2011-PhD-Toxopeus Practical Application of Viscous-flow Calculations for the Simulation of Manoeuvering Ships
2011-PhD-Toxopeus Practical Application of Viscous-flow Calculations for the Simulation of Manoeuvering Ships
Proefschrift
door
Maritiem Ingenieur
geboren te Alkmaar
ii
Samenstelling promotiecommissie:
Rector Magnificus, voorzitter
Prof. dr. ir. R.H.M. Huijsmans, Technische Universiteit Delft, promotor
Prof. dr. ir. M. Vantorre, Universiteit Gent
Prof. Dr.-Ing. A. Cura Hochbaum, Technische Universität Berlin
Prof. dr. A.E.P. Veldman, Rijksuniversiteit Groningen
Prof. B. Pettersen, Norwegian University of Science and Technology
Prof. ir. J.J. Hopman, Technische Universiteit Delft
Dr. S.R. Turnock, University of Southampton
Prof. dr. ir. T.J.C. van Terwisga, Technische Universiteit Delft, reservelid
Copyright c 2011 by S.L. Toxopeus, Wageningen, The Netherlands. All rights reserved.
Published by the Maritime Research Institute Netherlands (MARIN)
ISBN 978-90-75757-05-7 (print)
iii
Summary
Practical application of viscous-flow calculations for the simulation of
manoeuvring ships
Figure 1: Impression of the flow field and hull surface pressures, KVLCC2, β = −10◦
The present work was initiated in order to improve traditional manoeuvring simula-
tions based on empirical equations to model the forces and moments on the ship. With
the evolution of the capability of viscous-flow solvers to predict forces and moments on
ships, it was decided to develop a practical method to simulate the manoeuvrability of
ships in which viscous-flow solvers are utilised and to investigate whether this improves
the accuracy of manoeuvring predictions.
To achieve this goal, the virtual captive test approach is adopted, because of the effi-
cient use of computational resources compared to other methods. This procedure mimics
the approach for manoeuvring simulations in which experimental PMM is used to obtain
the forces and moments on the ship. This study extends the work of other researchers
by providing extensive verification and validation of the predicted forces and moments on
the hull and a detailed study of the sensitivity of the manoeuvring characteristics of the
ship to changes in the hydrodynamic coefficients in the simulation model.
Changes in the flow solvers were required to be able to calculate the flow around
iv Summary
ships in rotational motion. These changes are discussed as well as the acceleration tech-
niques that were developed to reduce the effort spent on grid generation and during the
computations.
In this thesis, it is demonstrated that good predictions of the loads on the hull in
manoeuvring motion can be obtained for a wide range of ship types. The trends in the
forces and moments as a function of the drift angle or yaw rate are simulated well.
The verification studies provide useful insight into the influence of grid density on
the predicted forces and moments. In several cases, validation of the calculations failed,
indicating modelling errors in the numerical results. In these cases, it was generally seen
that the magnitude of the transverse force was under-predicted, while the magnitude of
the yaw moment was over-predicted. For manoeuvring studies in the early design, the
comparison errors are within acceptable levels. However, improvements remain desired
and may be obtained using finer grids, larger domain sizes, different grid topologies with
refinement in the wake of the ship, other turbulence models or incorporating free surface
deformation.
The manoeuvring prediction program SurSim has been used to simulate the manoeu-
vrability of the HTC. A procedure is proposed to derive the hydrodynamic coefficients
required to model the forces and moments on the bare hull. This procedure is chosen to
enable accurate modelling of the linearised behaviour for course-keeping as well as realis-
tic modelling of the harbour manoeuvring characteristics, and to enable the modelling of
non-linear manoeuvres accurately.
To generate validation data for the manoeuvring predictions presented in this thesis,
free sailing manoeuvring tests for the HTC were performed. This test campaign resulted
in a very valuable data set which can be used for public validation studies. Besides
obtaining general characteristics of the manoeuvrability of a single-screw container ship,
unique information has been obtained on the drift angles and rates of turn combined with
propeller and rudder forces. Furthermore, repeat tests have been conducted for selected
manoeuvres. Based on these tests, the uncertainty in the characteristic manoeuvring
properties has been estimated.
By using hydrodynamic manoeuvring coefficients derived from the CFD calculations,
it has been shown that it is possible to improve the prediction of ship manoeuvres com-
pared to predictions using coefficients based on empirical equations. A considerable im-
provement in the turning circle predictions was obtained. The prediction of the yaw
checking and course keeping and initial turning abilities based on zig-zag simulations im-
proved as well, but further improvements are required for more reliable assessment of the
manoeuvring performance.
The sensitivity of the manoeuvring predictions to changes in the hydrodynamic co-
efficients was studied. It was found that for accurate predictions of the manoeuvrability
using coefficients derived from CFD calculations, accurate predictions of especially the
yawing moment must be made.
v
Contents
Summary iii
Review of figures x
Acronyms xix
1 Introduction 1
1.1 Problem definition and objectives . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 Outline of the thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
3 Mathematical model 25
3.1 Coordinate system and nomenclature . . . . . . . . . . . . . . . . . . . . . 25
3.2 Manoeuvring simulation program SurSim . . . . . . . . . . . . . . . . . . 26
3.2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.2.2 Equations of motion . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.2.3 Hull forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.2.4 Propeller forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.2.5 Rudder forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.3 Deriving the hydrodynamic coefficients for the user-defined hull forces . . . 34
3.4 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
References 127
Samenvatting 143
Acknowledgements 145
Appendices 149
Table pages . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
Figure pages . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177
ix
Review of tables
6.1 Experimental uncertainty estimate, zig-zag, 95% confidence interval, all tests . . 100
6.2 Experimental uncertainty estimate, zig-zag, 95% confidence interval, set 1 . . . . 101
6.3 Experimental uncertainty estimate, turning circle, 95% confidence interval, all tests101
6.4 Experimental uncertainty estimate, turning circle, 95% confidence interval, set 1 102
Review of figures
1 Impression of the flow field and hull surface pressures, KVLCC2, β = −10◦ . . . iii
7.1 Comparison between the original simulations and the free sailing experiments, 18 kn103
7.2 Estimated resistance curve (model scale values) . . . . . . . . . . . . . . . . . . . 105
7.3 Estimated propeller open water curves . . . . . . . . . . . . . . . . . . . . . . . . 106
7.4 Forces on the ship as function of rudder angle, HTC without propeller . . . . . . 107
7.5 Forces on the ship as function of rudder angle and propeller revolutions . . . . . 108
7.6 Forces on the ship as function of rudder angle, 18 kn, β = −10◦ , n = ns . . . . . 109
7.7 Forces on the ship as function of rudder angle, 18 kn, γ = 0.4, n = ns . . . . . . . 110
7.8 Forces on the bare hull as function of yaw rate or drift angle . . . . . . . . . . . . 111
7.9 RPM-Speed relation for HTC, scale 1:30.02, model scale values . . . . . . . . . . 113
7.10 Sensitivity study, HTC, zig-zag, 10 kn . . . . . . . . . . . . . . . . . . . . . . . . 114
7.11 Sensitivity study, HTC, turning circle, 10 kn . . . . . . . . . . . . . . . . . . . . . 115
7.12 Comparison between the simulations and the free sailing zig-zag experiments, 10 kn118
7.13 Comparison between the simulations and the free sailing 20◦ /20◦ zig-zag experi-
ments, 18 kn . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
7.14 Comparison between the simulations and the free sailing turning circle experi-
ments, 18 kn . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
List of Symbols
p−p∞
Cp pressure coefficient = 1 2
ρV∞
[−]
2
m mass [kg]
mij added mass/inertia in direction i due to acceleration in direction j [kg, kgm, kgm2 ]
p pressure [N/m2 ]
Qp propeller torque [N m]
R radius [m]
xvi List of Symbols
R resistance [N ]
Tp propeller thrust [N ]
X longitudinal force [N ]
Y transverse force [N ]
y2+ non-dimensional wall distance of first cell away from the wall [−]
Z vertical force [N ]
Acronyms
Chapter 1
Introduction
An increase in ship sizes can be seen in the last decades. One reason for the increase of
cargo ships is the improved economy of transporting goods with larger ships. For cruise
ships, the increase in size is probably also driven by the need to provide more diverse
amusement for passengers and the competition of cruise operators to provide cruises on
the largest cruise ship in the world. Furthermore, more and more goods are distributed
using ships and therefore an increase of traffic density is observed.
The enlargement of ships and the increased traffic density lead to new challenges in
the design of the ship, one of which is the demand for better manoeuvrability. Since
large ships must operate in existing harbours, they will experience shallow water effects
more severely and the traffic density requires better controllability of the ships. In the
past, the interest in improved manoeuvrability has resulted in requirements posed by the
International Maritime Organization (IMO), while recently also the North Atlantic Treaty
Organization (NATO) started working on manoeuvring criteria for naval ships.
Several methods are available to assess whether a ship’s manoeuvrability complies
with the requirements. When conducted properly, full scale trials will provide information
about ship manoeuvres free of scaling effects or other assumptions. However, due to
weather conditions or current, it may be difficult to obtain accurate trial results and when
the manoeuvrability is deemed insufficient, modifications to the ship will be extremely
expensive. Therefore, assessment of the manoeuvrability is generally made in the design
stage, using model tests or simulations. This approach is much less expensive and provides
more flexibility in the selection of e.g. the steering arrangements, with the drawback that
scale effects will influence the results (model tests) and inaccuracies may be present in
the simulations due to improper selection of the mathematical model or hydrodynamic
coefficients used.
For training purposes of the crew of ships, or feasibility studies regarding entries of
large ships in existing harbours, an increasing demand for full-mission bridge simulations
is observed. Furthermore, the manoeuvrability with new propulsors (e.g. pods), or new
control strategies or operations (joystick control, dynamic tracking, side-by-side opera-
tions) are more and more tried out in simulators before the application in the real ship.
2 Chapter 1 – Introduction
To represent reality as much as possible, the mathematical model of the ship should mimic
the response of the ship to rudder or engine commands as well as possible. This also poses
new requirements on the accuracy of the mathematical models used in the simulation of
ship manoeuvres.
In the last decade, considerable developments have been made in simulation of the
viscous-flow around ships in order to predict the flow in manoeuvring conditions and to
determine the associated forces and moments. With these developments, the possibility
to improve manoeuvring simulations and to partly replace model tests with simulations
emerged.
The viscous-flow solvers used for the present study are briefly discussed in chapter 4,
together with the grid procedures and boundary conditions used for the current work.
Subsequently, verification and validation of the predicted forces and moments on sev-
eral different ship hulls in manoeuvring motions will be presented in chapter 5 using vali-
dation data available in literature. It is demonstrated that for a wide range of ship types,
accurate predictions of the loads on the hull in manoeuvring motion can be obtained.
Free sailing manoeuvring model tests were performed within the context of this thesis
to provide detailed validation data for the manoeuvring simulations. The tests and an
estimation of the uncertainty in the manoeuvring parameters are discussed in chapter 6.
Coefficients will be derived from these virtual captive tests and will be used in the
simulation program SurSim to model the forces on the hull and subsequently simulate
standard manoeuvres. The modifications of the mathematical model and the results of
the simulations will be presented and discussed in chapter 7.
Finally, conclusions are drawn and recommendations for further work are given in
chapter 8.
For descriptions of the symbols and abbreviations used in this thesis, the reader is
referred to page xiii and page xix, respectively.
4 Chapter 1 – Introduction
Chapter 2
Background of manoeuvring
simulation
2.1 Introduction
Traditionally, ship manoeuvring studies have focused on assessing compliance with the
manoeuvring standards set by the IMO [69]. However, due to emerging owner and op-
erational requirements, the need has arisen for assessment of manoeuvring capabilities in
operations other than the manoeuvring conditions prescribed by the IMO requirements,
see, for example, Quadvlieg and Van Coevorden [111] or Dand [32]. For naval ships, the
NATO Specialist Team in Naval Ship Manoeuvrability is developing a Standardization
Agreement (STANAG) regarding common manoeuvring capabilities for NATO warships
for specific missions. Örnfelt [103] provides an overview of the NATO efforts towards this
STANAG. Preliminary criteria have been published in NATO Allied Naval Engineering
Publication (ANEP) 70 [96]. In Armaoğlu et al. [8] and Quadvlieg et al. [110], demon-
strations are given of the use of prediction tools to verify compliance with the STANAG
criteria.
The assessment of the manoeuvrability of ships in the design stage can be done ex-
perimentally or numerically, or by combining both. For most engineers, free sailing model
tests are generally the preferred option, since they provide immediate insight into the
manoeuvring characteristics of the ship and no assumptions are made regarding the hy-
drodynamics of the model. However, due to scaling of the model, deviations between
model scale and full scale manoeuvres may occur. Additionally, free sailing model tests
do not give quantitative insight into the forces and moments acting on the hull, which is
required when full mission bridge simulator studies, e.g. for training or feasibility studies,
6 Chapter 2 – Background of manoeuvring simulation
are to be conducted.
Furthermore, sometimes different design variants need to be compared before con-
structing and testing physical models of the ship. In those cases, the use of manoeuvring
simulation programs is preferred. However, to obtain reliable simulations, reliable mod-
els of the forces and moments acting on the ship are required, since these are needed to
calculate the accelerations, velocities and trajectories of the ship during the manoeuvre.
A flow chart of manoeuvring simulations is given in Figure 2.1. It is seen that to obtain
the forces on the ship, either experimental (mostly Planar Motion Mechanism (PMM))
or numerical (mostly empirical) techniques can be used. These forces are fed into a ma-
noeuvring simulation program in the form of coefficients or tables and with the program
the manoeuvring characteristics of the ship can be determined. If needed, modifications
to the design can be made in order to ensure compliance with manoeuvring requirements.
When the results are found to be satisfactory, the mathematical model can be used in a
simulator for e.g. training.
Hull Design
Hull forces
ok?
Manoeuvring- Simulator
no
characteristics training
yes
The traditional simulation tools use empirical descriptions of the forces and moments
on the ship’s hull and are generally based on regression analysis of captive manoeuvring
test data for a (preferably wide) range of ships. Due to the lack of resolution of hull
details or application outside the range of the regression database, the prediction of the
manoeuvrability may be unreliable. On the other hand, empirical simulation tools can
provide valuable information regarding the manoeuvring characteristics in a cost-effective
way during the early stages of the design.
