use of the simplified body-
shown along with nominal and theoretical distribution, revealing a force propeller model instead of the actual (discretized) propeller,
good agreement. especially for the surge. Nevertheless, surge, sway, and yaw show the
largest uncertainties for both EFD and CFD. A deeper analysis would
6.4. Ship motions and velocities, rudder angle, and immersion probes vali- require reducing the uncertainties and therefore run longer.
dation EFD distribution for the roll angle shows an evident bi-modal shape,
which is very well captured by CFD (see Fig. 13). This confirms the
Results are summarized in Figs. 13 and 14 and Tables 9 and 10. resonant condition for the roll motion, which contributes achieving
Specifically, Fig. 13 compares EFD and CFD ship motions primary large roll angles. It may be noted that the bi-modal (double peak)
and secondary variables showing PDF and quantile function, whereas shape of the roll PDF is mainly due to the sinusoidal shape of the roll
velocities, rudder angle, and immersion probes results are shown in signal. Additionally, even if there were no bulwarks on the model and
Fig. 14. The results are in good agreement. Primary variables (see the water is not trapped on the deck, for large roll angle the righting
Table 9) are validated with an average error of 9.6% and 8.8% for EV moment (𝑀𝑥,𝑠 , where 𝑠 stays for ship reference system) reduces due
and SD, respectively, and the corresponding uncertainties are equal to to the weight of the water on the deck as shown in Fig. 15, also
11.1% and 11.5%. Considering ship motions only, secondary variables contributing to the bi-modal shape of the roll PDF. As an example,
(see ‘‘Motion avg’’. row in Table 10) EVs are validated with an average Fig. 15 shows one roll period of a regular wave CFD run (described in
error equal to 10.6% and the validation uncertainty equal to 16.9%. the following subsection) and the corresponding total righting moment,
Secondary variables SDs are validated on average with an error equal along with the hull and deck contributions. It is clearly visible how
to 13.2% and the validation uncertainty equal to 17.3%. Planar motions the water on the deck reduces the total righting moment. Furthermore,

9
A. Serani et al. Ocean Engineering 237 (2021) 109600

Fig. 13. Comparison of EFD and CFD quantile (first and third column) and PDF (second and fourth column) of primary (first and second column) and secondary (third and fourth
column) variables for ship motion.

10
A. Serani et al. Ocean Engineering 237 (2021) 109600

Table 9
Primary variables validation by MBB.
Variable Unit EV SD
Value Uncertainty (𝑈 ) Value Uncertainty (𝑈 )
EFD CFD |𝐸|% EFD CFD 𝑈𝑣 EFD CFD |𝐸|% EFD CFD 𝑈𝑣
Surge m 5.17 3.56 3.76 8.66 10.86 10.41 21.42 11.40 23.39 6.55 12.04 9.16
Sway m 5.11 6.23 6.97 17.47 18.52 28.06 8.04 9.53 9.27 9.04 22.20 27.82
Heave m −0.24 −0.20 2.17 2.39 2.22 3.25 0.92 0.91 0.54 6.69 9.64 11.65
Roll deg 6.48 3.02 8.50 2.64 2.33 3.49 20.36 20.01 0.86 5.04 1.97 5.40
Pitch deg −0.04 −0.06 0.81 0.99 0.85 1.31 1.23 1.25 0.81 7.00 7.99 10.72
Yaw deg 1.40 1.03 11.56 6.95 12.51 17.61 1.60 2.07 14.69 4.45 16.00 21.17
Motion avg. 5.63 10.69 8.26 14.32
𝑥-velocity kn 23.34 23.13 7.34 16.40 9.18 17.73 1.43 1.05 13.29 6.02 6.74 7.79
𝑦-velocity kn −0.05 0.04 5.06 2.47 3.66 4.90 0.89 1.03 7.87 3.48 8.55 10.49
Vel. avg. 6.20 11.32 10.58 9.14
R3 m 2.78 3.32 17.31 10.76 7.13 13.66 1.56 1.84 8.97 2.44 2.01 3.40
R5 m 3.72 3.72 9.36 12.35 5.22 12.93 2.35 1.72 13.40 6.81 1.75 6.93
R avg. 13.33 13.29 11.19 5.17
Rudder deg −2.71 4.51 32.49 2.97 7.44 8.58 11.11 12.02 4.10 9.39 7.35 12.30
Average 9.58 11.08 8.84 11.53

Note: 𝐸 and 𝑈 are % 2SD.

