You are on page 1of 45

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

net/publication/364958555

Estimation of hydrodynamic derivatives of an appended KCS model in open and


restricted waters

Article in Ocean Engineering · December 2022


DOI: 10.1016/j.oceaneng.2022.112947

CITATIONS READS
3 299

2 authors:

Hafizul Islam Carlos Guedes Soares


Maritime Research Institute Netherlands University of Lisbon
49 PUBLICATIONS 481 CITATIONS 2,701 PUBLICATIONS 52,194 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Hafizul Islam on 01 November 2022.

The user has requested enhancement of the downloaded file.


Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

Estimation of Hydrodynamic Derivatives of an Appended KCS Model in


Open and Restricted Waters

Hafizul Islam and C. Guedes Soares∗

Centre for Marine Technology and Ocean Engineering (CENTEC), Instituto Superior Técnico, Universidade
de Lisboa, Portugal

Abstract

The paper studies and compares the manoeuvring capabilities of an appended KRISO Container Ship (KCS)
hull model in both open and restricted waters. The KCS hull model with a static rudder and body force
model-based propeller is simulated using the open-source computational fluid dynamics toolkit
OpenFOAM. Initially, open water simulations are performed for the propeller model to replicate
experimental open water curves and estimate parameters for the body force model. Next static drift, pure
sway, and pure yaw simulations are performed using both overset and static mesh, from which
hydrodynamic derivatives are calculated. Next, restricted water simulations are performed using overset
mesh for the ship in a channel with defined width and depth. Finally, manoeuvring derivatives are
predicted for restricted waters and compared against the open water values. The study concludes that
consideration of appendages notably influences the manoeuvring study outcome, and the ship
experiences notably larger forces during manoeuvring in restricted water and the manoeuvring
capabilities are also notably compromised.

Keywords: Ship manoeuvring; PMM; hydrodynamic derivatives; KCS; restricted waters; CFD.

1 Introduction
Ships have been an ideal mode for the transportation of bulk goods for centuries. With rapid globalization,
the quantity of goods being shipped has increased substantially, and so has the size of ships. Thus, modern
ships have become larger and bigger in terms of length and volume capacity, while ensuring optimum
performance and economy. Increasing environmental concerns have also forced designers to rethink
design requirements both in terms of seakeeping and manoeuvring. Evaluation of ship stability and
manoeuvrability are among the key requirements at the design stage to ensure safety and reliability.
Generally, for ship manoeuvrability prediction, two approaches are followed. The first is to conduct tests

1
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

in manoeuvring basins that provide characteristics parameters like tactical diameter, overshoot angle, and
so on. The alternative is the mathematical model-based approach (Abkowitz, 1964, 1988; Ogawa et al.,
1977; Sutulo and Guedes Soares, 2015) where ship motion equations are solved based on hydrodynamic
coefficients or manoeuvring derivatives. This approach also requires captive model tests to predetermine
the coefficients (Abkowitz, 1980; Sutulo and Guedes Soares, 2004). Apart from model tests, there are also
theoretical approaches, utilization of a full-scale database of measurements (Kose et al., 1992; Guedes
Soares et al, 1999, 2004), and empirical methods (Kijima and Nakiri, 2003; Sutulo and Guedes Soares,
2019), and numerical approaches (Sutulo et al. 2002, Araki et al., 2012; Sutulo and Guedes Soares, 2014;
Bonci et al., 2015). However, among the methods, experimental and numerical studies remain popular
due to their versatility and reliability.

Since manoeuvring basins are expensive and rarely available, the industry and academia generally rely on
captive model tests like the oblique towing test (OTT), rotating arm test (RAT), circular motion test (CMT),
and planar motion mechanism (PMM), to derive the manoeuvring derivatives and use the mathematical
models. However, with decades of development in numerical models like CFD, such studies are also being
replicated using numerical tanks with sufficient reliability.

While CFD-based manoeuvring studies were being investigated since the late 1990s, the efforts received
great momentum through the SIMMNAN (2008) workshop on ship manoeuvring. The workshop
introduced three benchmark ship models (KCS, KVLCC2, DTMB5415) and case studies as references for
validation studies. The workshop focused mostly on PMM simulations and received contributions from a
large number of researchers (Broglia et al., 2008; Cura-Hochbaum et al., 2008; Gullmineau et al., 2008;
Miller, 2008). This encouraged subsequent studies focusing on oblique motions (Wang et al., 2011); zig-
zag, turning circles (Sanada et al., 2012), and PMM motions (Simonsen et al., 2012; Yao et al., 2016); zig-
zag simulations with self-propulsion (Broglia et al., 2013; Mofidi and Carrica, 2014; Shen et al., 2015);
prediction of hydrodynamic derivatives from PMM results (Kim et al., 2015; Hajivand and
Mousavizadegan, 2015; Gadelho et al., 2018; Islam and Guedes Soares, 2018; Yao et el., 2021; Franceschi
et al., 2021).

Historically, manoeuvring derivatives were considered to be a property of the hull form and are
independent of appendages. Thus, the captive model tests were designed to record changes in forces and
moments while the ship goes through externally imposed motions. However, with the improvement of
knowledge, and experimental and computational capabilities, it was understood that rudder and propeller
configuration also create a notable impact on ship stability and manoeuvrability. Thus, in the subsequent

2
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

SIMMAN workshops (2014, 2019), captive model benchmark cases were introduced with an active rudder
and propeller. Subsequently, several researchers performed manoeuvring simulations with active rudder
and propeller-like effects (Broglia et al. 2015; Dubbioso et al. 2016; Hasanvand and Hajivand 2019;
Hasanvand et al. 2021; Shang et al. 2021; Kim et al. 2021, 2022a). However, not everyone could opt for
rotating propellers since they are computationally expensive and resorted to body force-based propulsion
models. Their efficiency has already been discussed in several works like Kawamura et al. (1997); Broglia
et al. (2013); Yao (2015); Gaggero et al. (2017); Shang et al. (2021).

On the other hand, ship manoeuvring capabilities also vary notably depending on their operating
environment. Thus, environmental influences like wave influence on manoeuvring capabilities have
already been elaborately discussed in European projects like SHOPERA (Papanikolaou et al., 2016) and
other researcher institutes (Sanada et al., 2018; Diez et al., 2022). Although ocean-going vessels spend
most of their lifetime in open waters, occasionally, they also need to pass through restricted waters to
cross channels or dock in ports. In restricted waters, with shallow depth, and narrow channel width, ships
experience shallow water and bank effect, which notably increases their turning parameters (ITTC, 2022).
Such conditions require ships to slow down, which further affects their manoeuvrability capabilities. As
such, understanding how restricted waters affect ship manoeuvrability is very important to ensure safe
navigation in all operational environments.

Among some of the previous works, Carrica et al. (2016) experimentally and numerically investigated the
manoeuvrability of a KCS model in shallow waters. Liu et al. (2019) investigated the DTC hull in shallow
waters and observed the effects of water depth on shop speed and dynamic sinkage. Xu and Guedes
Soares (2020) and Xu et al. (2020) also experimentally studied the DTC model in shallow waters and
developed nonlinear implicit models and system-based identification models for the estimation of
manoeuvring properties. Kim et al. (2022b) investigated a KCS model with an active rudder and body
force-based propeller model using CFD to assess the influence of different water depths on ship
manoeuvrability.

Although the influence of the rudder and propeller on manoeuvring simulations has been previously
studied, the quantitative impact of their absence on manoeuvring derivative prediction has been rarely
discussed. Detailed and comparative information on the influence of appendages on forces and moments
is also scarce. This paper aims to address some of these missing details by extending the previous work
by Islam and Guedes Soares, (2018), where a bare KCS hull was studied. In this study, an advanced
approach is described where PMM simulations are performed for the KCS model including a static rudder

3
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

and a body force model (rotor disk) based propeller. For this study, initially, simulations are performed
for the KCS propeller model in open water and compared with Experimental Fluid Dynamics (EFD) results
for validation. The simulation results are used to calibrate the rotor disk model used in the appended hull
simulations. Next, static drift simulations are performed using overset mesh, and pure sway and pure yaw
simulations are performed by applying imposed motion on the domain. Finally, static and dynamic
derivatives are computed from the simulation results and compared with bare hull results.

Next, after the open-water PMM simulations with an appended hull, manoeuvring derivatives are
predicted for an appended KCS hull using PMM simulations in restricted waters and compared with open-
water results. For the study, a restricted channel is considered roughly following the dimensions of the
Suez Canal. This is considered since Suez Canal is a very popular and important shipping route and is
frequented by large and ultra-large container vessels. The recent incident in the Canal involving the “Ever
Given” vessel also makes it an attractive case study.

2 Case Setup and Numerical Settings


Manoeuvring simulations for a KCS model are performed in open and restricted waters, using the open-
source CFD toolkit, OpenFOAM. The sub-sections describe the hull model, case setup, meshing, and
numerical model.

2.1 Ship Model and Case Setup


The KRISO Container ship (KCS) model used in this paper is a 3600TEU capacity container ship designed
by KRISO (formerly MOERI) for research purposes (SIMMAN 2014; Tokyo, 2015; Ino et al., 2021). The ship
is a very popular test model, as many experimental and CFD test results are open to the public and have
been discussed in many workshops and conferences like Gothenburg, Tokyo, and SIMMAN. Figure 1 shows
the body plan of the KCS model, and the ship hull, rudder, and propeller specifications are provided in
Table 1 (SIMMAN 2014). The simulations are performed for the hull with a fixed rudder. All simulations
are performed on the model scale.

Figure 1: Body plan of the KCS model

4
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

Table 1: Specifications of the KCS ship hull, rudder, and propeller.