Other methods to obtain mathematical models of the forces and moments on the ship
comprise conducting captive model tests for the ship under consideration, or by conducting
a series of free sailing model tests and subsequent system identification. Although these
2.1 – Introduction 7
techniques may yield accurate predictions of the manoeuvring characteristics of the ship,
the construction of the physical model and the use of the experimental facilities can be
costly. Furthermore, the data needs to be analysed and fed into a simulation program,
before the actual manoeuvring characteristics of the design can be assessed.
Therefore, new methods are required to obtain reliable and accurate manoeuvring
simulation models in a cost-effective manner. Such methods should not only be suitable
to predict the yaw checking and turning ability of the ship according to the IMO require-
ments, but also be applicable to operation in confined waterways or harbour manoeuvring
assessment studies, for example.
Recently, viscous-flow calculations provide an attractive means to improve manoeu-
vring simulations. Two approaches are available to use such calculations in the prediction
of the manoeuvring of ships:
virtual captive tests: the forces and moments for a range of forced motions (steady
drift, steady rotation, oscillatory sway or yaw, or combinations thereof) are cal-
culated. From the calculations, hydrodynamic coefficients can be derived which
are subsequently used in the simulation model to predict the forces and moments.
Sometimes, the forces and moments are obtained by interpolation between the data
points. This approach resembles the approach taken when using PMM tests and
the calculations are therefore referred to as virtual captive tests. This procedure is
further discussed in section 2.3.
coupling with body motions: the calculated forces and moments are directly used in
the equations of motions to obtain the accelerations, velocities and position of the
ship. This procedure is further discussed in section 2.4.
the design have to be considered within this time frame. Furthermore, incorporation of
this method in real-time simulators is not yet feasible1 , which means that for simulator
studies the method of using virtual captive tests is the only viable solution.
When using existing modular mathematical manoeuvring models, see e.g. section 2.2,
it is possible to use the virtual captive tests to improve low-fidelity sub-models and provide
manoeuvring advice within a reasonable time frame. For example, when it is expected
that the modelling of the propeller or rudder performance in the mathematical model is
sufficiently accurate, virtual captive tests need only be conducted for the bare hull. This
greatly simplifies the grid generation process and reduces the number of grid points (and
thereby computation time) required to arrive at an accurate estimation of the loads on
the ship.
Kang and Hasegawa [76] to extend the model to low-speed manoeuvring. For the SIM-
MAN 2008 workshop [130], several contributions were made in which the MMG model
was used to model the manoeuvring of the test case ships.
Other models differing from the MMG model are proposed by e.g. Oltmann [101] and
Hooft and Quadvlieg [66]. Both models utilise cross flow drag coefficients (see e.g. Hooft
[63]) to model non-linear effects in the forces and moments on the ship. Variations of
the model of Hooft and Quadvlieg form the basis of the MARIN in-house manoeuvring
simulation programs SurSim, FreSim (see Hooft and Pieffers [65]) and MPP which have
been used in submissions for the SIMMAN 2008 workshop [149]. SurSim has been used
as a basis for the simulations in this thesis, and is described in detail in section 3.2.
Another attempt to provide a practical but sufficiently accurate, general and physi-
cally sound mathematical model was made by Ankudinov and Jakobsen [6]. Guidelines
for the development of a standard simulation model are given. More information about
manoeuvring models and their applications can be found in Fossen [52] and Eloot [47].
ficients and use these to simulate the manoeuvring behaviour of a ship was presented by
Racine and Patterson [112]. For a novel hull form, for which accurate empirical formulae
to describe the forces and moments due to manoeuvres were not available, coefficients
were derived and the stability and trajectory of the vessel were assessed. The hull form
under consideration was the Newport News Experimental Model (NNemo). A sensitivity
study was performed to determine the scope of the required calculations and to reduce
the size of the CFD matrix. Unfortunately, validation of the simulated behaviour could
not be conducted at the time of the study although some of the phenomena found in the
simulations were apparently also found during free running model tests.
A detailed study using virtual PMM simulations is presented by Cura Hochbaum
[28]. Here, simulations of zig-zag and turning circle manoeuvres are conducted for a twin-
screw ferry. The time traces of the PMM simulations are compared to experimental PMM
results. Additionally, the coefficients derived from the numerical study are compared to
the coefficients derived from the experiments. Finally, the manoeuvre results are compared
to the results based on simulations using the experimental hydrodynamic coefficients and
to results obtained using free sailing experiments. The validation is encouraging and
it is demonstrated that the procedure works well, although some improvements in grid
resolution and modelling are proposed.
For the SIMMAN 2008 workshop [130], two participants provided manoeuvring simu-
lations for the KVLCC2 (see section 2.5.3) using coefficients derived from CFD results, i.e.
Cura Hochbaum et al. [30] and Toxopeus and Lee [149]. Cura Hochbaum et al. conducted
CFD calculations for static drift, oscillatory motion and for rudder deflections to arrive
at complete mathematical models for the KVLCC1 and KVLCC2. The agreement of
manoeuvring simulations using the mathematical models with the experiments was very
promising, especially for the KVLCC2. Toxopeus and Lee calculated the hydrodynamic
coefficients using CFD for the bare hull, and used empirical formulae to calculate the
forces due to the propeller and rudder. Comparison of the simulated manoeuvres with
the free sailing experiments showed that the mathematical model needed to be improved,
mainly by extending the range of drift angles and yaw rates used to derive the coefficients.
The present thesis demonstrates a procedure similar to the one used in Toxopeus and
Lee [149], but for a different test case and more attention is paid to the correct modelling
of the forces and moments on the ship.
the equations motions of the ship. The instantaneous forces on the hull are calculated
using CFD, while the forces due to the propeller and rudder are calculated using empirical
formulae based on the MMG model. With their model, they perform zig-zag manoeuvres
for two Very Large Crude Carrier (VLCC) variants. Comparison with the experiments
shows reasonable quantitative agreement.
Pankajakshan et al. [106] apply a coupled procedure to simulate overshoot manoeuvres
for the ONR Body-1 submarine model. The control surfaces and rotating propeller are
included in the viscous-flow calculations. The propeller is incorporated in the simulation
with sliding interfaces. The deflection of the planes is modelled using re-generation of the
grid for each deflection angle, based on interpolation between several grids spanning the
range of deflection of the control surface. The agreement between the simulated results
and the experiments is good.
Jensen et al. [74] show a turning circle simulation for a container ship. The rudder
is modelled with sliding interfaces, while the propeller is modelled using body forces.
Unfortunately, validation was not performed.
Venkatesan and Clark [160] present simulations for an overshoot manoeuvre for ONR
Body-1 (similar to the work by Pankajakshan et al. [106]) and compare the results to
experiments. The propeller is modelled using sliding interfaces, while the control surface
deflection is modelled using mesh deformation. The agreement is promising, but appears
to be slightly less than for Pankajakshan et al.
For the SIMMAN 2008 workshop [130] Carrica and Stern [23] performed Detached
Eddy Simulation (DES) of the KVLCC1 performing zig-zag and turning circles with
moving propeller and rudder. Overlapping grid techniques are used to model the moving
appendages and a level set approach is used to capture the free surface. The simulations
were found to be computationally very intensive and could not be finished before the
workshop. The results are promising and a good demonstration of the current capabilites
of CFD, but some issues remain to be solved. Carrica et al. [22] also conducted free
sailing manoeuvring RANS simulations for the 5415M (see section 2.6.4). Also in this
case, overlapping grids are used to model the hull, bilge keels, stabiliser fins, shafts,
struts and moving rudder and a level set method is used for free surface capturing. The
propeller is modelled through the body-force approach, neglecting local velocity effects.
The agreement between the simulations and free sailing experiments performed at MARIN
[150] was very good, leaving rather limited suggestions for improvements.
Another interesting example of coupling the RANS solution to rigid body motions
is given by Bettle et al. [14]. They study the rising stability of a submarine, i.e. the
development of the roll angle when a submarine needs to surface quickly. Calculations
for full scale Reynolds numbers have been performed. The forces due to the propeller,
ballast system and appendages have been incorporated using coefficient based models.
The simulations detected the underwater roll instabilities and were consistent with results
obtained using fully coefficient-based simulations and with observations during full scale
trials. With the CFD results, the main source of the instabilities, i.e. the rolling moment
generated by the sail, was identified.
12 Chapter 2 – Background of manoeuvring simulation
2.5.1 Series 60
Extensive flow field and force measurements on the well-known Series 60 hull form
were conducted at Iowa Institute of Hydraulic Research (IIHR) and the results were made
available to the public by Longo [89]. For a range of drift angles and Froude numbers,
the forces on the model were measured. Furthermore, wave patterns were measured for
a selected set of drift angles and speeds. For a drift angle of 10◦ , the mean flow was
obtained at several longitudinal stations. During the force measurements, the model was
free to sink, trim and heel and the displacements were recorded. During all other tests,
the model attitude was fixed. The water depth to ship’s draught ratio of h/Tm = 18.7
represented deep water conditions. The main particulars and body plan of the Series 60
are presented in Table 2.1 and Figure 2.2.
Because of the large amount of data obtained during the measurement campaign, this
case is very suitable for validation studies of viscous-flow calculations. Various researchers
have already reported such validation studies, such as Alessandrini and Delhommeau [3],
Cura Hochbaum [27], Campana et al. [21], Di Mascio and Campana [36], Tahara et al.
[140], Toxopeus [141, 142] and Di Mascio et al. [35].
The Series 60 test case is not further considered in this thesis.
2.5 – Validation cases for captive manoeuvring 13
Magnitude
Description Symbol proto model Unit
Scale λ 1:1 1:40 -
Length between perpendiculars Lpp 121.920 3.048 m
Breadth max. moulded B 16.250 0.406 m
Draught moulded fore Tf 6.500 0.163 m
Draught moulded aft Ta 6.500 0.163 m
Displacement volume moulded ∇ 7715 0.121 m3
Wetted surface area bare hull Swa 2528 1.580 m2
Position centre of buoyancy forward of midship xB -1.523 %Lpp
Block coefficient Cb 0.600 -
Length-Breadth ratio L/B 7.503 -
Breadth-Draught ratio B/T 2.500 -
Length-Draught ratio L/T 18.757 -
Wu et al. [165] conducted studies on the SUBOFF moving close to the sea floor. They
found that the bottom effects are proportional to Bc , with B the submarine beam and c
the clearance between the bottom and the submarine hull. The paper shows that RANS
solvers can be used as a practical tool to predict hydrodynamic aspects of submarines.
Etebari et al. [49] present results of a test campaign for the SUBOFF model undergoing
a steady turn. Stereo-Particle Image Velocimetry (PIV) was used to obtain the flow field
around the model. Pressure measurements at two axial cross sections on the model were
conducted and the forces and moments on the model were measured. Three configurations
of the SUBOFF were used: the bare hull, the fully appended model and the bare hull
with towed-array fairing.
Several series of unsteady measurements using the SUBOFF have been conducted, see
e.g. Whitfield [164] or Hosder [67]. Further experiments have been conducted by Granlund
and Simpson [56] on the added mass of the SUBOFF (AFF-2). This work has resulted in
the thesis by Granlund [55].
Magnitude
Description Symbol KVLCC1 KVLCC2 KVLCC2M Unit
Length between perpendiculars Lpp 320.000 320.000 320.000 m
Breadth max. moulded B 58.000 58.000 58.000 m
Draught moulded fore Tf 20.800 20.800 20.800 m
Draught moulded aft Ta 20.800 20.800 20.800 m
Displacement volume moulded ∇ 312631 312635 312650 m3
Wetted surface area bare hull Swa 27370 27257 27279 m2
Position centre of buoyancy forward of midship xB 3.494 3.497 3.527 %Lpp
Block coefficient Cb 0.810 0.810 0.810 -
Length-Breadth ratio L/B 5.517 5.517 5.517 -
Breadth-Draught ratio B/T 2.788 2.788 2.788 -
Length-Draught ratio L/T 15.385 15.385 15.385 -
Figure 2.4: Body plans of KVLCC1 and KVLCC2 (dotted: KVLCC1, continuous: KVLCC2)
forms, it was anticipated that the manoeuvrability of these ships would be different. The
KVLCC1 and KVLCC2 hull forms were two of the subjects of study during the CFD
Workshop Gothenburg 2000 [83] and the SIMMAN 2008 workshop [130]. The KVLCC2
was studied during the CFD Workshop Gothenburg 2010 as well.
Experiments have been conducted by Kim et al. [77]. For the KVLCC1 and KVLCC2
hull forms, they performed wave elevation and flow field measurements around the 1 : 58
scaled ship models. Furthermore, resistance and propulsion tests were conducted. The
attitude of the model was fixed, except during the resistance and propulsion tests. During
these tests, the trim of the model was modified to arrive at a level running trim during
the actual tests. The water depth to ship’s draught ratio of h/Tm = 13.5 represented
deep water conditions. The publication also contains results for the KRISO Container
Ship (KCS) (see 2.6.3). Similar flow field measurements have been conducted in a wind
16 Chapter 2 – Background of manoeuvring simulation
tunnel, using double-body models. Tests were only conducted for straight-ahead sailing.
The results were published by Lee et al. [85].
Captive model tests for the bare hull KVLCC2 were conducted by Istituto Nazionale
Per Studi Ed Esperienze Di Architettura Navale (INSEAN) in preparation for the SIM-
MAN 2008 Workshop [130], see also Fabbri et al. [50]. A set of PMM tests was performed,
comprising amongst others the measurement of the forces and moments for steady drift
motion and oscillatory yaw motion. During the tests, the model was free to heave and
pitch. The model tests were conducted for deep water (water depth to ship’s draught
ratio h/Tm = 8.3) as well as for restricted water depths using a false bottom: h/Tm = 1.2,
1.5 and 3.0. The scale factor was 1 : 45.71.