Table 10
Secondary variables (amplitudes) validation by bootstrap.
Variable Unit EV SD
Value Uncertainty (𝑈 ) Value Uncertainty (𝑈 )
EFD CFD |𝐸|% EFD CFD 𝑈𝑣 EFD CFD |𝐸|% EFD CFD 𝑈𝑣
Surge m 4.83 6.01 5.46 11.95 18.63 22.12 10.81 10.80 0.05 20.31 32.82 38.57
Sway m 6.36 7.75 10.76 15.07 21.52 45.19 6.46 12.79 48.99 10.87 33.87 67.93
Heave m 1.14 1.18 3.39 6.31 5.67 8.61 0.59 0.61 1.69 3.75 5.07 6.45
Roll deg 25.07 26.60 8.40 5.64 5.79 7.19 9.11 7.02 11.47 4.75 4.74 5.99
Pitch deg 1.45 1.69 15.58 5.75 6.12 8.11 0.77 0.72 3.25 4.40 6.46 7.47
Yaw deg 1.54 2.13 20.21 7.49 6.23 10.28 1.46 1.65 6.51 5.67 11.73 14.42
Motion avg. 10.63 16.92 11.99 23.47
𝑥-velocity kn 1.42 1.29 6.70 5.94 6.09 7.21 0.97 0.65 16.49 5.65 7.78 7.69
𝑦-velocity kn 1.03 1.22 16.96 6.23 5.53 9.53 0.56 0.73 15.18 4.32 8.26 11.60
Vel. avg. 11.83 8.37 15.84 9.64
R3 m 4.40 5.39 52.66 8.37 9.81 11.46 0.94 0.75 10.11 10.42 6.69 11.71
R5 m 5.91 5.20 18.49 9.40 8.88 10.07 1.92 0.78 29.69 11.45 6.00 11.71
R avg. 35.57 10.76 19.90 11.71
Rudder deg 13.15 15.92 17.04 6.34 5.82 8.70 8.13 8.32 1.17 4.68 3.91 6.16
Average 15.97 13.50 13.15 17.25

Note: 𝐸 and 𝑈 are % 2SD.