Scale 1.000 31.600


Main particulars
Lpp (m) 230.0 7.2785
Lwl (m) 232.5 7.3228
Bwl (m) 32.2 1.0190
D (m) 19.0 0.6013
T (m) 10.8 0.3418
Displacement (m3) 52030 1.6489

S w/o rudder (m2) 9530 9.5437


CB 0.651 0.651
CM 0.985 0.985
LCB (%), fwd+ -1.48 -1.48

Rudder

Type semi-balanced horn rudder semi-balanced horn rudder

S of rudder (m2) 115 0.1152


Lat. area (m2) 54.45 0.0545
Turn rate (deg/s) 2.32 13.04

Propeller
Type FP FP
No. of blades 5 5
D (m) 7.9 0.250
P/D (0.7R) 0.997 0.964
Ae/A0 0.800 0818
Rotation Right hand Right hand
Hub ratio 0.180 0.191

The study aims at predicting the manoeuvring derivatives of the KCS model in open water and restricted
conditions. Thus, simulations are performed with two case setups. One with open water or open sea
conditions with no reflecting side or bottom boundaries, and another with a restricted channel or domain
geometry.

For the restricted channel case setup, the simulation domain size is adjusted to match the scaled
dimensions of the Suez channel. The Suez Canal has a length of 193.3 km, with a maximum width of 205

5
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

m and a maximum draft of 24 m Suez Canal, Egypt (2020). Although part of the canal has a limited width
of 77.5 m, the maximum width is considered in this case to reduce the bank effect, and to reduce overall
uncertainty in simulation results. Since the studied hull model has a scaling of 1:31.6, the domain size is
adjusted accordingly.

2.2 Meshing and Boundary Settings


In this study, three different mesh configurations are used. For the open-water single-phase propeller
simulation, a cylindrical domain is used with a sliding mesh topology. For the PMM simulations in open
sea conditions, both a single domain and an overset mesh topology are used. As for restricted water
conditions, all simulations are performed with an overset mesh topology. All meshes are generated using
the mesh generation utility of OpenFOAM. Quantitative information on the meshes is provided in the
verification and validation study section.

2.2.1 Propeller simulation


For the propeller simulation, a cylindrical domain is chosen to represent a tunnel test. The domain has a
diameter of roughly 3 times the propeller diameter, while the length is set at roughly 5 times, with more
than one diameter space between the propeller and the inlet. As for the boundary conditions, the circular
cylinder is defined as a patch with the no-slip condition. The inlet is given fixed flow velocity, the propeller
is defined as a wall, and the outlet has zero return inflow condition. The flow at the inlet is assumed to be
uniform

A cylindrical block is generated using the blockMeshDict utility of OpenFOAM. The blockMeshDict
generates a structured mesh block that acts as the simulation domain. The domain is further refined three
times using the topoSet and refineMeshDict. The topoSet defines an area within the mesh, which is later
refined (split in half) by the refineMeshDict. Two cylindrical blocks are generated using topoSet to house
the cylinder which will allow rotating slipping motion. Finally, the propeller along with the shaft is
integrated into the domain using snappyHexMesh. SnappyHexMesh integrates the defined object or
geometry into the simulation domain by cutting and fitting the cells in the domain, to represent the
inserted geometry. The final mesh resolution is 0.5 million with a grid size of 0.06205m in the x, y, and z
directions. The mesh configuration is shown in Figure 2, with refinements in the lateral and vertical planes.

6
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

Figure 2: Mesh for the open water propeller simulation; (a) simulation domain, (b) mesh refinement in
the y-z plane, (c) meshing on the propeller and propeller shaft, (d) mesh refinement in the x-y plane.

2.2.2 Manoeuvring simulations at open sea


For the ship manoeuvring simulations, the domain or outer domain is generated following the general
ITTC-2011 guidelines, with the inlet placed one ship length before the bow, the outlet placed two ship
lengths after the stern, and the side boundaries at one ship length in each lateral direction from the sides
of the hull (to the starboard and port). The simulation domain for the overset mesh configuration is shown
in Figure 3, with representative dimensions of the domain.

For the simulations with open sea conditions, the bottom and sides of the open sea simulation domain
are modelled as symmetry planes. The symmetry boundary condition acts like a mirror surface, where the
fluxes and normal components of all variables across the symmetry are set to zero. The symmetry in this
case works as a non-reflecting boundary condition that ignores the interaction of the flow field with the
sides and the top.

The pressure boundary condition at the inlet is a zero-gradient (Neumann boundary condition), and the
outlet boundary supplies a constant pressure condition (Dirichlet boundary condition). Regarding the
velocity field, the inlet is set as a constant velocity condition, and the outlet is set to be a generic outflow
condition with zero return inflow. Initial values for the turbulence parameters are calculated following the
Reynolds number (Labanti et al., 2016). The OpenFOAM built in a rough wall function is used to reduce
the mesh dependency near the hull wall to capture the boundary layer and incorporate general hull

7
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

roughness. The flow at the inlet is assumed to be uniform. As for the inner domain in an overset mesh
configuration, all boundaries are set as overset faces, which allows the transfer of flux information
between the inner and outer domains.

Figure 3: Simulation domain for the overset mesh setup, with the name of the boundaries on the outer
domain.

For the open sea pure sway and yaw cases, the conventional approach for mesh generation is used.
Initially, a simulation domain is generated using blockMeshDict, which is then refined six times near the
hull area. Finally, the hull with a static rudder is integrated using snappyHesMesh. One exception, in this
case, is the cylindrical refinement in between the propeller screw cap and rudder to provide a higher
resolution for the rotor disk. The mesh configuration is shown in Figure 4. A mesh resolution of 3.5 million
cells is used for the study.

Next, an overset mesh is generated to perform the static drift simulations with the rotor disk model. In
the overset structure, the hull form, the rudder, and the rotor disk model are placed inside the inner mesh
which contains a relatively high mesh resolution. For this study, the inner domain is generated with a
length of 1.5L (L representing Lpp of the ship), a width of 1.5B (B representing ship width), and 2.6H (H
representing the ship height). The domain is generated using blockMeshDict, which is further refined four
times using refineMeshDict. The hull and the rudder are incorporated using the snappyHexMesh. The
rotor disk area was defined following the propeller diameter and the width (or chord length) of the

8
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

propeller blade sections. Next, the outer mesh is generated with a mesh resolution similar to or slightly
coarser than the inner domain in the overlapping area. The outer mesh is generated using blockMeshDict
and then three refinements are performed near the overset area to improve the mesh density. The two
domains are incorporated during the simulation using mergeMeshes. The overset mesh configuration is
shown in Figure 5. The mesh resolution used for the open water static drift studies is around 3.9 million
cells. The mesh resolution is discussed in greater detail in the grid dependency study section.

Figure 4: Static mesh for pure sway and pure yaw simulations; (a) simulation domain, (b) mesh refinement
in the x-y plane, (c) mesh refinement in the y-z plane, (c) meshing on the hull and rudder.

9
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

Figure 5: Overset mesh for static drift simulations; (a) simulation domain including both inner and out
domain, (b) mesh refinement in the y-z plane, (c) mesh refinement in the x-y plane, (c) meshing on the
hull and rudder.

2.2.3 Manoeuvring simulations in restricted waters


For the restricted channel simulations, the domain and mesh topology remained the same as in the open
water case, except for the adjustment in domain width and depth. Following the Suez channel dimensions,
a channel with a 6m width, and 0.68m depth (twice of ship draft) was generated for the simulations, while
keeping the domain length the same as open sea conditions. Figure 6 represents the restricted water
simulation domain with the boundary conditions.

10
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

In the simulations, the inlet, outlet, atmosphere, and inner mesh boundary conditions remain the same
as in the open sea condition simulations. However, for restricted water cases, the domain bottom and
side wall are defined as walls with small roughness. These walls allowed diffraction of the wake waves
generated by the passing ship.

Figure 6: Simulation domain for the PMM simulations of the appended KCS model in restricted waters.

Figure 7: General mesh assembly for the PMM simulations of the appended KCS model in restricted
waters.

11
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

The general mesh assembly used for the study is shown in Figure 7. Like previous cases, the domain is
generated using blockMeshDict, and after successive refinements, the hull and rudder are integrated into
the domain using snappyHexMesh. The propeller is modelled using the rotor disk model (rotor disk area
defined by propeller diameter and blade width) and no rotating mesh zone is included. The mesh
resolution used for the simulations is 3.8 million. Greater detail on cell size and resolution in the inner and
outer domains is discussed in the validation study section.

Unlike the open water cases, for PMM simulations in restricted water, an overset mesh topology is used
for all three types of imposed motions (drift, sway, and yaw) to ensure that the narrow channel effect is
experienced by the ship. In the open water case, the motion could be imposed on the entire domain for
sway and yaw cases since the assumption is that there is no interaction between the ship and the side
boundaries in open water conditions. However, in restricted water cases, the side boundaries are
modelled as walls and the interaction significantly influences the forces and moments experienced by the
ship.

2.3 Solver Numerical Model


For the study, the open-source CFD toolkit OpenFOAM is used. OpenFOAM is an open-source library that
numerically solves a wide range of problems in fluid dynamics from laminar to turbulent flows, with single
and multi-phases. The solver has been elaborately described by Weller (1998) and Jasak (2009). The
governing equations for the solver are the momentum or the Navier-Stokes equation for incompressible
two-phase flow (1) and the continuity equation (2). In vector form, the Navier-Stokes and continuity
equation are presented as:

 v 
  v.v   p   2v   g , (1)
 t 

.v  0 , (2)

where v is the velocity, p is the pressure,  is the dynamic viscosity, and g is the acceleration due to

gravity.

The volume of fluid (VOF) method is used to model the fluid as one continuum of mixed properties. This
method determines the fraction of each fluid that exists in each cell, thus tracking the free surface
elevation. The volume fraction is obtained by:

12
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947


 .( v)  0 (3)
t

with  is the volume fraction of water in the cell, varying from 0 to 1, with 0 representing a cell full of air
and 1 representing a cell full of water. The Finite Volume Method (FVM) is used to discretize the governing
equations. Pressure-velocity coupling is obtained through the PIMPLE (a combination of PISO and SIMPLE
algorithms) algorithm. All interpolation schemes in OpenFOAM, used for flux calculation in advection,
operate with the MULES (Multidimensional Universal Limiter for Explicit Solution) based solver.