Additional measurements for the KCS, KVLCC1 and KVLCC2 have been done by
Ueno et al. [153]. Using CMT, they obtained the forces and moments on the hull with
rudder and propeller as a function of the drift angle and rotation rate. The models were
free to trim, sink and heel. For the KVLCCs, the water depth to draught ratio was
h/Tm = 10.6, representing a deep water condition. The scale factor was 1 : 110.
The KVLCC2M hull form was conceived as a variant of the KVLCC2 but with a
slightly modified aft ship to reduce the complexity of the flow and therefore simplify
viscous-flow computations. The most visible difference is the fairing of the hull around
the propeller shaft position. The KVLCC2M was the main subject of the manoeuvring
studies in the CFD Workshop Tokyo 2005 [58]. Model tests for the KVLCC2M have
been extensively described by Kume et al. [82]. These tests comprised measurements of
forces, surface pressures and stern flow fields for the 1 : 64.4 scaled model at several drift
angles. During the tests, the movement of the model with respect to the carriage was
fully constrained. The water depth to ship’s draught ratio of h/Tm = 24.8 represented
deep water conditions.
Full scale ships of the KVLCCs do not exist. The main particulars of these ships and
2.5 – Validation cases for captive manoeuvring 17
body plans are given in Table 2.3, Figure 2.4 and Figure 2.5 respectively.
Several authors have studied the manoeuvrability of these VLCCs, such as the partic-
ipants of the Tokyo CFD Workshop [58], Toxopeus [143] (KVLCC2M, V&V), Simonsen
and Stern [126, 127] (KVLCC2, deep/shallow water, V&V), Broglia et al. [17] (KVLCC2,
blockage during PMM), Carrica and Stern [23] (KVLCC1, DES coupled with body mo-
tions), Cura Hochbaum et al. [30] (KVLCC1 / KVLCC2, virtual PMM), Toxopeus and
Lee [149] (KVLCC2, virtual PMM), Muscari et al. [95] (KVLCC2, RANS coupled with
body motions), Stern et al. [132] (KVLCC2, DES at large drift angles) and Phillips et al.
[107] (KVLCC2, virtual PMM). More information about the KVLCC results obtained at
SIMMAN 2008 can be found in Stern et al. [131].
2.5.4 HTC
The Hamburg Test Case (HTC) is a model of the container ship built by Bremer
Vulkan in 1986 as Ville de Mercure, and subsequently named Teresa del Mar, see Fig-
ure 2.6. Teresa del Mar was sold in 2010 and renamed Maria. After the Ville de Mercure
a number of other container ships with the same hull form were built. One of the sister
vessels, Catalina del Mar, is still sailing.
Captive model experiments were conducted on the HTC within the VIRtual Tank
Utility in Europe (VIRTUE) project by Hamburgische Schiffbau-Versuchanstalt (HSVA)
in order to provide additional material for CFD validation. These tests comprised force
measurements for the bare hull, the hull with rudder and the hull with propeller and
rudder. Furthermore, PIV measurements were conducted for the model equipped without
18 Chapter 2 – Background of manoeuvring simulation
rudder, with the model sailing at steady rotational motion. These experiments were
reported in VIRTUE deliverable D3.1.3, see Vogt et al. [161]. The scale of the model λ
was 1:24 during the HSVA tests. The water depth to ship’s draught ratio of h/Tm = 14
represented deep water conditions.
The HTC has been studied numerically by several authors, such as Drouet et al. [37],
Gao and Vassalos [54] and Toxopeus [144, 145, 148].
Hull form
In Table 2.4, the main particulars of the HTC are presented. Both model scale and full
scale (prototype) values are given.
HTC
Description Symbol proto model Unit
Scale λ 1:1 1:24 1:30.02 -
Length between perpendiculars Lpp 153.700 6.404 5.120 m
Breadth max. moulded B 27.500 1.1458 0.916 m
Draught moulded fore Tf 10.300 0.4292 0.343 m
Draught moulded aft Ta 10.300 0.4292 0.343 m
Displacement volume moulded ∇ 28342 2.0500 1.048 m3
Wetted surface area bare hull Swa 5567 9.6640 6.177 m2
Position centre of buoyancy forward of midship xB -0.571 %Lpp
Block coefficient Cb 0.650 -
Length-Breadth ratio L/B 5.582 -
Breadth-Draught ratio B/T 2.673 -
Length-Draught ratio L/T 14.922 -
The body plan of the HTC is presented in Figure 2.7. A photograph of the HTC
during the model tests at HSVA is presented in Figure 2.8.
The model was tested fixed in all degrees of motion. It was not equipped with bilge
keels. For the measurements of the model with rudder, the rudder forces were measured
separately. Turbulence was stimulated in all tests. Therefore, in the viscous-flow calcula-
tions it was assumed that the flow was fully turbulent.
2.5 – Validation cases for captive manoeuvring 19
Figure 2.8: HTC model during oblique motion test using CPMC
(photograph by HSVA)
20 Chapter 2 – Background of manoeuvring simulation
Propeller
The particulars of the propellers of the prototype HTC and of the model test propellers
(scaled to prototype values) are specified in Table 2.5.
Rudder
For the model tests, the rudder was divided into a movable and a fixed (headbox) part in
order to allow turning of the rudder without touching the hull surface. The particulars of
the rudder and a drawing are presented in Table 2.6 and Figure 2.9.
For validation of the predicted manoeuvres, results from free sailing manoeuvring tests or
full scale trials are required. This section presents some of the test cases that are available
in literature for which free sailing manoeuvring data is present.
The Esso Osaka received ample attention due to the existence of well-documented trials
in deep and shallow water published by Crane [26], and it was recommended by the 22nd
International Towing Tank Conference (ITTC) that this ship was used for validation of
force predictions and manoeuvring simulations. However, due to the fact that the hull
form is rather outdated and some doubts arose regarding the scatter in the results during
analysis of different model test campaigns [25], the interest in this validation case has
diminished in the last few years.
Recent studies in which the viscous-flow around the Esso Osaka for captive conditions
was simulated were published by e.g. El Moctar [46], Simonsen [122], Simonsen and Stern
[123, 124, 125] and Van Oers and Toxopeus [155].
22 Chapter 2 – Background of manoeuvring simulation
2.6.3 KCS
During the SIMMAN 2008 workshop, it was concluded that the existing free sailing ma-
noeuvres with the KCS were not sufficiently accurate to be used as reliable validation
data. Therefore, MARIN decided to conduct a new set of free sailing manoeuvres with
the KCS. The tests comprised zig-zag and combined turning-circle/pull-out manoeuvres,
with variations in GM values and for fully loaded and ballast condition, see Overpelt [104].
With a model scale of 1 : 39.89, the water depth to ship’s draught ratio of h/Tm = 17.5
represented deep water conditions.
2.6.4 5415M
In order to develop new hull form concepts for advanced naval mono-hull ships, a co-
operative research programme, called ”Thales”, was initiated between the Royal Nether-
lands Navy, the Italian Navy and the Danish Navy. The programme comprised develop-
ment of design requirements and procedures, determination of assessment procedures and
selection of design tools. To be able to select the best design tools, benchmark data were
necessary with which design tool data could be compared. Therefore, a model test pro-
gramme was conducted to generate this data. Model tests were conducted by INSEAN
(Italy), DMI (Denmark) and MARIN (The Netherlands). Additional information was
obtained from DTMB (USA).
For the model tests, the representative high speed displacement hull form of the US
Navy DDG 51 destroyer (in literature often designated DTMB 5415) was selected. This
hull form was fitted with a representative twin-propeller / twin-rudder arrangement and
centre-line skeg design comparable to European design practice. Furthermore, stabiliser
fins were fitted to the model. To distinguish this modified geometry and appendage
arrangement to the DTMB 5415, it is designated 5415M. Standard zig-zag experiments,
spiral tests and combined turning circle / pull-out tests were conducted. During these
tests, forces on the appendages were measured in order to validate manoeuvring simulation
programs in more detail. Several repeat runs for some of the manoeuvres were conducted
to assess the uncertainty in the manoeuvring parameters. The tests have been reported
2.7 – Conclusion 23
by Toxopeus and Lee [150]. With a model scale of 1 : 35.48, the water depth to ship’s
draught ratio of h/Tm = 29 represented deep water conditions.
The 5415M was selected as one of the test cases of the SIMMAN 2008 workshop. The
free sailing manoeuvre data are used to validate manoeuvring predictions in Carrica et
al. [22], Bhushan et al. [16], and in the SIMMAN 2008 proceedings [130]. Furthermore,
the roll decay and seakeeping tests performed with this free sailing model are used as
validation data in the NATO Research and Technology Organisation (RTO) Applied
Vehicle Technology (AVT)-161 working group.
2.7 Conclusion
Several different methods are given in literature to simulate ship manoeuvres using viscous
flow calculations. In this thesis, the virtual captive test approach is adopted, since this
is at present the most attractive approach when considering the computational costs.
Furthermore, this approach can directly be used to improve mathematical models for
manoeuvring simulators.
A survey of validation data existing in literature was made. For free sailing manoeu-
vres as well as for captive tests, data can be found that can be used to validate numerical
predictions. Noteworthy is the fact that validation data for captive steady drift motions
are much more abundant than validation data for steady rotation or combined motion.
Especially for more extreme conditions with large turning rates and drift angles (e.g. non-
dimensional turning rates γ of 0.8 - 1.0 and drift angles between 20◦ and 30◦ ), such as
occur during tight turning circles, not much validation data can be found in literature.
Furthermore, to the knowledge of the author, cases in which extensive captive test data
and free sailing manoeuvring test data are available have only been published for the Esso
Osaka and the KVLCCs (although the captive tests do not cover the complete range of
rotation rates experienced during turning circle manoeuvres). In the work leading up
to this thesis, much work was done on simulating the flow around the HTC for captive
conditions and simulations of standard free sailing manoeuvres were conducted within
the VIRTUE project. To generate validation data for these manoeuvres, MARIN decided
to perform free sailing manoeuvring tests for the HTC. Details about these tests will be
presented in chapter 6.
The tools used during this study and the procedure to predict the manoeuvrability of
the ship are discussed in the following chapters.
24 Chapter 2 – Background of manoeuvring simulation
Chapter 3
Mathematical model
The in-house manoeuvring simulation program SurSim will be used to simulate ship
manoeuvres within the present study. This chapter presents the system of coordinates
adopted in this thesis, details of SurSim and proposes the required steps to derive the
hydrodynamic coefficients for the bare hull forces and moments.
PS
y,r
x,u
d
b
SB
y,v V
SB PS
x,u
y,v
q,q f,p
z,w z,w
In ship manoeuvring, two reference frames are used. One frame is attached to the
ship and is called the ship-fixed coordinate system. The motion of the ship-fixed refer-
ence frame is described relative to an earth-fixed inertial reference frame. In the ship-fixed
coordinate system, x is directed forward, y to starboard and z vertically down, see Fig-
ure 3.1. For captive tests or simulation of captive motions, it is customary to use the
intersection of the waterplane, midship and centre-plane as origin O. Therefore, all forces
and moments are given with respect to O, with the longitudinal force X directed forward
positive and the transverse force Y positive when directed to starboard. A positive drift
26 Chapter 3 – Mathematical model
angle β corresponds to the flow coming from starboard. A positive non-dimensional yaw
rate γ corresponds to a turn to starboard when sailing at positive forward speed.
For free sailing manoeuvring tests or simulations, it is however customary to use the
centre of gravity G as the origin of the coordinate system. Therefore, the results of all
manoeuvres are presented with respect to G. In the results, the drift angle βG is given
relative to the centre of gravity is well.
In the earth-fixed coordinate system, xE is directed North, while yE is directed to the
East.
All forces and moments are presented non-dimensionally. The longitudinal force X
and transverse force Y are made non-dimensional1 using 21 ρV 2 Lpp T , the vertical force
Z using 21 ρV 2 Lpp B, the heeling moment K by 12 ρV 2 Lpp T 2 , the pitch moment M by
1
2
ρV 2 L2pp B and the yaw moment N by 21 ρV 2 L2pp T . This method of non-dimensionalisation
has been applied to all force components presented in this thesis. V is the speed of the
ship through the water.
Several subscripts are used to identify separate force components. In this thesis, a
subscript H indicates forces on the hull, R forces on the rudder, P forces on the propeller
and T total forces. Furthermore, f indicates a force contribution due to friction and p a
force contribution due to dynamic pressure (the hydrostatic force component is neglected,
since this force is cancelled by the displacement mass when free surface deformation is
neglected).
For a complete list of symbols used in this thesis, please refer to page xiii.
YT = YH + YP + YR (3.1)
1
Unless otherwise specified: for submarines forces and moments are generally made non-dimensional
using L2pp instead of Lpp T as reference area.
3.2 – Manoeuvring simulation program SurSim 27
in which the subscript T denotes the total forces, H the bare-hull forces, P the propeller
forces including hull-propeller interaction and R the rudder forces including hull-propeller-
rudder interaction. In this thesis, the results obtained with the bare-hull forces estimated
using the original empirical formulae are designated ”SurSim”. The results with bare-hull
coefficients obtained from the viscous-flow calculations will be designated ”CFD”.
The following sections describe the equations of motions and the different force com-
ponents. It must be noted that for each force component, user-defined models can also
be used.
NG = N − xG · Y (3.3)
Ship resistance
The straight ahead sailing resistance curve of the hull must be given as input to the
program. For this, the MARIN in-house program DESP can be used, which is based on
an improved version of the method of Holtrop and Mennen [62]. The resistance should
be predicted for the appropriate Reynolds numbers, in order to be able to compare the
simulation results directly with full scale trials or with model experiments.
28 Chapter 3 – Mathematical model
The slender-body method is a semi-empirical method to determine the linear force com-
ponents on the hull. According to this method, the linear manoeuvring coefficients are
determined by the rate of change of fluid momentum along the length of the ship. Only
a few empirical parameters based on careful validation with experiments are used. Since
this method utilises the full description of the hull form, the influence of changes in lo-
cal details of the hull can be investigated in the early design stage within a short time
frame (i.e. a few minutes when the hull form input has been prepared). A more detailed
description of the slender-body method can be found in e.g. Toxopeus [145].