the flow of water around the sharp deck edge causes additional roll nominally providing with the target speed corresponding to Fr = 0.33 in
damping, also contributing to the PDF double peak. Fig. 15 (bottom) calm water. CFD provided with an RPM equal to 922 for the same nom-
shows the moment contributions as a function of the roll angle during inal speed of Fr = 0.33 in calm water. As observed in Fig. 14 EFD and
the regular wave run. It may be noted how the resulting moment is CFD distributions of 𝑥-velocity are reasonably close, nevertheless CFD
asymmetric and also depends on the roll rate (positive versus negative shows a slightly smaller mean value and a smaller variability, which
rate), being the motion affected by the wave direction. Nonlinearities again may be due to the use of the simplified body-force propeller
emerge clearly for large roll angles especially for negative values, when model.
the deck faces the wave. Finally, to unveil the underlying dynamics and the motion correla-
Further indications of water on deck are derived from the immersion tion DMD of EFD and CFD time histories is performed. The ship 6 DoF
probes signals. Here, probes R3 and R5 are used, laying in between (surge, sway, heave, roll, pitch, and yaw) and the rudder angle are used
stations 7 and 8 at port and starboard sides (see Fig. 2), respectively. as DMD variables. Variables are standardized and an exact/full-rank
EFD and CFD distributions are in a fairly reasonable agreement, as DMD is applied (Diez et al., 2021). Twenty encounter periods from the
Fig. 14 shows. Both EFD and CFD distributions of both probes primary largest-variance runs are used for the analysis with both EFD and CFD.
variable present a significant peak corresponding to 5 m immersion. Modes 𝝓𝑘 are ranked based on their modal participation, ⟨[ℜ(𝑞𝑘 )]2 ⟩,
This corresponds to the peaks observed in the (bi-modal) roll distri- where 𝑞𝑘 are the coordinates of the system state (DMD variables)
bution in Fig. 13. It is interesting to note how EFD and CFD peaks in the modes basis. Couples of complex conjugate modes/frequencies
are very well aligned for starboard side probe 5, which sees the wave are grouped together in the modal participation analysis. Fig. 16
stepping away from the ship. For port side probe 3, which sees the (top) shows the participation of EFD and CFD modes. Fig. 16 (center)
wave approaching and impacting the ship, CFD overpredicts the peak presents the modal participation as a function of the mode angular
location, suggesting wave breaking and spray effects are significant and frequency, |ℑ(𝜔𝑘 )|. Finally, Fig. 16 (bottom) shows the magnitude of
not well captured by the level set method. the mode components for the two most energetic couples of complex
Based on the current analysis, EFD and CFD speed loss equal to 1.3 conjugate modes. It may be noted how overall the modal participation
and 3.4% respectively. This difference may be due to having used a is reasonably captured by CFD computations, but modes 3 and 4 that
different RPM for CFD compared to EFD. Namely EFD RPM equals 950, are significantly underrepresented in CFD compared to EFD (Fig. 16,

11
A. Serani et al. Ocean Engineering 237 (2021) 109600

Fig. 14. Comparison of EFD and CFD quantile (first and third column) and PDF (second and fourth column) of primary (first and second column) and secondary (third and fourth
column) variables for ship velocities, rudder angle, and immersion probes.

top). The most energetic and fast modes are very well captured by motion variables have a double origin. Firstly, the modes involving
CFD and correspond to the roll resonance condition, as Fig. 16 (center) planar motion variables (𝑘 = 3, 4) are underrepresented in CFD compu-
shows. Conversely, modes 3 and 4 correspond to a slower dynamics and
not well captured by CFD. Finally, Fig. 16 (bottom) shows how modes tations. Secondly, looking at the mode components derived from CFD
1 and 2 (most energetic, fast) mainly involves roll, pitch, yaw, rudder, it emerges how the surge is underrepresented if compared to EFD. This
and to a lesser extent heave, whereas modes 3 and 4 mainly involves again may be attributed to the simplified body-force propeller model
planar motion variables, i.e. surge, sway, yaw, and to a lesser extent
rudder. DMD results suggest that the differences between CFD and and will motivate future studies on the most convenient approach to
EFD in surge probability distribution and in general errors for planar propeller modeling.

12
A. Serani et al. Ocean Engineering 237 (2021) 109600

Fig. 17. Wave sequences causing SEs.

Fig. 15. One period of the roll motion and corresponding righting moments for a
regular wave CFD run.

Fig. 18. Clustering metrics results: (top) normalized within clusters sum of squares;
(bottom) average silhouette.

Table 11
EFD and CFD number of encounter waves and severe events (SE).
Data N. enc. N. severe Roll amplitude SE%
waves events (EV + 1.5SD) N. enc. waves
EFD 130 21 16.2
≥ 39 deg
(Usable) (92) (14) (15.2)
CFD 132 7 5.3
≥ 37 deg
(Usable) (102) (5) (4.9)