For the present study, the OpenFOAM version 2106 managed by the ESI group is used. Simulations are
used using both single and overset domain configurations, and with single (propeller- open water) and
multi-phase flows (manoeuvring simulations). The overset approach used here is a generic
implementation for the use of disconnected meshes (Chimera framework). It is based on the inverse
distance interpolation scheme (Tisovska, 2019). In overset mesh, the governing partial differential
equation is solved both on the background (outer) and interior (inner) mesh. In the outer mesh, the cells
in the overlapping region between the outer and inner regions are marked as holes and are removed from
the computational domain. The values in the inner domain are calculated, and then the values at the
boundary of the two domains are determined through interpolation. So, information is transferred both
ways. This additional interpolation increases computational cost notably, however, it also allows
substantial motion of the solid body within the domain which mesh morphing can not accommodate.

The integrated wall function model in OpenFOAM is used to reduce the mesh resolution requirement near
the hull surface to capture the boundary layer. For turbulence modelling, the two-equation SST K-Omega
(Menter, 1993) turbulence model is used. For imposing the PMM motions, the solid body motion library
is used. For the open sea condition, static drift motion is imposed only on the overset mesh, whereas,
sway and yaw motions are imposed on the entire simulation domain. For the restricted water simulations,
all motions are imposed on the overset mesh. For pure yaw motion, a combined motion is defined in the
motion library, as described by Islam and Guedes Soares (2018).

To simulate the propeller wake, the rotor disk model in OpenFOAM (a body force model; Wahono (2013),
Patrao (2017)) was used. The rotor disk model is an improvement on the actuator disk mode, which was
implemented by Partao (2017). The model replaces the physical geometry of the propeller with only its
effect on the flow. It simulates the time-averaged flow over the propeller by adding the blade forces as
source terms in the momentum equation, which simulates or replicates the wake development. The
source term is only added to the governing equations for the cells in the mesh located inside the pre-

13
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

defined disk area. The simplified approach of the rotor disk model only captures the time-averaged effects
of the rotor and ignores flow separation at the blades and associated non-linear effects. However, the
compromise allows a substantial saving in computational costs.

For the simulations, two desktop computers with Intel i9 processors with 18 cores, 128 GB RAM, and a
2TB SSD disk are used. The simulations are performed using 12 processors in parallel, and the simulation
duration varies depending on both the mesh and motion configurations. Total simulation time varied for
each case depending on the forced motion type. However, all simulations are continued until steady
output (convergence) is reached. The physical simulation time for the open-water propeller cases is
roughly 11 hours. The overset simulations are immensely resource-consuming with the sway and yaw rate
cases taking up to 40 days of simulation time.

3 Results
In the study, manoeuvring coefficients are determined for an appended KCS model with a static rudder
and body force-based propeller model using PMM simulations. The propeller is modelled using the rotor
disk model in OpenFOAM. The rotor disk model uses propeller information, rpm, and drag and lift
coefficients to reproduce the rotating propeller effect in the simulation domain. The rudder is kept static
since a moving rudder would have required another overset domain, which would have increased
computational expense substantially. Furthermore, the interaction or interpolation among different
overset domains would have increased uncertainty. Thus, the simplifications are considered. All presented
manoeuvring simulations are performed with a captive hull, that is all degrees of freedom of motion are
restricted, and the ship experiences only the imposed motion. Only the simulation cases for verification
and validation study have heave and pitch-free motion, to replicate the experimental conditions.

The results section is divided into three subsections. The first subsection describes open-water simulations
of the propeller model are performed and validated against available experimental data. The second
section describes the PMM simulations performed for the open sea conditions, while the third one does
the same for the restricted channel condition.

3.1 Open Water Propeller test


The open water simulations are performed to assess OpenFOAM’s capability in modelling a rotating
propeller with thrust and torque. Simulations are performed for the KCS propeller model for varying
advance ratios (J). For the simulations, a model scale propeller with a 0.25 m diameter is used, maintaining
the same scale ratio (31.6) for the studied ship model. The simulations are performed using the single-

14
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

phase solver pimpleFOAM, with air density, a CFL (Courant–Friedrichs–Lewy) dependent time step, and a
maximum CFL value of 1. The average time step observed in the simulations is 0.00006s. All simulations
are run for 1s of simulation time for stable results and the results are later converted for water density.
In total, 10 simulations are performed for varying J values and the propeller thrust and torque are
measured, from which efficiency is later calculated. Drag and lift values are also recorded for later use in
disk modelling. The simulation results are shown in Table 2. A comparison between the simulation results
and the experiments (SIMMAN, 2020) is shown in Figure 8.

Table 2: OpenFOAM simulation results for the KCS propeller model in open water.

Thrust Torque
Advance ratio, J Thrust, T Torque, Q Efficiency
coefficient, Kt coefficient, Kq
0.1 0.1564 -0.006066 0.443639 -0.06883 0.103
0.2 0.1456 -0.00567 0.413004 -0.06433 0.204
0.3 0.1324 -0.0052 0.375561 -0.059 0.304
0.4 0.1176 -0.004689 0.33358 -0.0532 0.399
0.5 0.1028 -0.004137 0.291599 -0.04694 0.494
0.6 0.0841 -0.00356 0.238555 -0.04039 0.564
0.7 0.0658 -0.00294 0.186646 -0.03336 0.623
0.8 0.0458 -0.002276 0.129915 -0.02582 0.641
0.9 0.0241 -0.00155 0.068361 -0.01759 0.557
1 -0.000155 -0.000729 -0.00044 -0.00827 0.008

The comparison shows reasonable agreement between the CFD and EFD results, with notable deviation
in a few cases. The torque prediction shows good agreement with experiments except for the lowest
advance ratio. As for thrust, the agreement improves with increasing J. As for the efficiency, CFD results
in general show an underprediction of efficiency and the difference is most significant for the J value of
1.0. Nevertheless, the overall agreement is well acceptable for the present study. The drag and lift
coefficient results derived from the simulations were used in the next stage of the study to model the
rotor disk to properly represent the pressure field around a rotating propeller. Pressure distribution on
the lateral and longitudinal cross-section of the domain is shown in Figure 9. It also shows the velocity
contours around the propeller.

15
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

0.8
Kt_EFD 10 Kq_EFD Efficiency_EFD
0.7 Kt_CFD 10Kq_CFD Efficiency_CFD

0.6
Kt, 10Kq, and Efficiency

0.5

0.4

0.3

0.2

0.1

0
0.000 0.200 0.400 0.600 0.800 1.000 1.200
-0.1
Advance ratio, J

Figure 8: Comparison between CFD and EFD results for the open water propeller test for the KCS model.

Figure 9: Flow field visualization from the propeller open water simulation, at j= 0.5. The Figure shows the
velocity contours and pressure distribution on the side and the cross-plane of the propeller.

3.2 PMM simulation of appended KCS hull in open waters


For the open water PMM simulation cases, initially, an uncertainty or verification study is performed for
the overset grid structure, followed by a validation study for the open sea condition. Next, static drift
simulations are performed for the open sea condition with six drift angles using an overset mesh topology.

16
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

Following that pure sway and pure yaw simulations were performed, for five sway and yaw rates, using a
static mesh. Finally, manoeuvring derivatives are calculated from the simulation results. All simulations
are performed in ship design Froude number 0.26, thus a velocity of 2.196m/s in model scale.

3.2.1 Verification and Validation


For the verification and validation study in open sea or open water cases, calm water simulations are
performed at ship design Froude number (0.26) using an overset mesh for an appended KCS hull at zero
static drift and heave and pitch-free motion. An uncertainty study for the static mesh is not performed
since it has been discussed in previous work (Islam and Guedes Soares, 2018; 2020).

For the verification study, ITTC-2017 (2017) guidelines, namely the correction factor (Stern et al., 2001)
and factor of safety (Celik et al., 2008) based approaches are used. However, contrary to the independent
time and grid-based dependency study, a constant CFL-based approach is chosen (Islam and Guedes
Soares, 2021). Three different mesh resolutions are used for the study with varying time steps to keep the
CFL number constant. The mesh resolutions used in the study are presented in Table 3. Since the overset
mesh is used, refinement is performed in both the outer and inner mesh. For the refinement, the number
of cells in x, y, and z direction were increased by the mentioned refinement factors, for both block meshes
of outer and inner domains. The settings for refineMeshDict and snappyHexMesh were kept the same.
However, since the initial cell sizes were made smaller, all successive refinements by refineMeshDict and
snappyHexMesh followed the same refinement ratio, all the way to the layer thickness outside the hull.
Slight variations are observed in refinement ratios between inner and outer mesh due to the limitation
related to the number of whole divisions that could be applied. The y+ value for the finest mesh is 220,
for the medium refinement 285, and for the coarse mesh, it is 340. Considering the large number of cells
generated due to overset topology and associated computational cost, a higher resolution with a smaller
y+ value is not applied. Nevertheless, the medium mesh which is used in later studies remains below 300
and the coarse mesh also exceeds by a relatively small margin.

Table 3: Mesh resolutions used for the uncertainty assessment for the KCS model using overset mesh.

Outer Mesh Inner Mesh


Total
Refinement Min layer Refinement
Mesh mesh X Y Z X Y Z
ratio thickness ratio
(mil)
1 2.09 0.1660 0.1590 0.0640 1.00 0.0200 0.0188 0.0125 0.0047 1.00
2 3.85 0.1250 0.1250 0.0500 1.28 0.0156 0.0156 0.0100 0.0039 1.23
3 8.3 0.0960 0.0975 0.0380 1.32 0.0120 0.0121 0.0077 0.0030 1.30

17
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

The uncertainty assessment results are shown in Table 4. The study shows monotonous convergence for
all cases except for the roll moment. The correction factor-based approach suggests relatively higher
uncertainty for the sway and roll moment results. However, considering that the presented study is for
calm water with zero drift, both the observed sway force and yaw moment are small and somewhat
arbitrary. Thus, a large part of the uncertainty, in this case, comes from the overset topology and the
iteration between the inner and outer mesh. Nevertheless, overall, the predicted uncertainty using both
methods remains quite low and within an acceptable limit.