The non-linear contributions to the forces and moments are calculated using e.g. the
so-called cross-flow drag theory, presented by e.g. Hooft and Quadvlieg [66] or Hooft [63].
0 0
+ Yβ|γ| · sin β |γ| + Y|β|γ · |sin β| γ + Yab0 · cosay β · sinby β · sign sin β (3.5)
up = u · (1 − w) (3.7)
up
J= (3.8)
n · Dp
KT = KT 0 + KT 1 · J + KT 2 · J 2 + KT 3 · J 3 + K T 4 · J 4 + KT 5 · J 5 (3.9)
KQ = KQ0 + KQ1 · J + KQ2 · J 2 + KQ3 · J 3 + KQ4 · J 4 + KQ5 · J 5 (3.10)
Tp = KT · ρn2 Dp4 (3.11)
Qp = KQ · ρn2 Dp5 (3.12)
The coefficients KT i and KQi in equations (3.9) and (3.10) are to be specified by the
user and can be obtained from propeller open-water tests. It should be noted that this
user-defined model is not valid for four-quadrant manoeuvre simulations. Based on the
formulae above, the longitudinal force on the ship due to the propeller is given by:
XP = (1 − t) · Tp (3.13)
30 Chapter 3 – Mathematical model
dh
de=d-dh
Hx
vr Hy
Lru
ur
Vrr
Dru
One of the most complicated aspects in determining the rudder forces is the determi-
nation of the flow velocity and direction at the rudder location as a consequence of the
drift angle, yaw rate and propeller action. Therefore, the formulae describing the flow
velocity are discussed first. To calculate the longitudinal velocity as a function of the
propeller loading, some basic formulae are required. The full analysis can be found in
e.g. Kuiper [81], but the basic principles are stated here to demonstrate the physics and
to modify the formulae for four-quadrant manoeuvres. A cross section of the flow to be
described is given in Figure 3.3.
3.2 – Manoeuvring simulation program SurSim 31
propeller
plane
up+uar up+ua up
p¥ p+Dp p p¥
p 2 p 2 p 2
S1= /4 D1 S= /4 Dp S0= /4 D0
strea
mline
Ahead speed with positive thrust Using axial momentum theory while assuming
inviscid and incompressible flow, the first relation is derived using the conservation of
mass:
π π π
up · D02 = (up + ua ) · Dp2 = (up + uar ) · D12 (3.14)
4 4 4
Conservation of momentum leads to:
π π
ρu2p · D02 − ρ (up + uar )2 · D12 + Tp = 0 (3.15)
4 4
Combining these two relations leads to the following equation for the propeller thrust:
π
Tp = Dp2 · ρ (up + ua ) · uar (3.16)
4
Applying Bernoulli’s law respectively in front and aft of the propeller disc, two addi-
tional equations are found:
1 1
p∞ + ρu2p = p + ρ (up + ua )2 (3.17)
2 2
32 Chapter 3 – Mathematical model
1 1
p∞ + ρ (up + uar )2 = p + ∆p + ρ (up + ua )2 (3.18)
2 2
Subtracting equations (3.18) from (3.17) and using Tp = ∆p· π4 Dp2 , the propeller thrust
is found to be:
π 2 1
Tp = Dp · ρ up + uar · uar (3.19)
4 2
such that uar = 2 · ua , following from equation (3.16).
Solving equation (3.19) for uar results in:
v
π π 8Tp
u
Tp − Dp2 · ρup · uar − Dp2 · ρu2ar = 0 ⇒ uar = −up ± tu2p +
u
(3.20)
4 8 ρπDp2
For positive inflow velocity up at the propeller and positive thrust of the propeller,
the axial induced velocity uar is larger than zero. Therefore, equation (3.20) reduces to:
v
8Tp
u
u
uar = tu2 + − up (3.21)
p
ρπDp2
Ahead speed with negative thrust When the forward speed up is positive but Tp
is negative and assuming that ur = up + uar > 0 (we still do not have flow reversal),
equation (3.21) is valid with the additional requirement that uar > −up , resulting in:
8Tp
> −u2p (3.22)
ρπDp2
If the thrust becomes too negative, we get flow reversal due to the propeller action.
In this case, it is assumed that the flow velocity ur at the rudder becomes zero. This
can be obtained by modifying equation (3.21), such that it reads for all conditions with
positive ahead speed up ≥ 0:
v !
8Tp
u
u
uar = tmax u2p + , 0 − up (3.23)
ρπDp2
Astern speed with negative thrust For astern speed and negative thrust, it is as-
sumed that there is no induced velocity from the propeller at the rudder location. In that
case, the change in axial velocity at the rudder becomes:
uar = 0 (3.24)
Astern speed with positive thrust For astern speed but with positive thrust, it is
assumed that the inflow velocity at the propeller position up equals zero. This results in
an axial induced velocity behind the propeller which depends on the propeller thrust and
the local undisturbed velocity u0 . The induced velocity is approximated using:
u0 = u · (1 − wp ) (3.25)
3.2 – Manoeuvring simulation program SurSim 33
v
8Tp
u
u
uar = max t
+ u0 , 0 (3.26)
ρπDp2
and the velocity at the rudder position is calculated with:
ur = u0 + uar (3.27)
Flow straightening Due to the drift angle of the ship and the rotation rate during
manoeuvres, the rudder will also experience a transverse velocity component vr . However,
due to the presence of the ship, the undisturbed flow will be rectified or straightened.
Therefore, the flow-straightening coefficients Cdb for the drift angle and Cdr for the yaw
rate are introduced:
vr = Cdb · v + Cdr · xr · r (3.29)
It must be noted that with this linear equation the effect of the hull on the flow is
strongly simplified (see e.g. Ogawa and Kasai [98]). It can be argued that, for a ship
in manoeuvring motion, vortices that are generated upstream may travel to the rudder
position, resulting in irregular inflow velocities and flow directions. This effect can be
studied using dedicated model tests or with CFD calculations, but is considered to be
outside the scope of the present study.
Rudder forces First the lift and drag coefficients are determined, see eq. (10) in [98]:
6.13 · Λ
CL = (3.30)
2.25 + Λ
CL2
CD = (3.31)
π·Λ
The hydrodynamic rudder angle describes the angle between the flow velocity at the
rudder position and the ship’s longitudinal axis and follows from:
vr
δh = arctan (3.32)
ur
34 Chapter 3 – Mathematical model
Rudder angle The rudder angle is determined by the required rudder angle δreq and
the actual rudder angle δ, mimicking a simplified steering machine. The required rudder
angle is set depending on the type of manoeuvre (e.g. auto-pilot, zig-zag, turning-circle).
To obtain the actual rudder angle for a new time step, the difference between the required
rudder angle and the actual rudder angle at the current time step is determined first:
∆δ = δreq − δ (3.39)
When |∆δ| is less than a certain threshold (e.g. 3◦ ), then the rudder rate is calculated by:
δ̇ = Cr · ∆δ (3.40)
with Cr a rudder rate constant obtained from Cr = δ̇max /3. For |∆δ| above the threshold,
the rudder rate is given by:
δ̇ = δ̇max · sign∆δ (3.41)
The actual rudder angle at the new time step is obtained by integrating the rudder
rate in time.
1. The linear coefficients for simple motions (slope of force or moment curves at β = 0◦
resp. γ = 0) are found as follows: For steady drift manoeuvres, the obtained forces
or moments are divided by cos β · sin β and the coefficients are taken from the
intersection at β = 0◦ of a linear or polynomial trend line through the data points.
For steady rotation, the same procedure is applied to the forces and moments divided
by γ.
2. Non-linear coefficients for pure transverse motion (β = 90◦ ) and pure rotation
(V = 0) are found using empirical relations (based on the work of Hooft [63],
e.g.). Currently, due to the unsteady nature of these manoeuvres and the com-
plexity of the flow around the hull, these motions are not solved using viscous flow
calculations.
4. The cross-terms, based on combined motions, are found in a similar way to step 3.
The known contributions of the coefficients from steps 1-3 are subtracted from the
calculated bare hull forces and the remainder is used to fit the cross-terms.
This approach is chosen to enable accurate modelling of the linearised behaviour for
course-keeping (step 1), realistic modelling of the harbour manoeuvring characteristics
(step 2), and accurate modelling of non-linear manoeuvres (steps 3 and 4). To ensure
appropriate responses for astern manoeuvres, it is assumed that the forces and moments
on the hull during astern manoeuvres are identical to those during ahead manoeuvring.
However, if different forces and moments are desired for astern motion, these can be ob-
tained by selecting the linear derivatives based on the sign of the longitudinal ship velocity,
for example, as follows for the coefficient Yβ0 , with Yβ,ahead
0
the appropriate coefficient for
0
ahead speed and Yβ,astern for astern speed:
Yβ0 = Yβ,ahead
0 0
· max(0, sign(cos β)) + Yβ,astern · max(0, −sign(cos β)) (3.42)
3.4 Conclusion
In this chapter, the mathematical model used by SurSim to simulate the manoeuvra-
bility of ships has been described. In the program, it is possible to provide user-defined
hydrodynamic coefficients for the bare hull forces and moments. This makes the program
well suited for the present work to investigate whether the use of viscous-flow calculations
can help in improving the prediction accuracy of simulations.
A procedure to derive the coefficients is proposed. This procedure is chosen to en-
able accurate modelling of the linearised behaviour for course-keeping as well as realistic
36 Chapter 3 – Mathematical model
Chapter 4
This chapter provides short background information regarding the viscous-flow solvers
used for the study discussed in this thesis. The grid procedures and the boundary condi-
tions used for the calculations are also presented in this chapter. Since the present work
focussed on the practical application of viscous-flow calculations and not on developing
the solvers or numerical procedures themselves, the theoretical details of the solvers will
not be discussed.
4.1 Background
Two different flow solvers have been used within this study. Initially, the calculations
were done with Parnassos, which is a solver optimised to calculate the flow around
ships in straight-ahead motion. For ships at drift angles or sailing in rotational motion,
it was found that considerable adjustments to the computational parameters (e.g. under-
relaxation) were required to reach the desired convergence levels. In the mean time, the
developments on the more general purpose solver ReFRESCO were started. With this
solver, it appeared to be easier to calculate the flow around ships in manoeuvring mo-
tions, especially for larger drift angles and rotation rates, and when using more advanced
turbulence models. Therefore, results obtained with both solvers are discussed in this
thesis.
4.2 Parnassos
Parnassos is one of MARIN’s in-house incompressible viscous-flow solvers. It is based
on a finite-difference discretisation of the Reynolds-averaged continuity and momentum
equations, using fully collocated variables and discretisation. The equations are solved
with a coupled procedure, retaining the continuity equation in its original form. Generally,
the governing equations are integrated down to the wall, i.e. no wall functions are used.
In Parnassos, multi-block structured grids are used. The implementation of the code is
optimised for solving the flow around ships with the mean flow directed along the ship’s
38 Chapter 4 – Viscous flow solvers
longitudinal axis. This makes the solver especially suitable for calculating and optimising
the resistance or propulsion of ships within a very short time frame, see e.g. Raven et al.
[114] or Van der Ploeg and Raven [154].
More detailed information about the solver can be found in Hoekstra and Eça [60],
Hoekstra [59], Raven et al. [113] or Eça and Hoekstra [40].
4.3 ReFRESCO
ReFRESCO is a MARIN spin-off of FreSCo [157], which was developed within the
VIRTUE EU Project together with Technische Universität Hamburg-Harburg (TUHH)
and HSVA. ReFRESCO is an acronym for Reliable and Fast Rans Equations solver for
Ships, Cavitation and Offshore. It solves the multi-phase unsteady incompressible RANS
equations, complemented with turbulence models and volume-fraction transport equations
for each phase. The equations are discretised using a finite-volume approach with cell-
centred collocated variables. The implementation is face-based, which permits grids with
elements with an arbitrary number of faces (hexahedrals, tetrahedrals, prisms, pyramids,
etc.). The code is parallelised using Message Passing Interface (MPI) and sub-domain
decomposition. Low order and higher-order spatial and temporal discretisation schemes
are available in the code. The equations are solved in a segregated approach, and the
pressure/velocity coupling is solved using the Semi-Implicit Method for Pressure Linked
Equations (SIMPLE) algorithm. The code is targeted, optimised and highly validated for
hydrodynamic applications, in particular for obtaining current, wind and manoeuvring
coefficients of ships, submersibles and semi-submersibles, see [159, 158, 51, 79]. Automatic
wall functions are available. In the present work, however, y2+ values below 1 are obtained
and all equations are integrated down to the wall.
as moving the grid in a rotational motion through a stationary flow (inertial reference
system), or by letting the flow rotate around the stationary ship (non-inertial reference
system). The latter is adopted in Parnassos and ReFRESCO. This approach has been
used by several authors, see for example section 3.2 in Batchelor [11] or section 1.15 in
Wesseling [162] and the applications to ships of e.g. Alessandrini and Delhommeau [3],
Cura Hochbaum [27] or Ohmori [99]. Using this system, the grid is attached to the hull
form and rotates with the ship. However, each water particle now should experience
centrifugal and Coriolis forces due to the rotation of the coordinate system. These forces
have to be added to the momentum equation as source terms.
with fi the force per unit volume that is exerted on a discrete flow volume.
Assuming a steady flow, the rotational motion is simulated by implementing the
centrifugal and Coriolis force as additional force terms, such that the modified momentum
equation reads:
ρuj ui,j + p,i − µui,jj + ρ u0i u0j = fi − ρ 2Ω × u −ρ Ω× Ω×r (4.2)
,j i i
with fi a remaining force term per unit volume (e.g. propeller forces), Ω the vector of
rotation, u = (u1 , u2 , u3 ) = (u, v, w) the velocity vector and r = (x − xR ) the radius of
rotation with xR the position of the centre of rotation. In the equation above,
the
Coriolis
force is represented by −2ρΩ × u while the centrifugal force is −ρΩ × Ω × r .
In ReFRESCO the Coriolis and
centrifugal
contributions are added to the external
force fi : i.e. fi = −ρ 2Ω × u − ρ Ω × Ω × r .
i i
4.6.2 ReFRESCO
For best performance of ReFRESCO, multi-block structured O-O grids3 are used in
general. Calculations for ships at drift angles or rotation rates are conducted by setting the
2
In an H-O grid, one set of grid lines is generally aligned with the incoming flow direction.