events is evaluated versus the threshold value 𝜎 for the roll amplitude.
Selecting a threshold of 𝜎 = 1.5SD (corresponding to a roll amplitude
of 39 and 37 deg for EFD and CFD, respectively), about 15 and 5% of
the events are considered as severe for EFD and CFD, respectively (see
Table 11). The corresponding wave sequences are shown in Fig. 17.
For all wave sequences, the EVw and SDw are evaluated for wave
heights, amplitudes, and encounter periods. The weights used are in-
versely proportional to the wave index, where 1 indicates the closest
to the severe event and 5 the farthest in the past. Average values for
all waves and only those causing severe events are summarized in
Table 12. It can be noted that all the wave sequences have an encounter
period close to the natural roll period, confirming that the ship is sailing
close to the resonant condition for roll. The common denominator of
Fig. 16. DMD analysis for EFD and CFD. the wave sequences causing the SEs is identified in small values of the
SD for both wave height and encounter period. This mean that the SEs
are caused by nearly-regular waves, close to resonance condition.
As an additional analysis, 𝑘-means clustering is performed for the
6.5. Causal sequence identification
roll amplitude caused by the sequence, wave height and encounter
period (within the sequence) SDs. Standardized data are used. The
Due to the use of sequences and the finite number of EFD and CFD optimal number of clusters for the representation of current data is
runs, the number of usable waves is lower than the actual number of defined by the WCSS and the silhouette. These are shown in Fig. 18,
recorded waves. Specifically, the number of usable waves is equal to which reveals that four clusters are a good compromise for current data
92 and 102 for EFD and CFD, respectively. The percentage of severe clusterization.

13
A. Serani et al. Ocean Engineering 237 (2021) 109600

Fig. 19. Clusters distribution: EFD (left) and CFD (right).

Clusters of EFD and CFD are shown in Fig. 19. Results are con- Table 12
sistent with earlier findings and show how reducing wave height and EFD and CFD wave sequences statistics, considering wave probe 2.

encounter period SDs produces larger roll angles (see cluster 1 for both Variable Avg. EVw Avg. SDw