For validation, predicted surge drag resistance (Fx) is compared with the experimental data from the
Tokyo 2015 workshop results. The predicted drag resistance coefficient from CFD is 4.11E-3, whereas, the
predicted drag from the experiment is 3.711E-3. Thus, the CFD results show a deviation of 10%. However,
the experimental data presented in the Tokyo workshop is for the hull form without the rudder. As such,
a higher drag prediction by the CFD model is expected. Nevertheless, the application of overset mesh also
influences the results and part of the over-prediction may come from the interpolation error between the
inner and outer mesh. The observed deviation of 10% is higher than the predicted uncertainty, thus, the
total drag coefficient results remain unvalidated.

Table 4: Uncertainty assessment for the appended KCS model using overset mesh, in open waters.

Property Fx (N) Fy (N) Mx (N-m) Mz (N-m)


Ø1 (fine) -93.51 8 6.21 -35.11
Output values Ø2 (mid) -95.89 8.35 7.56 -36.79
Ø3 (coarse) -101.11 10.62 7.94 -40.81
r12 = h2/h1 1.31 1.31 1.31 1.31
Refinement ratio
r32 = h3/h2 1.28 1.28 1.28 1.28
Convergence ϵ21/ϵ32 0.456 0.154 3.553 0.418
Order of accuracy p 2.53 6.55 5.07 2.85
Grid convergence GCI21fine 0.0325 0.0112 0.09269 0.05161
index (GCI) GCI32fine 0.0784 0.0842 0.0252 0.1338
Corrected 𝑈𝑐1 0.6492% 0.2249% 1.8539% 1.0323%
uncertainty 𝑈𝑐2 1.5690% 1.6833% 0.5035% 2.6758%
Correction Factor based approach
Corrected 𝑈𝑐1 0.96% 5.21% NA 2.55%
uncertainty 𝑈𝑐2 2.25% 35.85% NA 6.41%

18
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

3.2.2 Static Drift Simulations


For static drift simulations, four different drift angles are considered. An overset mesh topology is used
since the rotor disk model used for the propeller does not allow angular implementation. Thus, the entire
inner domain including the rotor disk area is rotated to the desired drift angle to ensure proper modelling
of propeller action. Applying rigid body motion to the entire domain may have impacted the incoming
flow field in case of relatively large drift angles, thus, it is avoided. All simulations were performed in the
ship design Froude number (0.26). In the simulations, drift angles are varied from 0 to 12 degrees, with 3-
degree intervals. From the simulations, total surge and sway force, and roll and yaw moments are
measured. All simulations are performed with a static hull, and all degrees of freedom are restricted
except the imposed drift motion. An initial acceleration period is applied to ensure the gradual yaw of the
ship to the desired drift angle. The simulation results are shown in Table 5 and the force and moment time
histories are presented in Figure 10. From the sway force and yaw moment, the sway force coefficient
(F’y) and yaw moment coefficient (M’z) are also calculated. For the non-dimensionalization of the results,
the following equations have been used.

𝑓𝑦 (𝑁)
𝐹′𝑌 = (1)
𝜌 × 𝑣2 × 𝐿2𝑝𝑝
𝑚𝑧 (𝑁)
𝑀′𝑍 = (2)
𝜌 × 𝑣2 × 𝐿3𝑝𝑝
Table 5: Static drift simulation results for the appended KCS hull using OpenFOAM.

Drift
Fx (N) Fy tot (N) Mx (N-m) My (N-m) Mz (N-m) F'y (e-3) M'z (e-3)
Angle

0 -95.89 8.35 7.56 1.96 -36.79 3.27E-05 -1.98E-05


3 -117.51 157.7 -191.11 539.82 486.74 6.18E-04 2.62E-04
6 -195.85 417.15 -414.81 670.67 1005.48 1.63E-03 5.41E-04
9 -375.44 787.67 -707.03 1051 1703.14 3.09E-03 9.17E-04
12 -632.18 1135.72 -825.6 874.5 2330.19 4.45E-03 1.25E-03

19
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

Figure 10: Force and Moment time history from the static drift simulations of the appended KCS model in
open waters.

The coefficient results are also shown in Figure 11, together with a comparison with bare hull
experimental data (Kim et al, 2015). Compared to experimental data and the bare hull results presented
by Islam and Guedes Soares (2018) for the same hull and similar setup, the results show notable variation.
Compared to the bare hull results, both the sway force and yaw moment coefficients show higher values
for all cases. Considering the inclusion of the rudder, an increase in lateral force and yaw moment is
expected. Furthermore, the interaction with the propeller also influences the results. However, part of
the difference in prediction might also come from the overset topology. Flow field visualizations for the
drift simulations are shown in Figure 12 to demonstrate how the flow passes through a drifting ship and
the pressure distribution on the hull. The pressure range for all three drift cases has been kept the same
for easier comparison. The slight inconsistencies observed in the left images come from the meshing.
Since successive refinements were applied to the domain, the inconsistencies appear at the boundaries
where the grid becomes coarser. Considering that the regions are located relatively far from the hull, they
rarely influence the simulation output.

20
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

5.0E-03 1.4E-03
Lateral force coefficient, CFD

Yaw moment coefficient,


1.2E-03 CFD
4.0E-03 EFD (HHI)
1.0E-03 EFD (HHI)
EFD (CWNU)
3.0E-03 8.0E-04 EFD (CWNU)

M'z
F'y

6.0E-04
2.0E-03 4.0E-04
2.0E-04
1.0E-03
0.0E+00
0.0E+00 -2.0E-04
0 5 10 15 0 5 10 15
Drift angle (deg) Drift angle (deg)

Figure 11: Static drift simulation results for the appended KCS hull using OpenFOAM and bare hull
results from experimental studies.

Figure 12: Hydrodynamic pressure distribution on the free surface and on the KCS hull during open water
static drift simulations.

3.2.3 Pure Sway Simulations


For pure sway simulations, a static mesh is used where sway motion is applied on the entire simulation
domain. Since the sides of the domain are modelled as symmetry, the motion did not have any impact on
the flow pattern. Simulations are performed for the design Froude number with changing sway rates. To
change the sway rate, the sway frequency is kept constant at 0.5 rad/s, and the sway amplitude is varied
to represent the sway rate. The simulation results are shown in Table 6. For the calculation of coefficients,

21
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

equations 1 and 2 are used. The coefficients are calculated based on the peak values, whereas, the lateral
force and yaw moment shown are for the total maximum change of value (peak to trough). Figure 13
shows the time history for the lateral force and yaw moment as predicted by the simulations.

Table 6: Pure sway simulation results for the appended KCS hull using OpenFOAM.

Freq.
Amp. (m) vdot' Fx Fy Mz F'y (x e-3) M'z (x e-3)
(rad/s)
0.5 0.133 0.05 -87.99 125.86 314.58 0.25 0.08
0.5 0.398 0.15 -95.39 379.52 980.51 0.74 0.26
0.5 0.663 0.25 -104.75 695.67 1641.24 1.36 0.44
0.5 0.928 0.35 -114.1 1078.39 2365.69 2.11 0.64
0.5 1.193 0.45 -125.8 1577.325 3185.7 3.09 0.86

Figure 13: Force and Moment time history from pure sway simulations of the appended KCS model in
open waters.

Predicted coefficient results are also shown in Figure 14. As before, the results show notably larger force
and moment predictions for changing sway rate, compared to the bare hull case (Islam and Guedes
Soares, 2018). To further illustrate pure sway simulation, flow field visualization for the simulation with a
sway rate of 0.25 is shown in Figure 15. The figure shows how pressure on the free surface and hull form
changes with the forced sway motion.

22
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

3.50 1.00
Lateral Force Coefficient,
3.00 0.80

Coefficient, M'z
Yaw Moment
2.50
0.60
2.00
F'y

1.50 0.40
1.00
0.20
0.50
0.00 0.00
0 0.2 0.4 0.6 0 0.2 0.4 0.6
Sway Rate, Vdot' Sway Rate, Vdot'

Figure 14: Pure sway simulation results for the appended KCS hull using OpenFOAM.

Figure 15: Hydrodynamic pressure distribution on the free surface and on the KCS hull during open water
pure sway simulation at a sway rate of 0.25.

3.2.4 Pure Yaw Simulations


Finally, for pure yaw simulations, five cases are simulated with a static yaw angle, and varying frequency
and amplitude. Like before, all simulations are performed at ship design Froude number. The results for
the average drag force and maximum sway force and roll and yaw moment are shown in Table 7. As
before, the maximum values represent the summation of peak and trough values, except the drag which
is a mean value. The force and moment time histories are also shown in Figure 16. As can be seen, the
simulations are run for different periods depending on the applied yaw rate.

23
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

Table 7: Pure yaw simulation results for the appended KCS hull in open waters.

Fx-tot Fy (N) Mx (N-m) Mz (N-m)


r' Angle(deg) Omega (Hz) disp (m) rdot'
(N) Max Max Max

0.10 4 0.0688 0.3553 0.1432 -97.73 275.24 9948.00 1077.88

0.20 4 0.1376 0.1777 0.5730 -100.85 302.77 4804.16 1140.65

0.30 4 0.2063 0.1184 1.2892 -103.58 355.50 3177.35 1196.68

0.40 4 0.2751 0.0888 2.2918 -103.58 424.52 2370.88 1274.10

0.50 4 0.3439 0.0711 3.5810 -106.31 510.70 1899.19 1363.99

Figure 16: Surge and lateral force and yaw moment time history for the pure yaw simulations, with the
ship motion for yaw rate 0.30, performed for the appended KCS model in open waters.