3
In O-O grids, grid lines are aligned perpendicular or parallel to the surface.
4.6 – Grid generation 41
Figure 4.1: Example grid, Parnassos, HTC, γ = −0.556 (coarsened for presentation)
boundary conditions to the proper inflow velocities. Unlike the grids used in Parnassos,
the computational domain does not need to be changed for each new calculation. For
simple geometries and grid topologies, the grids for ReFRESCO can be made using
in-house tools. For more generic applications, use is made of commercial grid generation
tools, such as ANSYS ICEM CFD, Numeca HEXPRESS, or PDC GridPro. In these
tools, the grid can be generated and boundary conditions can be defined. However, when
calculations are to be conducted for a range of drift angles or rotation rates, the re-
definition of inflow or outflow boundaries to accommodate each computational condition
requires user interaction. To avoid this, a new boundary condition has been implemented,
designated BCAutoDetect. With this type of boundary condition, it is possible to use a
single grid for different inflow angles, by automatically determining whether faces on the
exterior domain are inflow faces or outflow/pressure faces.
For each face on the exterior boundary, the angle between the velocity u and the
surface normal n is calculated, using:
!
u n
α = arccos · (4.4)
|u| |n|
Surface normals on the exterior boundary are always directed outward of the domain.
Based on the projected velocity, the boundary condition on a face is either set to inflow
(α ≥ αad ) or outflow/ pressure (α < αad ), with αad an adjustable angle. By default αad
is chosen to be 87◦ . This appeared to give slightly better convergence properties than an
angle of 90◦ .
To further facilitate the use of a single grid for all computations, the far field boundary
is generated as a cylindrical or spherical surface. An example grid for the HTC is given
42 Chapter 4 – Viscous flow solvers
in Figure 4.2.
In order to efficiently generate results for many drift angles, a procedure was imple-
mented to automatically increment the drift angle during a single simulation. Simulations
begin at a pre-set drift angle, until a specified number of iterations is reached, or when
the non-dimensional residuals are less than a specified convergence criterion. Next the
drift angle is incremented by ∆β, by changing the inflow conditions, and the solution is
continued from the solution at the previous drift angle. Starting the calculations from a
converged solution at a slightly different drift angle saves time compared to performing
each calculation separately. This procedure is repeated until the desired maximum inflow
angle is reached. In Figure 4.3, it is demonstrated that this approach (continuous lines)
provides the same results as those obtained with multiple single-drift angle calculations
(markers). This procedure was designated drift sweep and the application has already
been presented in e.g. Vaz et al. [158] and Bettle et al. [15].
Re = 14 × 106 , ∆β = 5◦ Re = 14 × 106 , ∆β = 5◦
0.015 0.005
istep = 1500 0.004 istep = 1500
0.01 istep = 750 istep = 750
istep = 3000 0.003 istep = 3000
0.005 steady 0.002 steady
0.001
0 0
N
Y
-0.001
-0.005 -0.002
-0.01 -0.003
-0.004
-0.015 -0.005
-20 -10 0 10 20 -20 -10 0 10 20
β [deg] β [deg]
4.8 Conclusion
In this chapter, the viscous-flow solvers used in this thesis have been presented. For
ship manoeuvres, not only the flow around the ship in oblique motion is of interest, but
also the flow around the ship when it performs a rotational motion. To compute the
flow around the ship in rotational motion, the flow solvers had to be modified. For this
work, the rotational motion was incorporated by using a non-inertial reference system
and supplementing the equations of motions with body forces representing the centrifugal
and Coriolis contributions to the flow.
44 Chapter 4 – Viscous flow solvers
To generate the grids for a range of drift angles and yaw rates, different approaches
were adopted depending on the flow solver used. For Parnassos, automated scripts
were developed with which the desired grids could be generated rapidly. For the more
generic solver ReFRESCO, all computations are conducted using one grid, but changing
the inflow angles and boundary conditions depending on the desired manoeuvring motion.
For this, a new boundary condition was developed, which removes the need to pre-process
each grid for each new manoeuvring condition.
To further improve the efficiency of the ReFRESCO calculations, a so-called drift
sweep procedure was developed to automatically calculate the forces and moments on the
ship for a range of drift angles.
The following chapter presents verification and validation of the viscous-flow calcula-
tions from which the hydrodynamic coefficients for the bare hull forces in the mathematical
model can be derived.
45
Chapter 5
In this chapter, the verification and validation (V&V) of viscous-flow calculations for ship
hulls in steady manoeuvring motion are discussed. Results of several test cases will be
presented to demonstrate that the flow solvers can be applied to compute the viscous-flow
around various hull forms and that accurate predictions of the forces and moments on the
manoeuvring ship can be obtained. The results for the HTC will be used in chapter 7
to derive hydrodynamic coefficients for the user-defined mathematical model for the hull
forces in SurSim, as presented in chapter 3.
merical error multiplied by a factor of safety Fs . The numerical error consists of three
components [115]: the round-off error; the iterative error and the discretisation error. The
viscous-flow solvers used in this study use double precision and therefore the round-off
error becomes negligible compared to the other errors. This means that the numerical
uncertainty USN is obtained by1
USN = UI + UG (5.1)
The uncertainty UI due to the iterative process depends on whether the equations are
sufficiently resolved. In simple flow problems, a reduction of the iterative error to machine
accuracy is sometimes possible, but for complex flows this may be too time consuming.
Estimations of UI based on changes in the solution during the iterative process are then
required. The discretisation uncertainty UG is obtained by estimating the discretisation
error δG and multiplying this with the factor of safety. When the differences between
solutions on progressively finer grids reduce (i.e. we are converging towards the exact
value), the discretisation error can be estimated with Richardson extrapolation (RE) and
the use of the Grid Convergence Index (GCI) [115]:
in which φi stands for the value of a considered quantity, φ0 is the estimate of the exact
solution, α is a constant, h is the typical cell size and p is the observed order of accuracy.
Additional or alternative error estimators are also used. With this estimator, the uncer-
tainty follows from: UG = |δG | · Fs . See Eça et al. [44, 45] for a complete discussion of the
procedure.
The aim of validation is to establish the comparison error E and the validation un-
certainty Uval and to obtain an interval that contains the modelling error δmodel . An error
δ is supposed to have a magnitude and sign, while an uncertainty U is used to designate
an interval containing an error of unknown magnitude and sign. The comparison error is
defined by the difference between the simulated value S and the experimental data value
D:
E =S−D (5.3)
This error contains all errors in the experiment as well as in the simulation. The validation
uncertainty is estimated by:
q
Uval = UD2 + USN
2 2
+ Uinput (5.4)
in which UD is the uncertainty in the experimental result and Uinput is the uncertainty in
the input parameters (e.g. fluid properties, geometry). Evaluation of Uinput is assumed to
be outside the scope of the present workq and is taken as zero for simplicity. Therefore
Uval can be calculated through Uval = UD2 + USN 2
. According to the American Society
1
According to the ASME V&V-20 procedure, Root mean square (RMS) addition cannot be applied
to the uncertainties, because of the dependency of e.g. UG on UI .
5.2 – KVLCC1, KVLCC2 and KVLCC2M 47
Of Mechanical Engineers (ASME) V&V-20 procedure [9], the modelling error δmodel is
defined by:
δmodel = E − (δSN + δinput − δD ) (5.5)
|E| >> Uval : the comparison error is governed by the modelling error (δmodel ≈ E), which
indicates that the model should be improved to reduce the comparison error.
|E| < Uval : the modelling error is within the noise imposed by the uncertainties contained
in the validation uncertainties. Conclusions about the modelling error can only be
drawn when the uncertainties are reduced. In this case, it is said that the model
and its solution are validated at a level of Uval .
In this thesis, solutions with |E| < Uval are indicated with a check-mark (✔), while
solutions with |E| >> Uval are not validated and therefore indicated with a cross (✘).
In the past, extensive procedures for V&V have been presented, such as the Interna-
tional Organization for Standardization (ISO) Guide to the Expression of Uncertainty in
Measurement [72], Stern et al. [133] and recently the ASME Standard for Verification and
Validation in Computational Fluid Dynamics and Heat Transfer [9]. In these procedures,
guidelines are given, but references to methods which can be used to obtain improved
uncertainty estimates are provided. The procedure based on a least-square version of the
Grid Convergence Index (GCI) as proposed by Eça et al. [44, 45] is followed in this thesis.
10−8
10−10
10−12
10−14
0 25000 50000
iteration number
The flow around the KVLCC2 hullform was also studied using ReFRESCO. One grid
with 5.388 × 106 cells was used for all drift angles, while the computations for different
rotation rates were made with a finer grid of 12.721 × 106 cells. Similar to the Parnassos
calculations, an undisturbed water surface and a fixed model attitude were assumed. The
Reynolds number was set to the slightly different value of 3.7×106 and the SST turbulence
model was used. The relation between the drift angle and the forces and moments on the
hull was obtained with the drift sweep procedure as introduced in section 4.6.2.
An overview of all experimental and computational results of the forces and moments
on the KVLCCs for captive conditions is given in the tables on page 151 through page 156.
spike in the convergence history. For each drift angle, the changes in Y force reduce to
below 1 × 10−10 . The convergence of the other force components is similar.
Table 5.1: Properties of grids for different drift angles, KVLCC2M, Parnassos
For a drift angle of 12◦ , a series of geometrically similar grids has been generated
in order to investigate the discretisation error. The grid coarsening has been conducted
in all three directions. For some of the grids, however, the distance of the first node
to the hull surface has been maintained in order to capture the velocity gradients in
the boundary layer. This might introduce scatter in the results due to non-geometric
similarity. Table 5.2 shows the number of nodes and y2+ values for these grids.
Table 5.2: Properties of grids for uncertainty analysis, KVLCC2M, Parnassos, β = 12◦
For grid 5, it was not possible to reach the required convergence criterion. Therefore
the results for this grid are dropped from further analysis.
For a drift angle of 12◦ , the predicted values of the friction (index f ) and pressure
(index p) components as well as the total force and moment coefficients for each force or
moment variable φ are given in Table 5.3 with their estimated uncertainties Uφ . Based
on an analysis of the results for each grid, it was decided to use the 6, 7 or 8 finest grids
for the uncertainty analysis. The number of grids ng used depended on the scatter in the
results for the coarsest grids. In the table, φ1 indicates the solution obtained on the finest
grid, φ0 the extrapolated solution and p the apparent order of convergence.
Item φ0 φ1 Uφ p
Xp - −2.32 × 10−3 32.3% 1
The absolute value of the uncertainty in the pressure components is larger than in the
friction components. The uncertainty in the longitudinal friction component Xf is about
one-third of the uncertainty in the longitudinal pressure component Xp . For the other
forces and moments, the uncertainty in the friction component is at least one order of
magnitude smaller than the uncertainty in the pressure component. Since most integral
forces and moments are dominated by the pressure component, this results in relatively
large uncertainties in the overall forces and moments.
In Figure 5.2, the convergence of the side force and yaw moment coefficients with grid
refinement is presented. It is seen that upon grid refinement, the estimated value for Y
(indicated by cfd) comes closer to the experimental value (indicated by exp). Considerable
scatter is visible in the data and therefore it is not easy to establish whether data points are
located in the asymptotic range of convergence. This is typical for this type of calculation,
5.2 – KVLCC1, KVLCC2 and KVLCC2M 51
as already observed previously by e.g. Eça et al. [38] and Hoekstra et al. [61].
Looking at the yawing moment N , the maximum difference between the estimated val-
ues for all grids is 5.1%. Because the difference between the estimated values is relatively
small and scatter on the data is present, monotonic divergence is found and extrapolation
to zero step size could not be made. This results in a relatively large uncertainty of 9.6%.
-0.054 -0.0225
-0.056 p=1.1 -0.023 U=9.6%
-0.058 U=5.6% -0.0235 exp
-0.024
-0.06 exp -0.0245
-0.062 -0.025
N
Y
-0.064 -0.0255
-0.066 -0.026
-0.068 -0.0265
-0.027
-0.07 -0.0275
-0.072 -0.028
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 0 0.5 1 1.5 2 2.5 3 3.5 4
relative step size relative step size
0.00
0.9
0.8
0.8
0.7
0.95
0. 0.2
4
0.9
0.7
0.95
0.02
0.4
0.
0.4
6
0.7
0.6
7
0.
0.
0.8
0.6
4
z/Lpp
0.8
0.04 0.9
0.80.6
0.40.7
0.8
0.
9
0.
0.2
95
0.7
0.06 95
0.
0.08 0.9
y/Lpp
0.00
0.95
0.6
0.9
7
0. 8
0.4 0.4 0.
0.70.8
0.8
0.4
0.6
0.02 0. 0.7
6
0.95
0.9
0.7
0.6
0.4
0.9
0.7
0.9
0.7
0.9
0.8
0.6
z/Lpp
0.04
0.70.80
0.9
0.9
0.8
5
0.4
0.6
0.4
0.9
.8
0.9
0.9
0.4
4
0.7.60.9
0.7 0.
0.06
0 .9
0
5
0.9
0.9
0.9
0.9
0.08 5
y/Lpp
0.00
0.6
0.
0.7
0.2
9
0.9
0.9
0.95
00 0.4 0.8
0.6.4.4
0.02
0.8
0.8
0.7
0.8
0.
7
0.
8
0.8
0.9
0.9
8
0.6
0.6
0.
0.6
0.6
z/Lpp
0.04
0.7
0.
0.6
0.9
0.95
7
0.9
0.9
0.
0.8
0.4 0.7 0.8
0.9
0.9
0.9
0.6
0.7
5 0.95
0.95
0.9
0.8
0.9
0.06 0.9
0.6
0.95
0.08 0.95
y/Lpp
Based on this comparison, it is seen that a trend may be found in the yaw moment N
against the drift angle due to the change in hull form: the yaw moment for the KVLCC1
is slightly higher than for the KVLCC2 and KVLCC2M hulls. This trend is visible in the
experiments and in the CFD results. However, no trend can be seen in the transverse force
Y or the force or moment against the yaw rate. Furthermore, one would expect the results
for the KVLCC2M to be closer to the KVLCC2 results than to the KVLCC1 results.