EFD and CFD). EFD and CFD results are remarkably consistent. EFD CFD 𝐸% EFD CFD 𝐸%
𝐴 [m] All 2.30 2.17 −5.46 0.65 0.64 −0.62
6.6. Deterministic reconstruction of severe and rare events by regular-wave SE 2.00 2.85 42.2 0.51 0.60 15.9
𝛥% −12.9 31.1 −21.5 −6.76
CFD simulations
𝐻 [m] All 4.60 4.35 −5.43 1.18 1.21 3.52
SE 4.01 5.70 42.1 0.89 1.11 18.0
The deterministic reconstruction of SEs is based on regular wave 𝛥% −12.8 31.0 −24.4 −7.68
CFD simulations. It may be noted that the choice of running regular 𝑇𝑒 [s] All 16.16 15.87 −1.79 1.45 1.36 −6.74
as opposed to irregular waves is determined after severe and rare SE 16.39 15.96 −2.62 1.07 1.08 72.6
𝛥% 1.42 0.57 −26.4 −21.0
events are identified along with the corresponding wave sequences,
and statistical and clustering analyses is performed. For the current Note; 𝐸 are % of EFD.
test case and both EFD and CFD data, severe/rare events are caused by
wave sequences characterized by encounter waves with almost identi-
cal height and period (small SD of these parameters within the wave highly nonlinear regime, causing one of the propellers to operate out
sequence), close to resonance condition for the roll motion. Therefore, of the water with the subsequent reduced thrust and slip back from
it can be concluded that, in this case, (approximately) regular waves the wave crest into the throat, as Fig. 21 shows. It is also interesting
are causing severe/rare events and therefore regular waves are used to note how the inflection point for the roll (inception of capsizing)
for the recreation of such events. Accordingly, two proof-of-concept occurs when the pitch is at its peak (bow down). Finally, a comparison
approaches are shown here: of the ship attitude in deterministic wave at the first eight roll peaks
is shown in Fig. 22. The results are compared to those obtained in
1. Reconstruction of SEs using wave amplitude and period identi-
irregular waves, for one of the sequences causing severe roll angles
fied as condition causing SEs;
(first column), showing a remarkable agreement.
2. Reconstruction of rare events (capsizing) increasing the wave
amplitude.
7. Conclusions and future work
A regular wave is used, characterized by wave amplitude and encounter
period as per Table 12: 𝐴0 = 2.85 m and 𝑇𝑒 = 15.96 s. Fig. 20 The stochastic assessment and validation of URANS simulations of
compares the CFD stochastic wave sequences causing the SEs and the free-running 5415M model in stern-quartering waves at sea state 7 is
deterministic regular wave used for the first proof of concept. For the presented. Validation studies also include roll decay at zero speed and
second proof of concept a parametric analysis increasing the wave self-propulsion RPM studies.
amplitude is conducted. Specifically, three further CFD simulations are Roll decay and self-propulsion studies are successfully performed.
performed with 𝐴 = 1.1, 1.2, and 1.4 times 𝐴0 . Specifically, EFD 𝑘𝑥𝑥 ∕𝐵 value of 0.413 produce a 6% error for the roll
The results of the deterministic reconstruction of severe and rare damped period when used with CFD. Therefore, a 𝑘𝑥𝑥 ∕𝐵 value equal
events are shown in Fig. 21 wave elevation at probe 2 and ship motions. to 0.385 is selected for CFD runs in waves, achieving an error for the
Using the deterministic wave, the roll motion reaches the SE threshold, roll period equal to only 0.3%. CFD self-propelled simulations in calm
whereas the rare event (capsizing) is achieved with 𝐴 = 1.4𝐴0 . It may water achieve the target Fr = 0.33 with an RPM lower than EFD of
be noted that in this latter case the encounter wave frequency suddenly about 3%. This is used for CFD runs in waves.
increases (see Fig. 21 top, after about 100 s) as the ship is upside- The stochastic validation of the input wave is achieved by spectral
down and stopping its course, as well as the numerical carriage. It analysis, autocovariance, and bootstrap methods. EFD data include 7
is also worth noting that capsizing is not associated with surf riding runs for a total of 130 encounter waves. CFD simulations include 8
ahead and broaching. Roll resonance produces large roll angles in the runs for a total of 215 encounter waves. EFD versus nominal spectrum