However, for pure yaw simulations, the coefficients are calculated not from the maximum force or
moment. For varying yaw rate, lateral force and yaw moment data are collected at zero yaw angle, and
for varying yaw acceleration rate, lateral force and yaw moment are calculated at maximum yaw angle
(Lewis, 1988). The results are shown in Table 8. In the table, r’ is the yaw rate and the rdot’ is the yaw
acceleration rate. Comparing the previous OpenFOAM results (Islam and Guedes Soares, 2018), the yaw
results follow a similar trend as the sway results. However, the high deviation with experimental data
(From HHI and CNU) persists, especially for sway force coefficient results.

24
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

Table 8: Calculated yaw and yaw acceleration force and moment coefficients for the appended KCS hull
in open water using OpenFOAM.

r'
Fy_r' Mz_r' rdot' Fy_rdot' Mz_rdot'
0.10 0.108 0.0342 0.14 0.560 0.2665
0.20 0.253 0.0562 0.57 0.598 0.2786
0.30 0.396 0.0860 1.29 0.605 0.2847
0.40 0.555 0.1244 2.29 0.669 0.2795
0.50 0.707 0.1771 3.58 0.691 0.2864
The coefficient results are also represented graphically in Figure 17. The Mz_rdot’ results show an
anomaly for the rdot’ value of 2.29, which goes downwards, instead of upwards. This might be an
exceptional case where the conditions promote lower resistance. However, the issue might also come
from a numerical anomaly. Since it was limited to one point, further investigation was not performed for
the time being. Pressure distribution on the free surface and the hull form is presented in Figure 18, for
the simulation with a yaw rate of 0.30. The reflection observed at the inlet of the domain after 66s comes
from the yaw motion of the domain. However, the phenomenon is short-lived, and it hardly influences
the overall results.

0.8 0.20
Lateral Force Coefficient,

Yaw Moment Coefficient,

0.7
0.6 0.15
0.5
Mz_r'
Fy_r'

0.4 0.10
0.3
0.2 0.05
0.1
0.0 0.00
0.00 0.20 0.40 0.60 0.00 0.20 0.40 0.60
Yaw rate, r' Yaw rate, r'

0.75
0.34
Coefficient, Fy_rdot'

0.65
Yaw Moment Coefficient,

0.32
Lateral Force

0.55 0.30
0.28
Mz_rdot'

0.45
0.26
0.35 0.24
0.22
0.25
0.00 1.00 2.00 3.00 4.00 0.20
Yaw Acceleration rate, rdot' 0.00 1.00 2.00 3.00 4.00
Yaw Acceleration rate, rdot

Figure 17: Calculated yaw and yaw acceleration force and moment coefficients for the appended KCS hull
using OpenFOAM.

25
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

Figure 18: Hydrodynamic pressure distribution on the free surface and the KCS hull during open water
pure yaw simulation at a yaw rate of 0.30.

3.2.5 Manoeuvring Derivatives for Appended Hull


Finally, from the PMM simulation results, linear and non-linear hydrodynamic derivatives are calculated
to understand how the presence of appendages (static rudder and model propeller) can affect the hull
manoeuvring characteristics. Hydrodynamic derivatives or coefficients represent the rate of change in
force or moment, with the rate of change in velocity or acceleration. The value of hydrodynamic
derivatives helps to determine the stability and manoeuvrability characteristics of a ship. Thus, they are
determined in advance to ensure the proper functioning of a marine vehicle. Hydrodynamic coefficients
are also known as added mass and damping coefficients. The added mass coefficients are acceleration
dependent, whereas, the damping coefficients are velocity-dependent (Lewis, 1988).

26
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

By using curve fitting to the data for forces and moments as a function of drift angle β, sway rate v’max,
yaw rate r’max, and yaw acceleration rate r’dot(max); the hydrodynamic derivatives or coefficients, Xvv,
Yv, Yvvv, Nv, Nvvv, 𝑌𝑣̇ , 𝑁𝑣̇ , 𝑌𝑟 , 𝑁𝑟 , 𝑌𝑟̇ and 𝑁𝑟̇ can be predicted. Among the derivatives, Yv and Nv are linear,
and the rest are non-linear. Thus, for calculating the derivatives for open water simulations, the slopes of
the rate of change of force and moment were calculated against changing drift, sway, and yaw rates. The
results are shown in Table 9, including a comparison with bare hull CFD and EFD results reported by Kim
et al. (2015) and Islam and Guedes Soares (2018).

The linear velocity derivatives show notable disagreement with previous results, especially for the Yv.
However, following theoretical hydrodynamics, Yv is expected to have a relatively large “-ve” value. Thus,
the hull shows higher stability. As for Nv, a small positive or negative value is expected. So, the hull shows
better stability considering linear derivatives.

Next, the linear acceleration derivatives from pure sway results also show a notable difference from the
bare hull results. 𝑦𝑣̇ is expected to have a relatively large negative value, thus it shows lower stability.
Whereas, 𝑁𝑣̇ is expected to have a small positive or negative value, thus it shows better moment
acceleration stability.

Finally, the angular velocity and acceleration derivatives calculated from pure yaw results also show
notable disagreement, especially for the angular acceleration derivatives. The results suggest lower
stability or control in terms of angular velocity, and angular acceleration force. Whereas, improvement is
only observed for the yaw added mass moment of inertia.

Table 9: Hydrodynamic derivatives of the KCS hull model calculated from PMM simulations.

Derivatives CFD_appended hull CFD_bare hull EFD_HHI EFD_CWNU


Static Yv -0.30 0.14 0.14 0.12
Drift Nv -0.05 0.06 0.07 0.05
Pure 𝑦𝑣̇ -0.25 -4.10 -4.20 -3.80
Sway 𝑁𝑣̇ -0.02 -2.10 - -
𝑌𝑟 -0.05 -0.10 -0.03 -0.09
𝑁𝑟 -0.01 -0.02 -0.06 -0.10
Pure Yaw
𝑦𝑟̇ 0.57 0.04 0.08 0.16
𝑁𝑟̇ 0.27 0.01 0.06 0.09

Overall, the appended hull shows lower linear acceleration, angular velocity, and angular acceleration
stability comparing to a bare hull. The study suggests that the inclusion of appendages, like the rudder

27
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

and hull, should give more reliable manoeuvring characteristics of the ship and the belief that
manoeuvring derivatives are strictly a hull property may not be justified.

One limitation of the work might be the missing wave impact. Ocean-going vessels spend most of their
lifetime at sea with waves, while manoeuvring assessments are mostly performed in calm waters. Thus,
the mismatch needs to be addressed. Furthermore, recent SIMMAN workshops are also encouraging free-
running manoeuvring studies with active rudder and propeller and free ship motions, which mimic the
operating condition. Thus, future studies will focus on these areas to further enhance understanding of
the topic.

3.3 PMM simulation of appended hull form in restricted waters


For the restricted channel cases, the same set of PMM simulations are performed and derivatives are
predicted as described in section 3.2. However, unlike open waters, restricted channels do not allow large
manoeuvring motion. As such, relatively small drift angles, and sway and yaw rates were imposed for the
study. Following channel restrictions, the simulations were performed at a reduced speed of 1.284 m/s in
model scale (Fr 0.152) or 14 knots in full scale. The simulated speed is somewhat higher than the generally
permitted speed in the canal, however, a higher speed was used to assess the extreme case. The rotor
disk model was also adjusted to match the desired thrust and torque at the imposed ship speed. The PMM
simulations were performed with a captive hull, with all degrees of freedom restricted. However, for the
calm water simulations used for the uncertainty study, heave and pitch motions were kept free. To limit
computational cost, three cases were simulated for each forced motion.

3.3.1 Verification Study


Although the mesh topology for overset meshing is similar to the previous section 3.2, a verification or
grid uncertainty assessment was repeated for the restricted water cases to observe the additional
uncertainty introduced due to shallow water and channel side effects (if any). For the study, three
different grid resolutions were used, and refinements were performed for both the inner and outer
domains. The mesh resolutions used for the study are shown in Table 10. The y+ value for the three mesh
resolutions are 130, 175, and 210, respectively.

The Uncertainty assessment was performed using the factor of safety-based approach, following the ITTC-
2017 guidelines. However, as before, a constant CFL-based approach was taken with time step adjustment
along with grid refinement. For the study, surge and sway force, and roll and yaw moment results were

28
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

considered. The uncertainty assessment results are shown in Table 11. The simulations were performed
at Froude number 0.152, with a heave and pitch free motion.

Table 10: Mesh resolutions used for the grid dependency study for the simulation of appended KCS hull
in restricted waters.

Total Min
Refinement Refinement
Mesh mesh X Y Z X Y Z layer
ratio ratio
(mil) thickness
1 8.24 0.1000 0.0940 0.0310 1.00 0.0120 0.0117 0.0090 0.0029 1.00
2 3.78 0.1250 0.1250 0.0400 1.29 0.0156 0.0156 0.0117 0.0039 1.33
3 1.97 0.1500 0.1500 0.0520 1.30 0.0196 0.0188 0.0160 0.0047 1.37

Table 11: Uncertainty assessment for the zero drift simulation cases for the appended KCS model in
restricted water.