Therefore, it is concluded that the differences between the CFD results for the Y force
against the drift angle and Y and N against the yaw rate are within the accuracy of the
calculations and are not representative for the differences between the hull forms. Based
on the above observations, it is expected that the three hulls will have similar manoeuvring
behaviour. This has been demonstrated during the free sailing manoeuvring tests with
the KVLCC1 and KVLCC2, see MARIN Report No. 21571-1-SMB [87], in which it was
found that indeed there were no significant differences in the manoeuvring performances,
other than a small difference in the directional stability derived from pull-out tests.
A comparison between the Parnassos and ReFRESCO results for the KVLCC2
shows only marginal differences, even though a different turbulence model was used. Only
in the N moment against γ, a consistent difference between the results of the two solvers is
seen. The differences are within the uncertainty of the predictions and a reduction of the
uncertainties is required to investigate whether the differences are caused by modelling
errors. Therefore, no conclusions can be drawn about the influence of either the solver,
the grid layout or the turbulence model on the results.
5.2.5 Validation
The uncertainties in the measurements of the forces and moments are specified by Kume
et al. [82] and summarised in Table 5.4. The values for β = 12◦ were calculated by in-
terpolation between the uncertainties for β = 9◦ and β = 18◦ . Using the measurement
uncertainties and assuming that the simulation numerical uncertainty USN is only influ-
enced by the discretisation uncertainty UG (i.e. USN = UI + UG = UG ), Table 5.4 can
be constructed. The uncertainty Uinput due to uncertainties in the input parameters (e.g.
fluid properties,qgeometry) is assumed to be zero and therefore Uval can be calculated
through Uval = UD2 + USN 2
.
It is seen that for the longitudinal force X and yaw moment N the comparison error |E|
is smaller than the validation uncertainty Uval which means that the solution is validated
at levels of 8.6% and 10.4% respectively. These levels are judged to be good. If lower
validation levels are desired then the numerical uncertainty needs to be reduced. For the
transverse force Y validation is not achieved (|E| > Uval ), which indicates modelling errors.
It is seen that the magnitude of the Y force is under-predicted. Changes in turbulence
model or domain size or the inclusion of free surface may lead to improvements of the
comparison error.
54 Chapter 5 – Verification and validation of steady motion calculations
0 0.02
−0.01 Exp (KVLCC2)
Exp (KVLCC2M)
KVLCC1 (Parnassos)
X
−0.02
−0.1 0 KVLCC2 (Parnassos)
KVLCC2 (ReFRESCO)
−0.03 KVLCC2M (Parnassos)
Y
−0.2
N
−0.04
−0.02
0 10 20 30 40
β
−0.3 −0.04
−0.4 −0.06
0 10 20 30 40 0 10 20 30 40
β β
0.04 0.01
0.03 0
0.02 −0.01
Y
0.01 −0.02
0 −0.03
−0.01 −0.04
0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8
γ γ
Figure 5.4: Comparison between experiments and calculations, KVLCC hull forms (Lines rep-
resent CFD results)
5.3 HTC
The HTC has been described in section 2.5.4. In this section, RANS calculations for this
hull will be presented and compared to available experimental validation data. The captive
model tests have been conducted for speeds of 1.05 m/s and 1.89 m/s, corresponding to
10 kn and 18 kn on full scale. On model scale (λ = 24), this results in Reynolds numbers
of Re = 6.29 × 106 and Re = 12 × 106 , respectively. All calculations presented in this
thesis have been performed at model scale Reynolds numbers. To clearly distinguish the
different speeds used for the calculations or model tests and avoid confusion with the
speeds for other scale factors and the free sailing tests, calculations for the lower speed
will be identified with 10 kn, while the calculations for the higher speed will be identified
with 18 kn.
Calculations using Parnassos were made for the HTC hull form, using the Menter
one-equation turbulence model. Free surface and appendages were not modelled in the
calculations. These calculations were previously documented in VIRTUE Deliverable 3.1.1
[152]. Table 5.5 shows the conditions for the 10 kn computations. The domain sizes and
variation in the number of grid nodes in the stream-wise, normal and girth-wise (nξ , nη
and nζ ) directions are given, together with the y2+ values that were obtained.
Table 5.5: Properties of grids for different drift angles, HTC, Parnassos, 10 kn
Calculations have been conducted with ReFRESCO as well. Also in this case, free
surface and appendages were not modelled. All calculations were conducted for the 18 kn
condition and the SST turbulence model was selected for turbulence closure. One single
grid was used for all calculations. The steady drift calculations were performed with the
drift sweep procedure as introduced in section 4.6.2.
An overview of all experimental and computational results of the forces and moments
on the HTC for captive conditions is given in the tables on page 157 and page 158.
smaller than the discretisation uncertainty and therefore the iterative uncertainty UI can
be neglected, see Eça et al. [41].
For ReFRESCO, all calculations were run until the maximum non-dimensional resid-
ual of the pressure resp,max (the so-called L∞ norm) between successive iterations had
dropped well below 1 · 10−5 or when further iterative convergence was not obtained. The
changes in the non-dimensional integral quantities (forces and moments) were well below
1 × 10−7 . An example of the convergence history of the transverse force Y is given in
Figure 5.5. In this calculation, the drift angle is increased after a predefined number of
iterations from 0◦ to 30◦ in steps of 2.5◦ . Each increase of the drift angle is visible as the
spikes in the convergence history. For each drift angle, the changes in Y force reduce to
below 1 × 10−8 . The convergence of the other force components is similar.
10−8
10−10
10−12
10−14
0 25000 50000 75000 100000
iteration number
The absolute value of the uncertainty in the pressure components is larger than in the
friction components, as was already found during the uncertainty study for the KVLCC2M
hull form, see section 5.2. The uncertainty in the longitudinal friction component Xf is
about one-third of the uncertainty in the longitudinal pressure component Xp . For the
other forces and moments, the uncertainty in the friction component is at least one order
of magnitude smaller than the uncertainty in the pressure component. Since most integral
forces and moments are dominated by the pressure component, this results in relatively
large uncertainties in the overall forces and moments. In Raven, Van der Ploeg and
Table 5.6: Properties of grids for uncertainty analysis, HTC, β = 10◦ , Parnassos
Item φ0 φ1 Uφ p
Xp −5.33 × 10−4 −3.40 × 10−3 111.1% 0.67
Xf - −1.22 × 10−2 2.6% 2
Eça [113], an extensive study to improve the uncertainty and accuracy of the pressure
resistance component is presented.
-0.013 0.022
-0.014 p=0.9
0.02
p=1.4
-0.015 U=14.1% U=7.5%
-0.016 exp 0.018 exp
-0.017 0.016
-0.018
K
X
-0.019 0.014
-0.02 0.012
-0.021
-0.022 0.01
-0.023 0.008
0 1 2 3 4 5 6 0 1 2 3 4 5 6
relative step size relative step size
-0.035 -0.022
p=0.6 p=3.5
-0.04 U=19.9% -0.0225 U=1.2%
exp -0.023 exp
-0.045
-0.0235
N
Y
-0.05
-0.024
-0.055 -0.0245
-0.06 -0.025
0 1 2 3 4 5 6 0 1 2 3 4 5 6
relative step size relative step size
In Figure 5.6 the forces and moments on the ship are shown for the different grids.
The scatter in the results is much smaller than found for the KVLCC2M results. For a
relative step size below 3, the results appear to converge. The observed convergence rate
p, however, is found to be small for both X and Y (p = 0.9 and 0.6 respectively). Due to
the slow convergence, the difference between the extrapolated value φ0 for zero step-size
and the value φ1 is large and hence the uncertainty is relatively large.
Noteworthy is the fact that based on the trends with the current grids, the estimations
(indicated by cfd) for X, Y and N for increasing numbers of grid nodes do not converge
to the experimental values (indicated by exp). This may be caused by either modelling
errors or by uncertainties in the experimental values.
A similar grid study was conducted for ReFRESCO. Using an automatic procedure
in GridPro, a series of geometrically similar grids was generated, using grid coarsening in
all three directions for each block in the grid. For each grid, the number of cells in the
grid ncells and the number of faces on the hull surface nhull are given in Table 5.8, which
includes also the y2+ values that were obtained during the calculations at drift angles of
0◦ and 30◦ and a non-dimensional rotation rate of 0.4. For each grid, the full range of
drift angles between 0◦ and 30◦ was calculated.
For drift angles of 10◦ and 30◦ , the uncertainty estimates are presented in Table 5.9
and Table 5.10 respectively. Based on an analysis of the results for each grid, it was
decided to use the four finest grids for the uncertainty analysis. The number of grids ng
5.3 – HTC 59
used was chosen based on the scatter in the results for the coarsest grid.
In Figure 5.7 and Figure 5.8 the forces and moments on the ship are graphically
presented for the different grids. The scatter in the results is relatively large and the
results are less consistent than those obtained with Parnassos. For a relative step size
of 1.6 or below, the results appear to converge. The observed convergence rate p, however,
is found to be large for all force components, hence the uncertainty is relatively large due
to the use of a larger safety factor Fs . Finer grids are required to reduce the uncertainty
in the calculations.
Table 5.8: Properties of grids for uncertainty analysis, HTC, 18 kn, ReFRESCO
Item φ0 φ1 Uφ p
Xp - −3.15 × 10−3 44.0% 1
Item φ0 φ1 Uφ p
Xp - 6.48 × 10−3 109.5% 1
-0.0135 0.021
U=6.6% U=11.6%
-0.014 0.02
exp exp
-0.0145 0.019
-0.015
0.018
K
X
-0.0155
-0.016 0.017
-0.0165 0.016
-0.017 0.015
0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3
relative step size relative step size
-0.025 -0.0205
U=31.0% -0.021 U=6.9%
-0.03 exp exp
-0.0215
-0.035 -0.022
-0.04 -0.0225
N
Y
-0.045 -0.023
-0.0235
-0.05 -0.024
-0.055 -0.0245
0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3
relative step size relative step size
0 0.124
-0.002 U=90.0% p=7.1
0.122
-0.004 exp U=3.5%
-0.006 0.12 exp
-0.008 0.118
K
X
-0.01 0.116
-0.012 0.114
-0.014
-0.016 0.112
-0.018 0.11
0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3
relative step size relative step size
-0.21 -0.068
-0.22 p=3.6 -0.07 p=6.2
-0.23 U=9.1% -0.072 U=7.5%
-0.24 exp -0.074 exp
-0.076
-0.25
N
Y
-0.078
-0.26 -0.08
-0.27 -0.082
-0.28 -0.084
-0.29 -0.086
0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3
relative step size relative step size
Table 5.11: Properties of grids for uncertainty analysis, HTC, γ = −0.2, Parnassos
The predicted values of the friction (index f ) and pressure (index p) components as
well as the total force and moment coefficients are presented in Table 5.12 with their
estimated uncertainties. Based on an analysis of the results for each grid, it was decided
to use the four finest grids (grids 3, 4, 6 and 8) for the uncertainty analysis.
Item φ0 φ1 Uφ p
Xp −1.83 × 10−3 −2.07 ×10−3 25.8% 5.46
Xf - −1.17 × 10−2 3.3% 1
Similar to what was found for steady drift, the absolute uncertainty in the pressure
5.3 – HTC 63
components is larger than in the friction components. Compared to the calculations for
steady drift, the relative uncertainties for the rotational motion results are in most cases
smaller.
In Figure 5.9 the longitudinal force X, transverse force Y , heel moment K and yawing
moment N are given for the different grids. It is seen that the results do not differ much
between the individual results, but convergence is not always found due to scatter. For a
relative step size below 3, reasonably consistent results are however found.
-0.013 -0.001
p=6.1 -0.0012 p=5.6
-0.0135
U=5.1% -0.0014 U=12.6%
-0.014 exp -0.0016 exp
-0.0145 -0.0018
K
X
-0.015 -0.002
-0.0155 -0.0022
-0.0024
-0.016 -0.0026
-0.0165 -0.0028
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5
relative step size relative step size
-0.002 0.009
U=6.1% 0.0085 U=10.9%
-0.003 exp exp
0.008
-0.004 0.0075
-0.005 0.007
N
Y
-0.006 0.0065
0.006
-0.007 0.0055
-0.008 0.005
0 0.5 1 1.5 2 2.5 3 3.5 4 0 0.5 1 1.5 2 2.5 3 3.5 4
relative step size relative step size
Using the grids summarised in Table 5.8, the uncertainties in the ReFRESCO results
for non-dimensional rotation rates of γ = 0.2 and γ = 0.4 were determined. Based on
an analysis of the results for each grid, it was decided to use the four finest grids for the
uncertainty analysis, which is presented in Table 5.13 and Figure 5.10 for γ = 0.2 and
Table 5.14 and Figure 5.11 for γ = 0.4.
The convergence with grid refinement for rotational motion is slightly better (more
realistic apparent orders of convergence) than the grid sensitivity for steady drift presented
in Table 5.9.
Item φ0 φ1 Uφ p
Xp - −2.39 × 10−3 30.1% 2
-0.0115 0.004
-0.012 p=0.5 p=1.3
U=12.7%
0.0035 U=41.7%
-0.0125
-0.013 exp 0.003 exp
-0.0135 0.0025
K
X
-0.014 0.002
-0.0145
-0.015 0.0015
-0.0155 0.001
0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3
relative step size relative step size
0.01 -0.0055
0.009 p=1.6 -0.006 p=0.8
0.008 U=19.1% -0.0065 U=24.9%
0.007 exp -0.007 exp
-0.0075
0.006
N
Y
-0.008
0.005 -0.0085
0.004 -0.009
0.003 -0.0095
0.002 -0.01
0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3
relative step size relative step size
Item φ0 φ1 Uφ p
Xp −3.37 × 10−3 −3.16 × 10−3 20.5% 2.19
Xf −1.20 × 10−2 −1.16 × 10−2 5.1% 0.96
X −1.52 × 10−2 −1.47 × 10−2 4.3% 1.66
Yp 1.61 × 10−2 1.50 × 10−2 9.7% 2.02
Yf 3.39 × 10−4 3.52 × 10−4 4.9% 1.98
Y 1.64 × 10−2 1.53 × 10−2 9.3% 2.02
Zp - 7.30 × 10−2 7.2% 2
-0.012 0.0048
-0.0125 p=1.7 0.0046 p=2.0
-0.013 U=4.3% 0.0044 U=9.3%
-0.0135 exp 0.0042 exp
-0.014 0.004
-0.0145 0.0038
K
X
-0.015 0.0036
-0.0155 0.0034
-0.016 0.0032
-0.0165 0.003
-0.017 0.0028
0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3
relative step size relative step size
0.017 -0.016
0.016 p=2.0
-0.017 p=1.4
0.015 U=9.3% U=9.1%
0.014 exp
-0.018 exp
0.013 -0.019
0.012
N
Y
0.011 -0.02
0.01 -0.021
0.009
0.008 -0.022
0.007 -0.023
0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3
relative step size relative step size
0.02
0.00
0.8
0.2
0.4 00
0..47.7
00
0.4
0.9
5
1 8
0.