14
A. Serani et al. Ocean Engineering 237 (2021) 109600

motion variables present the higher errors/discrepancies compared to


EFD, which may be attributed to the use of the simplified body-force
propeller model, especially for the surge. This is confirmed by the
DMD analysis, where two modes clearly emerges from motion time
histories. The first and fast dynamics is associated to roll resonance,
mainly involves roll, pitch, yaw, rudder, and to a lesser extent heave,
and is very well captured by CFD. The second and slower dynamics
mainly involves planar motion variables and to a lesser extent rudder,
and is underrepresented in CFD results. Finally, immersion probes
EVs and SDs show an average error close to 12% with an associated
uncertainty close to 9%, likely due to the use of the level set method
in combination with rough waves and large motions. For the same
variables, amplitudes are assessed showing similar trends although with
larger errors and uncertainties.
Fig. 20. Comparison of stochastic and deterministic CFD waves.
Conditions causing severe events (large roll angles) are identified
considering sequences of five encounter waves and their statistics (av-
erage EVw and SDw of wave amplitude/height and encounter periods).
Severe events are identified when the roll amplitude exceeds its EV +
1.5SD, which corresponds to 39 and 37 degrees for EFD and CFD,
respectively. About 15 and 5% of EFD and CFD sequences, respectively,
are identified as causing extremely large roll angles. These sequences
are characterized by an almost regular wave inducing roll resonance.
Specifically, the within-sequence SDw of both wave height and en-
counter period reduces significantly if compared to all sequences in
EFD and CFD records. This result is confirmed by 𝑘-means clustering
analysis, which shows how the cluster with the largest roll amplitudes
corresponds to small within-sequence SDs of wave height and period.
EFD and CFD produce remarkably close results. Based on this analysis,
two proof-of-concept results are achieved. First, severe events (large
roll angle) are reconstructed by regular wave CFD simulations. Second,
a rare event (capsizing) is achieved by increasing the wave amplitude.
It may be noted how large roll angles arise in the nonlinear regime and
identify conditions where the ship is on the verge of capsizing. Results
show how capsizing is mainly due to roll resonance, as opposed to surf
riding and broaching instabilities.
Overall the objectives of current study are achieved, namely CFD
(URANS) validation for heavy weather, identification of severe and
rare events, and deterministic reconstruction as a proof of concept. The
results are in general affected by the use of a simplified non-interactive
body-forced model for the propeller, that, for the objective of the
present study (statistical validation of relevant variables in irregular
waves), was deemed reasonable within the activities of the NATO
AVT-280 task group. To achieve better validation results, the use of
more sophisticated body-force models, as well as a fully discretized
propellers, is under investigation within the NATO AVT-348 task group
on the ‘‘Assessment of Experiments and Prediction Methods for Naval
Ships Maneuvering in Waves’’. Nevertheless, since these methodologies
Fig. 21. Deterministic reconstruction of severe and rare events by CFD simulations in can be quite expensive, a good compromise between numerical accu-
regular waves: WP2 and 6 DoF.
racy and computational efficiency can be given by coupling potential
flow and RANS solvers, as shown in Chase and Carrica (2013).
It may be finally emphasized that stochastic CFD simulations are
shows an error of 41% for the zero-th moment. Therefore, selected CFD too expensive to be advocated as a method to identify rare events
runs are used to achieve an input wave as close as possible to EFD. on a large sample basis. Nevertheless, if rare events are identified
Specifically, 5 runs are used with 132 encounter waves. These provide either anecdotally by captain, by expensive model tests data, and/or by
with an error equal to 6.1% for the zero-th moment, 1.5% for the wave reduced order models (lower fidelity but faster such that large samples
SD and 5.8% for the wave amplitude EV. EFD and CFD stochastic can be achieved) and assuming sufficient environmental information is
uncertainties are both close to 10%. A reduction of the uncertainty available, CFD is able to (a) reproduce statistically the response (based
would require both EFD and CFD to run longer. Note that EV and SD on a smaller sample) and (b) deterministically reconstruct severe events
errors and uncertainties are normalized with 2SD, which is used as a for detailed studies.
metric for the data range. Future work will include further analysis for the identification and
The statistical assessment and validation of CFD versus EFD re- assessment of the severe events based on stochastic distribution of regu-
sponse is achieved by bootstrap methods. Ship motions SDs are vali- lar waves (He et al., 2013) and sequential sampling strategy (Mohamad
dated on average with an error of 8.3% and an associated uncertainty and Sapsis, 2018). Extensions to other test cases (not in resonance
of 14.3%. Ship velocity EVs are validated with an error equal to 6.2% conditions) are also advisable and will be considered for future studies.
and an uncertainty equal to 11.3%. Rudder angle SD is also validated Finally, the current work lay the ground for the deterministic and
with error and uncertainty equal to 4.1 and 12.3% respectively. Planar stochastic assessment of ship maneuvering in waves.

15
A. Serani et al. Ocean Engineering 237 (2021) 109600

Fig. 22. Comparison of ship attitude conditional to deterministic wave amplitudes and roll peaks.