Property Fx (N) Fy (N) Mx (N-m) Mz (N-m)


Ø1 (fine) -56.91 1.23 4.1 -16.65
Output values Ø2 (mid) -51.49 2.64 4.25 -22.22
Ø3 (coarse) -53.3 2.21 4.22 -17.88
Convergence ϵ21/ϵ32 -2.994 -3.279 -5.000 -1.283
Order of
p 0.85 0.92 1.25 0.19
accuracy
Grid
GCI21fine -0.4926 5.4280 0.12202 8.43565
convergence
index (GCI) GCI32fine 0.1759 -0.7458 -0.0227 -4.7767
Corrected 𝑈𝑐1 9.8527% 108.5605% 2.4404% 168.7130%
uncertainty 𝑈𝑐2 3.5176% -14.9157% -0.4547% -95.5340%

The results show oscillatory convergence and low-order accuracy for all four measurements. Due to the
relatively high speed at shallow depths, the solutions become notably more sensitive to mesh distribution
and resolution. As such, the coarse mesh resolution somehow provides a higher estimation of forces and
moments compared to the mid mesh. The overall uncertainty remains substantial for the lateral force and
yaw moment. However, the simulated case is a zero-drift calm water case. As such, these two values are
somewhat arbitrary and are slightly influenced by the channel width and draft. For total drag resistance,
the uncertainty remains at around 10%. A higher mesh resolution for the coarse mesh might have
improved the estimation. However, considering the sophistications involved in the simulations including
a restricted channel, overset mesh, the propeller model, and turbulence effects, there is no assurance
that a higher resolution would have resulted in a monotonous convergence. Thus, it was not attempted.

29
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

Unfortunately, validation data is not available for the presented case setup. As such, direct validation
could not be performed, and verification could not be ascertained.

3.3.2 Static Drift Simulations


For the static drift simulations, a total of four cases are considered starting from 0 to 4 drift, with an
interval of 1. For all cases, the rudder is kept static and gradual drift motion is imposed on the overset
mesh including the hull, rudder, and propeller rotor disk model. The simulations are performed at Froude
0.152, with a channel depth-to-ship draft ration of 2, and width of 6 m. The simulation results are shown
in Table 12. The results include maximum force and moment predictions and calculated lateral force and
yaw moment coefficient. For the non-dimensionalization of results, equation 1 and 2 was used. The time
history for the sway force and yaw moment are shown in Figure 19.

Table 12: Static drift simulation results for the appended KCS hull in restricted waters.

Drift Angle Fx vis (N) Fx tot (N) Fy tot (N) Mx (N-m) My (N-m) Mz (N-m) F'y (e-3) M'z (e-3)
0 -34.8 -54.16 1.14 4.01 1252 -11.9 0.013 0.019
1 -35.65 -67.54 43.83 -109.06 1409.4 122.02 0.548 0.192
2 -37.8 -72.54 89.81 -218.35 1389.4 274 1.029 0.432
3 -39.51 -84.8 140.26 -359.12 1500 418.7 1.608 0.659

Figure 19: Lateral force and yaw moment time history for static drift simulation in restricted water
conditions for the KCS model.

30
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

2.0 0.7
Lateral Force Coefficient, F'y

Yaw Moment Coefficient,


0.6
1.5 0.5

M'z (e-3)
0.4
(e-3)

1.0 0.3
0.2
0.5
0.1
0.0
0.0
0 1 2 3 4
0 1 2 3 4 Drift Angle (deg)
Drift Angle (deg)

Figure 20: Lateral force and yaw moment coefficient results for the KCS appended hull model against
varying drift angles.

The lateral force and yaw moment coefficient results are graphically represented in Figure 20. As can be
observed from the results, even a small increase in drift angle increases the experienced force and
moment substantially. This indicates that the ship requires notably more power to manoeuvre in
restricted conditions. Hydrodynamic pressure distribution on the water and hull surface during drift
simulations is also shown in Figure 21. For easy comparison, the figure has the same pressure range as for
open water drift cases (Figure 12).

Figure 21: Hydrodynamic pressure distribution on the free surface and on the KCS hull during restricted
water static drift simulations.

31
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

3.3.3 Pure Sway Simulations


For pure sway, three simulations are performed with varying sway rates. The sway simulations are also
performed in a channel with a width of 6 m and a depth-to-ship draft ratio of 2. The simulation results are
shown in Table 13, and the time history from the simulations is shown in Figure 22. In the Table, vdot’
represents the sway rate. The applied amplitude and frequency determine the sway rate. Surge and sway
force, and roll, pitch, and yaw moment results are presented, together with lateral force and yaw moment
coefficients. The oscillations observed in the time history are mostly due to the sophistication involved in
the simulation. Due to the shallow draft, the simulation CFL number witnessed high oscillation, as such,
the time step fluctuated as well. Thus, a smooth convergence was not observed, like in the open water
cases. Ship interaction with the diffracted waves creates the low-frequency responses observed.

Figure 22: Force and moment time history for the pure sway simulations of the KCS model in restricted
waters.

32
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

Table 13: Pure sway simulation results for the appended KCS hull in restricted waters.

Freq. Amp.
vdot' Fx (N) Fy (N) Mx (N-m) My (N-m) Mz (N-m) F'y (e-3) M'z (e-3)
(rad/s) (m)
0 0.000 0 -54.16 1.14 4.01 1252 -11.9 0.007 -0.009

0.5 0.136 0.15 -63.29 340.285 4809.36 1345.63 593.48 1.950 0.467

0.5 0.227 0.25 -60.2 468.15 8064.12 1199.26 833.4 2.683 0.656

0.5 0.317 0.35 -65.41 607.23 11285.95 1360.78 1179.2 3.480 0.929

Following the results, the coefficients show a steady rise in lateral force and yaw moment with increasing
sway rate. However, they show a significant rise in the value of forces and moments compared to the
relatively small change in sway rate. One interesting result from the study is the pitch moment, which
decreases the sway rate by 0.25. The pitch moment is observed since the pitch motion was restricted in
the simulations and the shallow draft was building pressure at the bow. The unexpected result for the
sway rate of 0.25 has not been further investigated in this work. The coefficients are graphically
represented in Figure 23. Flow field visualization for the sway rate of 0.25 is also shown in Figure 24. The
figure also shows how pressure at the side of the domain changes with changing sway positions.

4.0 1.0
Yaw Moment Coefficient,
Lateral Force Coefficient,

3.5
0.8
3.0
M'z (e-3)

2.5 0.6
F'y (e-3)

2.0
1.5 0.4
1.0
0.2
0.5
0.0 0.0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
Sway Rate, Vdot' Sway Rate, Vdot'

Figure 23: Lateral force and yaw moment coefficients for the appended KCS hull from pure sway
simulations in restricted waters.

33
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

Figure 24: Hydrodynamic pressure distribution on the free surface, sides, and on the KCS hull during
restricted water pure sway simulation at sway rate 0.25.

3.3.4 Pure Yaw Simulations


Finally, pure yaw simulations are performed with three different yaw rates. Due to the channel geometry,
the yaw angles are restricted to 2 and the frequency is increased to match desired yaw and yaw
acceleration rates. The average drags and maximum sway forces and moments encountered by the hull
form during yaw motions are shown in Table 14. As before, the simulations are performed in the restricted
channel with a depth-to-draft ratio of 2, and a width of 6m. The results show that total drag increases
notably due to yaw motion. However, for the yaw rate of 0.3, the drag somehow reduces. As for the roll
moment, it decreases with increasing yaw rate, which is expected due to the yaw motion. The simulation
time histories are shown in Figure 25. The time history shows secondary interactions which come from
the diffracted ship wake waves from the channel side and bottom that interacted with the hull.

34
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

Figure 25: Force and moment time history for the pure yaw simulations of the KCS model in restricted
waters.

The sway force and yaw moment coefficient results for varying yaw rate and yaw acceleration rate are
shown in Table 15. The results are also represented graphically in Figure 26. The results show a steep rise
in both lateral force and yaw moment with changing yaw rates. For the changing yaw acceleration rate,
the change in force and moment is relatively gradual. However, for the yaw moment at a yaw acceleration
rate of 2.58, a decreased value is observed. To further illustrate the results, pressure distribution on the
free surface and the hull surface is shown in Figure 27, for a yaw rate of 0.30.

Table 14: Pure yaw simulation results for the appended KCS hull in restricted waters.

Fy (N) Mx (N-m) Mz (N-m)


r' Angle(deg) Omega (Hz) Disp (m) rdot' Fx-tot (N)
Max Max Max
0.20 2 0.1609 0.0444 1.1459 -70.31 249.82 1722.70 753.46
0.30 2 0.2413 0.0296 2.5783 -75.41 268.61 1193.84 793.70
0.40 2 0.3217 0.0222 4.5837 -72.96 304.89 866.66 926.93

35
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

0.8 0.20

Yaw moment coefficient,


Lateral force coefficient,

0.7
0.6 0.15

Mz_r' (e-3)
Fy_r' (e-3)

0.5
0.4 0.10
0.3
0.2 0.05
0.1
0.0 0.00
0.0 0.2 0.4 0.6 0.0 0.2 0.4 0.6
Yaw rate, r' Yaw rate, r'

1.2 0.7
Lateral force coefficient,

Yaw moment coefficient,


1.0 0.6

Mz_rdot' (e-3)
0.5
Fy_rdot' (e-3)

0.8
0.4
0.6
0.3
0.4 0.2
0.2 0.1
0.0 0.0
0.0 2.0 4.0 6.0 0.0 2.0 4.0 6.0
Yaw acceleration rate, rdot' Yaw acceleration rate, rdot'

Figure 26: Lateral force and yaw moment coefficients for the appended KCS hull from pure yaw
simulations in restricted waters.

Table 15: Lateral force and yaw moment coefficients predicted from pure yaw simulations for the
appended KCS hull model in restricted waters.
Yaw acceleration
Yaw rate, r' Fy_r' (x e-3) Mz_r' (x e-3) rate, rdot' Fy_rdot' (x e-3) Mz_rdot' (x e-3)
0.20 0.06 0.06 1.15 0.82 0.54
0.30 0.31 0.10 2.58 0.96 0.44
0.40 0.69 0.17 4.58 1.14 0.60

36
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

Figure 27: Hydrodynamic pressure distribution on the free surface, sides, and on the KCS hull during
restricted water pure yaw simulation at a yaw rate of 0.30.

3.3.5 Manoeuvring derivatives for the appended hull in restricted waters


As before (section 3.2.5), from the simulation results, hydrodynamic derivatives were predicted for the
appended hull form determining the rate of change in force and moment, with changing drift, sway, and
yaw rates. The results are shown in Table 16, including a comparison with open-water appended hull
results.