.9
.6.06.
.9 0
2
.60.2
0.02
0.9
0.95
00.6
0.8
9
0.4
0.
1
0.4
0.8
0.2 00.7
z/Lpp
00.7.7
0.04
.70.95
00.8.42
00.4.8 1
0..66
0.9
0.06
0.02
0.9
0.
95
0.08
1
ReFRESCO fine, x=0.48Lpp, γ=0.4 Parnassos MNT norud, x=0.48Lpp, γ=0.4
0.
95
0.10
0.16 0.14 0.12 0.10 0.08 0.06 0.04 0.02 0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14 0.16
y/Lpp
Figure 5.12: Comparison between ReFRESCO and Parnassos, HTC, x = 0.48Lpp , 18 kn
(solid lines: ReFRESCO, dotted lines: Parnassos)
neglect of the free water surface and may lead to errors in the prediction of the forces on
the hull.
The experimental data of the wake field in a plane 0.48Lpp aft of midship have kindly
been made available by HSVA2 . In the upper half of Figure 5.13 a comparison of the
non-dimensional axial velocity fields between the experiments (dotted lines) and the cal-
culations (solid lines) for a non-dimensional yaw rate γ of 0.4 is presented. The non-
dimensional axial velocity along a horizontal line in this plane at z = 0.05Lpp is given
in the lower half of the figure. This figure shows that in most parts of the plane, the
viscous-flow calculations correspond reasonably well with the experiments, especially at
port side (windward). At the leeward side, the wakes of three different vortices are found
in the Parnassos results: around y = 0.01Lpp the vortex generated at the stern, around
y = 0.04Lpp the vortex generated at the bow and at y = 0.07Lpp the vortex generated
at the starboard bilge. In the ReFRESCO results and in the experiments, only the
wakes of two vortices are seen. Based on an analysis of the flow along the length of the
ship calculated by ReFRESCO, it was observed that the bow vortex merges with the
vortex generated at the bilge. Apparently, this effect is not captured by Parnassos. The
difference between the experiments and the calculation is mainly caused by the relatively
coarse grid used for the Parnassos calculations.
Overall, it is concluded that the flow field at the propeller plane is predicted reasonably
well.
2
The actual immersion of the PIV probe, which served as reference for the position of the field of view
during the wake field measurements, was unknown, because the immersion was measured before the final
adjustment of the model condition when the model still had a considerable trim by the stern due to the
weight of the PIV equipment. Based on the apparent location of the propeller hub and the ship hull, a
vertical shift by about 0.011Lpp has been applied.
5.3 – HTC 67
0.02
0.00
0.
0.8
0.7
0.9
0.
95
1 0.6
0.
8
0.6 8
95
0.
7 0.
1
0.7
0.9
0.7
0.02
0.9 0.9
0.9
0.8
0.9
0.7
0.
0.8
95
0.6
0.9
1
0.6 1
z/Lpp
0.8
0.8
0.04 0.6
0.07.6
0.70.8
0.95
1
0.6
0.9
0.8 1
0.9
0.06 5
0.9
0.9
5
0.08 1
0.00
0.9
0.8
0.95
0.95
0.6
0.7 0.9 1 0.8
0.9
0.9
00..8
8 0.6
0.02
1
1
0.9
0.6 0.7
00.
0.8
.7 70.6
1
9
0.95
0.9
0.95
0.
z/Lpp
0.95
0.8
0.
0.04 6
1
0.
95
0.9
.4
0.9
00.6
1
0.
0.8 0.8
7
0.
1
0.9
0.06
1
0.08 0.
95
1.0
0.8
0.6
u
0.4
0.2
1.0
0.8
0.6
u
0.4
0.2
y/Lpp
Figure 5.13: Comparison with experiments, HTC, x = −0.48Lpp , 18 kn (solid lines: calcula-
tions, dotted lines: experiments)
68 Chapter 5 – Verification and validation of steady motion calculations
−0.005 0.12
−0.005 Exp (HSVA)
Parnassos
0.1
ReFRESCO
−0.01
−0.01 0.08
X
K
X
0.06
−0.015
−0.015 0.04
−0.02
0 5 10 15 20 25 30
0.02 β
−0.02 0
0 10 20 30 0 10 20 30
β β
0 0
−0.05 −0.002
−0.1 −0.004
M
Y
−0.15 −0.006
−0.2 −0.008
−0.25 −0.01
0 10 20 30 0 10 20 30
β β
0.3 0
0.25
−0.02
0.2
N
−0.04
Z
0.15
−0.06
0.1
0.05 −0.08
0 10 20 30 0 10 20 30
β β
Figure 5.14: Comparison between experiments and calculations, HTC, 18 kn, steady drift
70 Chapter 5 – Verification and validation of steady motion calculations
−3
−0.012 x 10
6
5
−0.014
4
X
−0.016
K
3
2
−0.018
1
−0.02 0
0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8
γ γ
−3
0.025 x 10
0
0.02
−1
0.015
Y
−2
0.01
−3
−0.005 Exp (HSVA)
0.005 Parnassos
ReFRESCO
0 −0.01
−4
0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8
X
γ γ
−0.015
0.09 0
−0.02
−0.0050 5 10 15 20 25 30
0.08 β
−0.01
0.07
Z
−0.015
−0.02
0.06
−0.025
0.05 −0.03
0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8
γ γ
Figure 5.15: Comparison between experiments and calculations, HTC, 18 kn, steady yaw
5.3 – HTC 71
5.3.6 Validation
The uncertainties in the measurements of the forces and moments have not been specif-
ically determined. However, an estimate of the uncertainty UD can be made based on
results obtained for similar test conditions during measurements at the start of VIRTUE
and measurements conducted later in the project. Some tests were repeated and for
each test, the uncertainty was estimated using an arbitrary factor of safety of 1.25,
i.e. UD = 1.25 × abs (φtest1 − φtest2 ). The overall data uncertainty for the HTC ex-
periments is taken from the average of the data uncertainties obtained for several re-
peat tests. With these estimated data uncertainties and assuming that the simulation
numerical uncertainty USN is only influenced by the discretisation uncertainty UG (i.e.
USN = UI + UG = UG ), Table 5.15 and Table 5.16 can be constructed.
The Parnassos result for β = 10◦ shows that for the longitudinal and transverse
forces X and Y the comparison errors |E| are smaller than the validation uncertainties
Uval which means that the solution is validated at levels of 13.3% and 21.1% respectively.
These levels are judged to be relatively high and indicate that the numerical uncertainty
needs to be reduced if lower validation levels are desired. For the heel moment K and
the yawing moment N validation is not achieved (|E| > Uval ), which indicates modelling
errors. Changes in turbulence model or domain size or the inclusion of free surface may
lead to improvements of the comparison error.
The ReFRESCO result for Y for β = 10◦ is validated at a level of 27.2%, which is
high. The numerical uncertainty USN needs to be reduced to reduce the validation level.
The yaw moment N is validated at an acceptable level of 8.3%. The other components
are not validated and indicate modelling errors. For the higher drift angle β = 30◦ , Y and
N are validated at levels of 12.8% and 9.5% respectively, which are reasonable values.
For the steady rotation case, γ = 0.2 with Parnassos (the results have been ob-
tained by mirroring the γ = −0.2 results for comparison with the ReFRESCO results),
the comparison errors |E| are larger than the validation uncertainties for all forces and
moments. Since experimental results for the 10 kn condition with γ = 0.2 were not avail-
able, the experimental values for the 18 kn condition were used for comparison. This
mainly introduces an error in the longitudinal force X due to a difference in the frictional
resistance and therefore a large comparison error in X is to be expected. For the other
components, the influence of the different speed is expected to be of less influence and
therefore the validation should still provide insight into the accuracy of the calculations.
Even though a comparison between two speed conditions is made, it is concluded based
on the validation that the large values of the comparison errors indicate that modelling
errors are present in the simulation results.
For ReFRESCO, the comparison errors |E| for γ = 0.2 are smaller than for Par-
nassos. Together with the higher numerical uncertainties USN , this leads to validation
of X, Y , K and N at levels of 11.4%, 20.6%, 66.8% and 22.2%, respectively. These levels
are high and USN needs to be reduced to obtain lower validation levels.
The ReFRESCO results for γ = 0.4 for Y , K and N are validated at levels of 12.7%,
72 Chapter 5 – Verification and validation of steady motion calculations
10.5% and 10.5% respectively, which are judged to be reasonable. The comparison errors
|E| for all components are found to be small. For X, modelling errors appear to be
present, which might be improved by incorporating free surface deformation.
more than two to three orders of magnitude smaller than the discretisation uncertainty
and therefore the iterative uncertainty UI can be neglected, see Eça et al. [41].
id β nξ nη nζ hi hi Nodes Comment
(β = 0 ) (β = 18◦ ) ×10−3
◦
For Re = 14 × 106 , the discretisation error has been investigated. In Table 5.18 and
Figure 5.16, the results for β = 0◦ are presented. The graphs show that scatter exists in
the data: the data points are not exactly aligned along the curve. Reasons for this might
be e.g. the non-evenly spaced cell nodes, the use of numerical limiters or lack of perfect
geometrical similarity between the grids.
For this high Reynolds number, i.e. when convection dominates, and when using an
unstructured-grid Quadratic Upwind Interpolation for Convective Kinematics (QUICK)
scheme for convective fluxes, it is expected that ReFRESCO will be second order ac-
curate [45]. The observed order of convergence p depends on the force component under
consideration. For the friction force, a value just below 1 is found, indicating that the
convergence with grid refinement follows a linear order of accuracy. For the other com-
ponents, a much higher order is found, which is most probably caused by scatter and
insufficiently fine grids.
In the previous study by Toxopeus and Vaz [151], the convergence appeared to be
better. However, comparing the old (FreSCo) results with the new (ReFRESCO)
results with a finer grid added, it is seen that now the four finest grids show a more
consistent trend than the four finest grids in the previous study. The present results are
therefore judged to be more reliable. The overall uncertainty U in X is 4.5% which is
judged to be small.
In Table 5.19 and Figure 5.17, the results for β = 18◦ are presented. In this case, the
apparent order of convergence p ranges from 0.74 for Xf to 6.07 for Xp . This indicates
5.4 – DARPA SUBOFF 75
Item φ0 φ1 Uφ p 1
−4 −4 Oscillatory convergence
Xp −1.22 × 10 −1.29 × 10 12.2% 6.71 2
−3 −4 Monotonic divergence
Xf −1.00 × 10 −9.67 × 10 4.4% 0.97 3
−3 −3 Oscillatory divergence
X −1.11 × 10 −1.10 × 10 4.5% 5.05
-0.0001 -0.00084
-0.00015 p=6.7 -0.00086 p=1.0
-0.0002 U=12.2% -0.00088 U=4.4%
-0.00025 -0.0009
-0.0003 -0.00092
Xp
Xf
-0.00035 -0.00094
-0.0004 -0.00096
-0.00045 -0.00098
-0.0005 -0.001
-0.00055 -0.00102
0 1 2 3 4 5 6 0 1 2 3 4 5 6
relative step size relative step size
-0.0007
-0.0008 p=5.0
-0.0009 U=4.5%
exp
-0.001
-0.0011
X
-0.0012
-0.0013
-0.0014
-0.0015
0 1 2 3 4 5 6
relative step size
that a finer grid needs to be used to obtain a solution closer to the so-called asymptotic
range, where the order of convergence will be equal to or lower than the order of the
discretisation scheme.
For the components of the transverse force Y and yaw moment N the apparent orders
of convergence are between 0.62 and 1.21, which may indicate that for the transverse force
and yawing moment the grid density is closer to the asymptotic range. The uncertainty
in X is found to be relatively large. The large value is caused by the fact that for the
overall force X monotonic convergence was not obtained. However, the value is acceptable
from an engineering viewpoint. The uncertainty in the overall transverse force or yawing
moment is judged to be small.
Item φ0 φ1 Uφ p
Xp 2.70 × 10−4 2.41 × 10−4 31.7% 6.07
Xf −1.13 × 10 −3 −1.06 × 10−3 7.9% 0.74
X - −8.22 × 10−4 11.3% 2
1 Oscillatory convergence
Yp 5.21 × 10−3 5.36 × 10−3 3.6% 1.17 2 Monotonic divergence
Yf 3.26 × 10−4 3.01 × 10−4 10.3% 0.62 3 Oscillatory divergence
Y 5.53 × 10−3 5.66 × 10−3 3.0% 1.21
Np 3.45 × 10−3 3.40 × 10−3 1.9% 0.81
Nf - 1.75 × 10−5 8.2% 2
-0.0007 0.0074
-0.0008 U=11.3% 0.0072 p=1.2
exp 0.007 U=3.0%
-0.0009 0.0068 exp
-0.001 0.0066
0.0064
X
-0.0011 0.0062
-0.0012 0.006
0.0058
-0.0013 0.0056
-0.0014 0.0054
0 0.5 1 1.5 2 2.5 3 3.5 4 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5
relative step size relative step size
0.0036
p=0.8
0.0035
U=2.0%
0.0034 exp
0.0033
N
0.0032
0.0031
0.003
0.0029
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5
relative step size
sented in this section were made according to these instructions. The experimental values
are obtained from flow field and pressure measurements conducted by Huang et al. [68].