Declaration of competing interest Turbulent Free-Surface Flow Simulations around a Ship with High
Resolution Grids using CFDShip-Iowa’’. CNR-INM is supported by the
The authors declare that they have no known competing finan-
University of Iowa, USA under a subaward to the same grant. The Office
cial interests or personal relationships that could have appeared to
influence the work reported in this paper. of Naval Research, USA grants N00014-17-1-2083 and N00014-17-1-
2084 under administration Drs. Thomas Fu, Woei-Min Lin and Ki-Han
Acknowledgments Kim partially sponsored IIHR research. Dr. Andrea Serani is also grate-
ful to the National Research Council of Italy, for its support through the
The authors acknowledge the permission granted by Cooperative
Short-Term Mobility Program 2018. The research is performed within
Research Navies (CRN), The Netherlands participants to utilize the
experimental data. The University of Iowa is supported by the Korea NATO STO Task Group AVT-280 ‘‘Evaluation of Prediction Methods for
Institute of Science and Technology under grant #19026000 ‘‘Unsteady Ship Performance in Heavy Weather’’.

16
A. Serani et al. Ocean Engineering 237 (2021) 109600

References Rousseeuw, P.J., 1987. Silhouettes: a graphical aid to the interpretation and validation
of cluster analysis. J. Comput. Appl. Math. 20, 53–65.
Belenky, V., Pipiras, V., Weems, K., 2015. Statistical uncertainty of ship motion data. Sadat-Hosseini, H., Carrica, P., Stern, F., Umeda, N., Hashimoto, H., Yamamura, S.,
In: 12th International Conference on the Stability of Ships and Ocean Vehicles, Mastuda, A., 2011. CFD, system-based and EFD study of ship dynamic instability
STAB, Glasgow, UK, June, pp. 14–19. events: Surf-riding, periodic motion, and broaching. Ocean Eng. 38 (1), 88–110.
Carlstein, E., et al., 1986. The use of subseries values for estimating the variance of a Sadat-Hosseini, H., Kim, D.H., Toxopeus, S., Diez, M., Stern, F., 2015. CFD and potential
general statistic from a stationary sequence. Ann. Statist. 14 (3), 1171–1179. flow simulations of fully appended free running 5415M in irregular waves. In:
Chase, N., Carrica, P.M., 2013. Submarine propeller computations and application to World Maritime Technology Conference, Providence, RI, Nov, pp. 3–7.
self-propulsion of DARPA Suboff. Ocean Eng. 60, 68–80. Sadat-Hosseini, H., Stern, F., Olivieri, A., Campana, E.F., Hashimoto, H., Umeda, N.,
Diez, M., Broglia, R., Durante, D., Olivieri, A., Campana, E.F., Stern, F., 2018. Statistical Bulian, G., Francescutto, A., 2010. Head-wave parametric rolling of a surface
assessment and validation of experimental and computational ship response in combatant. Ocean Eng. 37 (10), 859–878.
irregular waves. J. Verif. Valid. Uncertain. Quantif. 3 (2). Sanada, Y., Kim, D.-H., Sadat-Hosseini, H., Stern, F., Hossain, M.A., Wu, P.-C., Toda, Y.,
Diez, M., Serani, A., Campana, E.F., Stern, F., 2021. Data-driven modelling of ship Otzen, J., Simonsen, C., Abdel-Maksoud, M., Scharf, M., Grigoropoulos, G., 2020.
maneuvers in waves via dynamic mode decomposition. In: Proceedings of the Assessment of experimental and CFD capability for KCS added power in head and
9th International Conference on Computational Methods in Marine Engineering, oblique waves. In: Proceedings of the 33rd Symposium on Naval Hydrodynamics,
MARINE 2021. Osaka, Japan.
Efron, B., 1981. Nonparametric estimates of standard error: the jackknife, the bootstrap Schmid, P.J., 2010. Dynamic mode decomposition of numerical and experimental data.
and other methods. Biometrika 68 (3), 589–599. J. Fluid Mech. 656, 5–28.