Following theoretical hydrodynamics, considering the linear velocity derivatives, the appended hull in
restricted waters shows lower lateral stability, whereas, the yaw moment improves here. Next, for the
linear acceleration derivatives, 𝑦𝑣̇ , shows a large positive value, which indicates large instability. This is
justified due to the channel restriction on the sides. However, the value is substantially larger compared
to the open water cases and shows large instability. As for 𝑁𝑣̇ , it is also a relatively large value indicating
instability in the yaw moment. As for the angular velocity, both parameters are relatively large, showing
higher instability compared to the open water case. The situation is the same for yaw acceleration
derivatives.

37
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

Table 16: Hydrodynamic derivatives of the KCS hull model with static rudder and modelled propeller
calculated using PMM simulations.

CFD_appended CFD_appended
Derivatives
hull_RestrictedWater hull_OpenWater
Yv -0.009 -0.30
Static Drift
Nv 0.001 -0.05
𝑦𝑣̇ 0.792 -0.25
Pure Sway
𝑁𝑣̇ 0.223 -0.02
𝑌𝑟 -0.590 -0.05
𝑁𝑟 -0.051 -0.01
Pure Yaw
𝑦𝑟̇ 0.715 0.57
𝑁𝑟̇ 0.471 0.27

Overall, the results suggest that the manoeuvring capabilities of a hull form change substantially while
passing through restricted channels. As mentioned before, the applied ship velocity for the ship was
higher compared to the generally allowed vessel speed in channels. Thus, with a lower speed, the ship
might show higher stability. However, the lower speed also causes a slower rudder response decreasing
ship manoeuvrability. As such, much care and attention are needed while passing such large vessels
through restricted channels to ensure safe manoeuvring.

4 Conclusion
The paper presents manoeuvring simulations for a KCS hull with a static rudder and a body force model-
based propeller in both open and restricted waters. Initially, open-water propeller simulations are
performed to generate information for the body force model, which is later used in the PMM simulations.
Static drift, pure sway, and pure yaw simulations are for varying conditions; and from the simulation
results, hydrodynamic derivatives are predicted. The predicted derivatives for open water are compared
against bare hull simulation and experimental results, and restricted water derivatives are compared
against the open water ones.

The open water manoeuvring simulation results suggest that due to the presence of a rudder and
propeller, the ship encounters larger lateral force and yaw moment, which impacts its manoeuvring
behaviour. Following the derivatives, the appended hull shows lower linear acceleration, angular velocity
and acceleration stability compared to the bare hull. Thus, manoeuvring derivatives may not just be a hull

38
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

property, and the inclusion of appendages (rudder and propeller) should provide a better estimation of
the ship’s manoeuvring capabilities compared to a bare hull investigation.

The restricted water simulations show that ship resistance increases substantially while passing through
restricted waters or confined channels. Compared to the open water case, following the comparison of
manoeuvring derivatives, the ship requires notably more effort to remain stable and manoeuvre in
restricted waters. The bank effect gives the ship a yaw moment, whereas, the shallow draft initiates a
pitch moment. While lower forward speed might make ships more stable in restricted waters, it will also
make the ship more difficult to manoeuvre. Thus, a balanced approach with sincere attention is needed
while manoeuvring large ships in restricted waters.

Overall, the paper provides quantitative information on how the ship manoeuvring capabilities are
influenced based on the model consideration and environmental conditions. Thus, it helps improve the
understanding of ship manoeuvring characteristics and how to improve it by optimizing designs for all
involved components. The paper also confirms that OpenFOAM can prove to be a useful and reliable tool
for such sophisticated studies.

Acknowledgment
The work was performed within the NAVAD project “Simulation of manoeuvrability of ships in adverse
weather conditions” which is co-funded by the European Regional Development Fund (Fundo Europeu de
Desenvolvimento Regional - FEDER) and by the Portuguese Foundation for Science and Technology
(Fundação para a Ciência e a Tecnologia - FCT) under contract 02/SAICT/032037/2017. This work
contributes to the Strategic Research Plan of the Centre for Marine Technology and Ocean Engineering
(CENTEC), which is financed by the Portuguese Foundation for Science and Technology (Fundação para a
Ciência e Tecnologia - FCT) under contract UIDB/UIDP/00134/2020.

References
Abkowitz, M.A., 1964. Lecture on Ship Hydrodynamics: Steering and Manoeuvrability, HyA Report. No.
Hy-5, Lyngby, Denmark.

Abkowitz, A.M. Measurement of Hydrodynamic Characteristics from Ship Maneuvering Trials by System
Identification. 1980. Available online: https://trid.trb.org/view/157366 (accessed on August 2022).

Abkowitz, M.A., Liu, G., 1988. Measurement of ship resistance, powering and manoeuvring coefficients
from simple trials during a regular voyage. Trans. Soc. Nav. Archit. Mar. Eng., 96: 97–128.

39
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

Araki, M., Sadat-Hosseini, H., Sanada, Y., Tanimoto, K., Umeda, N., Stern, F., 2012. Estimating maneuvering
coefficients using system identification methods with experimental, system-based, and CFD free-running
trial data. Ocean Engineering, 51: 63–84.

Bonci, M.; Viviani, M.; Broglia, R.; Dubbioso, G., 2015. Method for estimating parameters of practical ship
manoeuvring models based on the combination of RANSE computations and System Identification.
Applied Ocean Research, 52: 274-294.

Broglia, R., Dubbioso, G., Durante, D., 2013. Simulation of turning circle by CFD: Analysis of different
propeller models and their effect on manoeuvring prediction. Appl. Ocean Res., 39: 1–10.

Broglia, R., Dubbioso, G., Durante, D., Di Mascio, A., 2015. Turning ability analysis of a fully appended twin
screw vessel by CFD. Part I: single rudder configuration. Ocean Engineering105: 275–286.

Broglia, R., Muscari, R., Mascio, A.D., 2008. Numerical simulations of the pure sway and pure yaw motion
of the KVLCC1 and 2 tankers. In: SIMMAN 2008 Proceedings, pp. F2–F9.

Capitao P.A., 2017. Description and validation of the rotorDiskSource class for propeller performance
estimation. In Proceedings of CFD with OpenSource Software, 2017, Edited by Nilsson. H.

Carrica, P.M., Mofidi, A., Eloot, K., Delefortrie, G., 2016. Direct simulation and experimental study of zigzag
maneuver of KCS in shallow water. Ocean Engineering112: 117–133.

Celik, I. B., Ghia, U., Roache, P. J., Freitas, C. J., Coleman, H., Raad P. E., 2008. Procedure for Estimation
and Reporting of Uncertainty Due to Discretization in CFD Applications. Journal of Fluids Engineering,
130(7): 078001.

Cura-Hochbaum, A., 2006. In: Virtual PMM Tests for Manoeuvring Prediciton. 26th Symposium on Naval
Hydrodynamics, pp. 23-28, Rome, Italy.

Cura-Hochbaum, A., Vogt, M., Gatchell, S., 2008. Manoeuvring prediction for two tankers based on RANS
simulations. In: SIMMAN 2008 Proceedings, pp. F22–F27.

Diez, M., Serani, A., Campana, E.F., Stern, F., 2022. Time-series forecasting of ships maneuvering in waves
via dynamic mode decomposition. J. Ocean Eng. Mar. Energy (2022). https://doi.org/10.1007/s40722-
022-00243-0

Dubbioso, G., Durante, D., Di Mascio, A., Broglia, R., 2016. Turning ability analysis of a fully appended twin
screw vessel by CFD. Part II: single vs. twin rudder configuration. Ocean Engineering117: 259–271.

Franceschi, A., Piaggio, B., Tonelli, R., Villa, D., Viviani, M., 2021. Assessment of the Manoeuvrability
Characteristics of a Twin Shaft Naval Vessel Using an Open-Source CFD Code. J. Mar. Sci. Eng., 9: 665.

Frisch, U., 1995. Turbulence: The Legacy of A.N. Kolmogorov, 1st ed., Cambridge University Press:
Cambridge, UK.

Gadelho, J.F.M., Rodrigues, J.M., Lavrov, A., Guedes Soares, C., 2018. Heave and sway hydrodynamic
coefficients of ship hull sections in deep and shallow water using Navier-Stokes equations. Ocean
Engineering, 154: 262-276.

40
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

Gaggero, S., Villa, D., Viviani, M., 2017. An extensive analysis of numerical ship self-propulsion prediction
via a coupled BEM/RANS approach. Applied Ocean Research, 66: 55-78.

Guedes Soares, C., Sutulo, S., Francisco, R. A., Santos, F. M., and Moreira, L., 1999. Full-Scale
Measurements of the Manoeuvring Capabilities of a Catamaran. Proceedings of the International
Conference on Hydrodynamics of High Speed Craft; London. 1-12.

Guedes Soares, C., Francisco, R. A., Moreira, L., Laranjinha, M., 2004. Full-Scale Measurements of the
Manoeuvering Capabilities of Fast Patrol Vessels, Argos Class. Marine Technology. 41(1): 7-16.

Gullmineau, E., Queutey, P., Visonneau, M., Leroyer, A., Deng, G., 2008. RANS simulation of a US Navy
frigate with PMM motions. In: SIMMAN 2008 Proceedings, pp.F28–F33

Hajivand, A., Mousavizadegan, S.H., 2015. Virtual maneuvering test in CFD media in presence of free
surface. Int. J. Nav. Architect. Ocean. Engineering7: 540–558.

Hasanvand, A., Hajivand, A., 2019. Investigating the effect of rudder profile on 6DOF ship turning
performance. Appl. Ocean Res. 92: 101918.

Hasanvand, A., Hajivand, A., Ali, N.A., 2021. Investigating the effect of rudder profile on 6DOF ship course-
changing performance. Appl. Ocean Res. 117: 102944.

Ino, T., Stern, F., Larsson, L., Visonneau, M., Hirata, N., Kim, J. Numerical Ship Hydrodynamics An
Assessment of the Tokyo 2015 Workshop: An Assessment of the Tokyo 2015 Workshop. Springer, 2021.

Islam, H., Guedes Soares C. 2019.Uncertainty analysis in ship resistance prediction using OpenFOAM,
Ocean Engineering, 191:105805,

Islam, H., Guedes Soares, C., 2021. Assessment of uncertainty in the CFD simulation of the wave-induced
loads on a vertical cylinder. Mar. Struct., 80: 103088.

ITTC, 2002. The Specialist Committee On Esso Osaka - Final Report and Recommendations to the 23rd
ITTC.

ITTC-2011. ITTC- Recommended Procedures and Guidelines: Practical Guidelines for Ship CFD Application.
7.5-03-02-03.

ITTC-2017 – Recommended Procedures and Guidelines. Uncertainty Analysis in CFD Verification and
Validation Methodology and Procedures. ITTC – 7.5-03-01-01 (2017).

Jasak, H., 2009. OpenFOAM: Open Source CFD in research and industry. International Journal of Naval
Architecture and Ocean Engineering, 1(2):89-94.

Kawamura, T., Miyata, H., Mashimo, K., 1997. Numerical simulation of the flow about self-propelling
tanker models. J. Mar. Sci. Technol., 2: 245–256.

Kijima, K., Nakiri, Y., 2003. On the practical prediction method for ship manoeuvering characteristics.
Transaction of the West Japan Society of Naval Architects 105: 21–31.

41
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

Kim D., Tezdogan T., Incecik A., 2022a. A high-fidelity CFD-based model for the prediction of ship
manoeuvrability in currents. Ocean Engineering, 256:111492.

Kim, D., Tezdogan, T., Incecik, A., 2022b. Hydrodynamic analysis of ship manoeuvrability in shallow water
using high-fidelity URANS computations. Appl. Ocean Res. 123: 103176.

Kim H., Akimoto H., Islam, H., 2015. Estimation of the hydrodynamic derivatives by RaNS simulation of
planar motion mechanism test. Ocean Engineering, 108: 129-139.

Kim, I.-T., Kim, C., Kim, S.-H., Ko, D., Moon, S.-H., Park, H., Kwon, J., Jin, B., 2021. Estimation of the
manoeuvrability of the KVLCC2 in calm water using free running simulation based on CFD. Int. J. Nav.
Archit. Ocean Eng., 13: 466-477.

Kose, K., Misiag, W., Zhu J., Hirao, S., 1992. Database System Approach for Maneuvering Performance
Prediction. Journal of The Society of Naval Architects of Japan, 172: 375-382.

Labanti, J., Islam, H., Guedes Soares, C., 2016. CFD assessment of Ropax hull resistance with various initial
drafts and trim angles. In Maritime Technology and Engineering 3, Guedes Soares, C., Santos, T.A., Eds.,
Taylor & Francis Group: London, UK, 2016, pp. 325–332.

Lewis, E., 1988. Principles of Naval Architecture. Jersey City, NJ: The Society of Naval Architects and Marine
Engineers.

Liu, Y., Zou, Z., Zou, L., Fan, S., 2019. CFD-based numerical simulation of pure sway tests in shallow water
towing tank. Ocean Engineering, 189: 106311.

Menter, F.R., 1993. Zonal Two Equation k-ω Turbulence Models for Aerodynamic Flows. In Proceedings of
the 23rd Fluid Dynamics, Plasmadynamics, and Lasers Conference, Orlando, FL, USA, 6–9 July 1993, AIAA
Paper: Virginia, USA, 1993, p. 2906.

Miller, R.W., 2008. PMM calculation for the bare and appended DTMB 5415 using the RaNS solver
CFDSHIP-Iowa. In: SIMMAN 2008 Proceedings.

Ogawa, A., Koyama, T., Kijima, K., 1977. MMG report-I, on the mathematical model of ship manoeuvring.
Bull Soc. Nav. Archit., 575: 22–28.

Papanikolaou, A., Zaraphonitis, G., Bitner-Gregersen, E., Shigunov, V. , El Moctar, O., Guedes Soares, C.,
Reddy, D.N., Sprenger, F., 2016. Energy efficient safe Ship OPERAtion (SHOPERA). Transportation Research
Procedia, 14(2016): 820-829.

Sanada, Y., Elshiekh, H., Toda, Y., Stern, F., 2019. ONR Tumblehome course keeping and maneuvering in
calm water and waves. J Mar Sci Technol, 24: 948–967.

Sanada, Y., Park, S., Kim, D-H., Wang, Z., Stern, F., Yasukawa, H., 2021. Experimental and computational
study of hull–propeller–rudder interaction for steady turning circles. Physics of Fluids, 33: 127117.

Shang H., Zhan C., Liu Z., 2021. Numerical Simulation of Ship Maneuvers through Self-Propulsion. J. Mar.
Sci. Engineering, 9: 1017.

42
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

Shen, Z., Wan, D., Carrica, P.M., 2015. Dynamic overset grids in OpenFOAM with application to KCS self-
propulsion and maneuvering. Ocean Engineering 108: 287–306.

SIMMAN 2008 Workshop. [Online] Available at: http://www.simman2008.dk/[Accessed July 2022].

SIMMAN 2014 Workshop. [Online] Available at: https://simman2014.dk/ship-data/moeri-container-


ship/geometry-and-conditions-moeri-container-ship/ [Accessed July 2022].

SIMMAN 2020 Workshop. [Online] Available at: https://simman2020.kr [Accessed July 2022].

Simonsen, C., Stern, F., 2005. RANS maneouvering simulation of Esso Osaka with rudder and a body-force
propeller. J. Ship Res. 49 (2): 98–120.

Simonsen, C.D., Otzen, J.F., Klimt, C., Larsen, N.L., 2012. Manoeuvring predictions in the early design phase
using CFD generated PMM data. In: 29th Symposium on Naval Hydrodynamics, Gothenburg.

Stern F, Wilson RV, Coleman HW, Paterson EG. Comprehensive Approach to Verification and validation of
CFD Simulations—Part 1: Methodology and Procedures. ASME Journal of Fluids Engineering, 123(4): 793–
802.

Suez Canal, Egypt (2020). [Online] Available at: https://earth.esa.int/web/earth-watching/image-of-the-


week/content/-/article/suez-canal-
egypt/#:~:text=After%20several%20enlargements%2C%20it%20is,and%20the%20Great%20Bitter%20La
ke [Accessed July 2022].

Sutulo, S., and Guedes Soares, C., 2004. Synthesis of Experimental Designs of Manouevring Captive-Model
Tests with Large Number of Factors. Journal of Marine Science and Technology. 9(1)32-42.

Sutulo, S., Guedes Soares, C., 2014. An algorithm for offline identification of ship manoeuvring
mathematical models from free-running tests. Ocean Engineering, 79: 10–25.

Sutulo, S., and Guedes Soares, C., 2015. Development of a core mathematical model for arbitrary
manoeuvres of a shuttle tanker. Applied Ocean Research. 51293-308.

Sutulo, S., and Guedes Soares, C., 2019. On the application of empiric methods for prediction of ship
manoeuvring properties and associated uncertainties. Ocean Engineering. 186106111.

Sutulo, S., Moreira, L., and Guedes Soares, C. 2002. Mathematical Models for Ship Path Prediction in
Manoeuvring Simulation Systems. Ocean Engineering. 29(1)1-19.

Tisovska, P., 2019. Description of the overset mesh approach in ESI version of OpenFOAM. In Proceedings
of CFD with OpenSource Software, 2019, Edited by Nilsson. H.

Tokyo 2015 Workshop. [Online] Available at: https://www.t2015.nmri.go.jp/kcs.html [Accessed July


2022].

Wahono, S., 2013. Development of Virtual Blade Model for Modelling Helicopter Rotor Downwash in
OpenFOAM. Report no: DSTO-TR-2931, Australian Government Department of Defence.

43
December 2022
Ocean Engineering 266(3):112947 DOI: 10.1016/j.oceaneng.2022.112947

Wang, H.M., Xie, Y., Liu, J.M., Zou, Z.J., He, W., 2011. Experimental and numerical researches on the
viscosity hydrodynamic in hydrodynamic forces acting on a KVLCC2 model in oblique motion. In:
Proceedings of the International Conference in Remote Sensing, Environment and Transportation
Engineering (RSETE), pp. 328–331.

Weller, H.G., Tabor, G., Jasak, H., Fureby, C., 1998. A tensorial approach to computational continuum
mechanics using object-oriented techniques. Comput. Phys., 12: 620–631.

Wilson, R., Carrica, P., Stern, F., 2006. Unsteady RANS method for ship motions with application to roll for
a surface combatant. Comput. Fluids 35(5): 501–524.

Xu, H. T., and Guedes Soares, C., 2020. Manoeuvring Modelling of a Containership in Shallow Water based
on Optimal Truncated Nonlinear Kernel-based Least Square Support Vector Machine and Quantum-
Inspired Evolutionary Algorithm. Ocean Engineering, 195:106676.

Xu, H. T., Hinostroza, M. A., Wang, Z., Guedes Soares, C., 2020. Experimental investigation of shallow water
effect on vessel steering model using system identification method. Ocean Engineering, 199:106940.

Yao, J.X., 2015. On the Propeller Effect when Predicting Hydrodynamic Forces for Manoeuvring Using
RANS Simulations of Captive Model Tests. Doctoral Dissertation. Technical University of Berlin, Germany.

Yao, J., Jin, W., Song, Y., 2016. RANS simulation of the flow around a tanker in forced motion. Ocean.
Engineering 127: 236–245.

Yao J., Liu Z., Song X., Su Y., 2021. Ship manoeuvring prediction with hydrodynamic derivatives from RANS:
Development and application. Ocean Engineering, 231: 109036.

44
December 2022

View publication stats

You might also like