He, W., Diez, M., Zou, Z., Campana, E.F., Stern, F., 2013. URANS study of Delft Serani, A., Stern, F., Campana, E.F., Diez, M., 2021. Hull–form stochastic optimization
catamaran total/added resistance, motions and slamming loads in head sea includ- via computational–cost reduction methods. Engineering with Computers 1–25.
ing irregular wave and uncertainty quantification for variable regular wave and Silverman, B.W., 2018. Density Estimation for Statistics and Data Analysis. Routledge.
geometry. Ocean Eng. 74, 189–217. Toxopeus, S., Sadat-Hosseini, H., Visonneau, M., Guilmineau, E., Yen, T.-G., Lin, W.-
Huang, J., Carrica, P.M., Stern, F., 2008. Semi-coupled air/water immersed bound- M., Grigoropoulos, G., Stern, F., 2018. CFD, potential flow and system-based
ary approach for curvilinear dynamic overset grids with application to ship simulations of fully appended free running 5415M in calm water and waves. Int.
hydrodynamics. Internat. J. Numer. Methods Fluids 58 (6), 591–624. Shipbuild. Prog. 65 (2), 227–256.
Judge, C., Mousaviraad, M., Stern, F., Lee, E., Fullerton, A., Geiser, J., Schleicher, C., van Walree, F., Serani, A., Diez, M., Stern, F., 2020. Prediction of heavy weather
Merrill, C., Weil, C., Morin, J., et al., 2020. Experiments and CFD of a high-speed
seakeeping of a destroyer hull form by means of time domain panel and CFD
deep-v planing hull–part II: Slamming in waves. Appl. Ocean Res. 97, 102059.
codes. In: Proceedings of the 33rd Symposium on Naval Hydrodynamics, Osaka,
Ketchen, D.J., Shook, C.L., 1996. The application of cluster analysis in strategic
Japan.
management research: an analysis and critique. Strateg. Manag. J. 17 (6), 441–458.
van Walree, F., Visser, C., 2010. Analysis And Post Processing of Measured Signals
Kutz, J.N., Brunton, S.L., Brunton, B.W., Proctor, J.L., 2016. Dynamic Mode
Vol. 2: Seakeeping Water on Deck Tests. Technical Report 22810-1-SMB, Maritime
Decomposition: Data-Driven Modeling of Complex Systems. SIAM.
Research Institute Netherlands.
Lloyd, S., 1982. Least squares quantization in PCM. IEEE Trans. Inform. Theory 28 (2),
Visonneau, M., Queutey, P., Deng, G.B., Toxopeus, S., Sadat-Hosseini, H., Stern, F.,
129–137.
2020. Evaluation of Prediction Methods for Ship Performance in Heavy Weather,
Miecznikowski, J.C., Wang, D., Hutson, A., 2010. Bootstrap MISE estimators to obtain
STO-TR-AVT-280, Chapter 6:ONRT in calm water and waves. Technical Report,
bandwidth for kernel density estimation. Comm. Statist. Simulation Comput. 39
NATO.
(7), 1455–1469.
Wang, J., Zou, L., Wan, D., 2017. CFD simulations of free running ship under course
Mohamad, M.A., Sapsis, T.P., 2018. Sequential sampling strategy for extreme event
statistics in nonlinear dynamical systems. Proc. Natl. Acad. Sci. 115 (44), keeping control. Ocean Eng. 141, 450–464. http://dx.doi.org/10.1016/j.oceaneng.
11138–11143. 2017.06.052.
Noack, R., 2005. SUGGAR: a general capability for moving body overset grid assembly. Wilson, R.V., Carrica, P.M., Stern, F., 2006. Unsteady RANS method for ship motions
In: 17th AIAA Computational Fluid Dynamics Conference, p. 5117. with application to roll for a surface combatant. Comput. & Fluids 35 (5), 501–524.
Petacco, N., Gualeni, P., 2020. IMO second generation intact stability criteria: General Zhang, L., Zhang, J., Shang, Y., 2021. A practical direct URANS CFD approach for
overview and focus on operational measures. J. Mar. Sci. Eng. 8 (7), 494. the speed loss and propulsion performance evaluation in short-crested irregular
head waves. Ocean Eng. 219, 108287. http://dx.doi.org/10.1016/j.oceaneng.2020.
108287.

17

You might also like