These experiments were conducted at a Reynolds number of 12 × 106 . The first and sec-
ond cases defined in the instructions are comparisons of the pressure Cp and friction Cf
coefficients along the hull, see Figure 5.18.
These graphs show that the differences in pressure coefficient between the results are
negligible. A very small difference between the SST and SA results is found at the stern,
which explains the difference in the longitudinal pressure coefficients Xp . For the skin
friction coefficient, it is seen that the results with the SA model are in general slightly
closer to the experimental data than the SST results. The differences between the results
explain the differences in forces found in Table 5.20.
5.4 – DARPA SUBOFF 77
1.5
Exp
ReFRESCO−SST
1 ReFRESCO−SA
Cp
0.5
−0.5
0 0.2 0.4 0.6 0.8 1
x/L
oa
−3
x 10
7
Exp
6 ReFRESCO−SST
ReFRESCO−SA
5
4
Cf
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
x/Loa
Figure 5.18: Pressure (top) and friction (bottom) coefficients along the hull, β = 0◦
The predicted distribution of the pressure coefficient is close to the experiments. The
trends in the predicted distribution of the friction coefficient correspond well to the trends
found in the experiments. Although some discrepancies at the bow and stern area are
found, it is concluded that the prediction of the pressure and skin friction coefficients is
good. It is noted that the discrepancies at the bow and stern were also present in the
calculations by Bull [19] and Yang and Löhner [166] and in all results submitted for a
collaborative CFD study within the SHWG [146].
The difference between the SST and SA results for the streamwise Vx and radial
velocities Vr at x = 0.978Loa in the aft part of the hull, see Figure 5.19 (top) is considered
to be negligible. Comparing the computed results with the experiments, it is observed
that the trends in the development of the boundary layer are very well predicted by both
solvers, but quantitative discrepancies are seen. Especially the magnitudes of the radial
velocities are different. It is seen that in the experiments the radial velocity changes sign
between (r − R0 )/Rmax = 2 and (r − R0 )/Rmax = 0.8, suggesting outward radial flow in
the far field. This may be caused by the use of an open-jet wind tunnel.
In this study, also the correlation between the measured and the predicted Reynolds
0 0
shear stresses is investigated by comparison of −VVx2Vr . Following the eddy-viscosity as-
0
78 Chapter 5 – Verification and validation of steady motion calculations
In Figure 5.19 (bottom), the Reynolds shear stresses for the aft-most longitudinal sta-
tion are presented. It is observed that the curve representing the SST results corresponds
very well with the measurements. The results using the SA turbulence model are also
close to the measurements, but under-predict the peak of the distribution.
2
Exp
ReFRESCO−SST
1.5 ReFRESCO−SA
(r−R0)/Rmax
0.5
Vr Vx
0
−0.2 0 0.2 0.4 0.6 0.8 1
Velocity/V0
−3
x 10
1.5
1
−Vx’Vr’/V20
Exp
ReFRESCO−SST
ReFRESCO−SA
0.5
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
(r−R0)/Rmax
Figure 5.20 presents comparisons of the pressure Cp and friction Cf coefficients along
the hull at the leeward plane of symmetry. Figure 5.21 shows the axial Vx , tangential Vα
and radial Vr velocities (top part of the figure) and Reynolds shear stress (lower part of
the figure), given for the leeward symmetry plane located at x = 0.978Loa . These graphs
show that the distribution of the pressure coefficient along the length of the ship and the
velocity distribution at the stern is quite well represented. The difference between the
5.4 – DARPA SUBOFF 79
SST and SA results is considered to be small. However, with the SA turbulence model,
the radial velocity Vr appears to be too negative compared to the SST results and the
results obtained using other solvers and turbulence models during the SHWG CFD study
[146]. The distribution of the Reynolds shear stress shows reasonable correspondence with
the measurements.
1.5
Exp
ReFRESCO−SST
1 ReFRESCO−SA
Cp
0.5
−0.5
0 0.2 0.4 0.6 0.8 1
x/Loa
−3
x 10
7
5 ReFRESCO−SST
4 ReFRESCO−SA
Cf
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
x/Loa
Figure 5.20: Pressure (top) and friction (bottom) coefficients along the hull, β = 2◦ (leeward
meridian)
2
Exp
ReFRESCO−SST
1.5 ReFRESCO−SA
(r−R0)/Rmax
0.5
Vr Vt Vx
0
−0.2 0 0.2 0.4 0.6 0.8 1
Velocity/V0
−3
x 10
1.5
1
−Vx’Vr’/V20
Exp
ReFRESCO−SST
ReFRESCO−SA
0.5
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
(r−R0)/Rmax
Figure 5.21: Velocities (top) and Reynolds stresses (bottom) at x = 0.978Loa , β = 2◦ (leeward
meridian)
The longitudinal force components obtained from the calculations for β = 0◦ are given
in Table 5.20. As can be expected for submarine hull forms, the largest part (about 90%)
of the total resistance is caused by friction. This means that for the bare hull, the form
factor is relatively low, i.e. (1 + k) = X/Xf = 1.13 for SST and 1.07 for SA. From the
experiments, the form factor is estimated to be (1 + k) = X/Xf (ITTC) = 1.13. This is a
normal value for a bare hull submarine.
The comparison error E between the ReFRESCO prediction of X and the mea-
surement is about 3.7%, which is judged to be good for practical applications when also
the uncertainty in the experimental data is taken into consideration. It is found that
the total resistance predicted using the SST turbulence model is slightly higher than the
experimental value, while the SA results are slightly lower. The skin friction coefficient
predicted using SA is lower than the coefficient found using SST, as can also be observed
in Figure 5.18. In the aft ship, the pressure coefficient predicted using SA is marginally
higher than the pressure predicted with the SST model. This explains the lower pressure
resistance found in the SA results.
Figure 5.22 presents the force and moment components obtained from the calculations
and the values from the experiments for oblique inflow. In Tables 5.21 through 5.23 the
5.4 – DARPA SUBOFF 81
5.4.5 Validation
The uncertainties in the measurements of the forces and moments have not been deter-
mined. However, validation of the solution can still be performed, when a data uncertainty
UD is assumed. To obtain an estimate of the uncertainty in the experimental data, the
uncertainty UD in the experimental data is estimated using the difference between two
measurements for the same condition and a factor of safety of 1.25. For example, the
uncertainty in the longitudinal force X for β = 0◦ is estimated by:
For other incidence angles, the same procedure can be applied. With this estimated un-
certainty and assuming that the simulation numerical uncertainty USN is only influenced
by the discretisation uncertainty UG (i.e. USN = UI + UG = UG ), Table 5.24 can be con-
structed. For β = 18◦ , it is seen that for the longitudinal force X the comparison error
|E| is smaller than the validation uncertainty Uval which means that the solution of X is
validated at a level of 27.2%. This level is judged to be high and indicates that especially
the experimental uncertainty needs to be reduced if lower validation levels are desired.
For all other forces and moments validation is not achieved (|E| > Uval ), which in-
dicates modelling errors. The magnitude of the Y force is under-predicted, while the
magnitude of the N moment is over-predicted. Changes in turbulence model or in the
domain size but also using higher grid densities in the wake or time accurate solution
procedures may lead to improvements of the comparison error.
5.4 – DARPA SUBOFF 83
−3
x 10
1.5
Exp
1 ReFRESCO−SST
0.5
X
−0.5
−1
−1.5
0 5 10 15 20 25 30 35 40
β
0.005
−0.005
Y
−0.01
−0.015 Exp
ReFRESCO−SST
−0.02
−0.025
0 5 10 15 20 25 30 35 40
β
−3
x 10
1
−1 Exp
−2 ReFRESCO−SST
N
−3
−4
−5
−6
0 5 10 15 20 25 30 35 40
β
Figure 5.22: Comparison between experiments and calculations, SUBOFF, steady drift
5.5 – Walrus 85
5.5 Walrus
In work conducted for the Royal Netherlands Navy (RNLN) into the influence of the
seafloor on the manoeuvrability of submarines, validation studies were performed by Bet-
tle under the supervision of Toxopeus [15]. The flow around the Walrus bare hull form,
with deck and sail was computed with ReFRESCO. Other appendages were not included
in the study. The SST turbulence model was used and all equations were integrated down
to the wall (y2+ values were below 1). Calculations have been done for a range of clear-
ances c between the sea bottom and the submarine. In this thesis, only the deep water
calculations are considered. For the steady drift calculations, the drift sweep procedure
as introduced in section 4.6.2 was used.
It should be noted that for the Walrus results the coordinate system and non-di-
mensionalisation of the forces and moments differ from those adopted in the rest of this
thesis: all forces and moments are made non-dimensional using a reference area of L2pp
instead of Lpp T . The origin of the right-handed coordinate system is located at the
intersection of the longitudinal axis of symmetry of the hull, midship and centre-plane.
The experiments for the Walrus-class submarine were conducted by the David W.
Taylor Naval Ship Research and Development Center (DTNSRDC) [34]. Experimental
results were obtained for the three configurations of the early Walrus design listed in
Table 5.25.
Table 5.25: Designations and descriptions of Walrus experimental configurations [34]
It should be noted that the design of the Walrus as used for the DTNSRDC model
tests differs slightly from the real Walrus class submarine design, which was used for the
present study. The largest differences compared to the real design are a slightly smaller
length (0.9%Loa ) and the absence of the Toekan (exhaust diffuser). It is expected that
this discrepancy will have only a small effect on the overall forces and moments.
The calculations were performed at Re = 5.2 × 106 (to resemble the condition of
the free sailing experiments) whereas the DTNSRDC experiments were conducted at two
higher Reynolds numbers: 9 million and 14 million. The main effect of Reynolds number
in this range is to reduce the viscous drag. In order to better compare the results, the
axial force evaluated in the calculations at zero drift angle was scaled to Re = 14 million
using the ITTC 1957 friction line:
Cf,ITTC,Re=14×106
XRe=14×106 = Xp + Xf,Re=5.2×106 · (5.10)
Cf,ITTC,Re=5.2×106
This was done for the overall force on the hull and sail as well as on the hull surface alone.
86 Chapter 5 – Verification and validation of steady motion calculations
2 2
u u
0 v 0 v
w w
-2
p -2
p
-4 -4
-6 -6
-8 -8
c = 11.9 B c = 11.9 B
-10 α = 0 , β = 0 -10 α = 0 , β = 0
o o o o
6 6
1.9x10 Cell 1.9x10 Cells
-12 -12
0 2000 4000 6000 8000 0 2000 4000 6000 8000
Iterations Iterations
Excellent convergence was achieved, with L∞ and L2 norms of the residuals dropping
below 10−5 and 10−7 , respectively, in all cases of β = 0◦ . The integrated forces and
moments were unchanging to 7 significant digits (the precision with which this data was
written to the results file) for the last several hundred iterations. Figure 5.23 shows the
convergence histories of a selected calculation with a drift angle of zero.
For non-zero drift angles, the L2 residuals dropped at least four orders of magnitude.
The effect on the integral quantities is small: the non-dimensional changes in the forces
and moments were well below 1×10−8 . This is more than two to three orders of magnitude
smaller than the discretisation uncertainty and therefore the iterative uncertainty UI can
5.5 – Walrus 87
be neglected, see Eça et al. [41]. An example of the convergence history of the non-
dimensional Y force for the drift-sweep calculation is shown in Figure 5.24. Each spike in
the line indicates the beginning of a new drift angle.
Item φ0 φ1 Uφ p
Xp −2.09 × 10−4 −2.16 ×10−4 11.2% 7.56
Xf −1.32 × 10−3 −1.31 × 10−3 1.0% 1.98
X - −1.53 × 10−3 1.7% 2
1
−5 Oscillatory convergence
Zp −2.62 × 10 −2.72 × 10−5 9.9% 0.49 2 Monotonic divergence
Zf - −4.86 × 10−6 38.5% 1
3 Oscillatory divergence
Z −3.04 × 10−5 −3.20 × 10−5 10.2% 0.69
Mp −5.47 × 10−5 −5.44 × 10−5 2.4% 5.45
Mf 2.39 × 10−5 2.38 × 10−5 2.1% 3.22
M −3.08 × 10−5 −3.06 × 10−5 2.6% 6.83
-0.00019 -0.00124
-0.0002 p=7.6 -0.00125 p=2.0
-0.00021 U=11.2% -0.00126 U=1.0%
-0.00022 -0.00127
-0.00128
Xp
Xf
-0.00023
-0.00129
-0.00024 -0.0013
-0.00025 -0.00131
-0.00026 -0.00132
-0.00027 -0.00133
0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3
relative step size relative step size
-0.00149
U=1.7%
-0.0015
-0.00151
-0.00152
X
-0.00153
-0.00154
-0.00155
-0.00156
0 0.5 1 1.5 2 2.5 3
relative step size
but is shifted down to smaller values. It is possible that this discrepancy is a result of
the sailplanes being present in the experiments but not in the calculations. The drag
and lift on the sailplanes would tend to increase M , consistent with the shift observed
in Figure 5.26. It also appears from the scatter in the data and the differences between
positive and negative drift angles that there is more uncertainty in M (relative to the
scale used for the plot) than in the other integral quantities.
The comparison with experiments showed that CFD gives accurate predictions for
the forces and moments on the submarine for the case of unrestricted water over a range
of drift angles.
90 Chapter 5 – Verification and validation of steady motion calculations
2
CFD, 1x106 Cells, Re = 5.2x106
CFD, 4x106 Cells, Re = 5.2x106 x
CFD, Extrap. to Re=14x106 20
1 x Expt, Re = 9x106 x
6
+ Expt, Re = 14x10
x +
x
+
+
103 X’
103 Y’
++
++
0 x 0 ++
+
+
x +
x
+
x
x
+
-1 + + x
++ ++++ + + x -20
+ + x
+
-2
-20 -10 0 10 20 -20 -10 0 10 20
β (deg) β (deg)
12 1.5
10 1 x
x
8 x
x 0.5 +
x
+
6 +
103 K’
103 Z’
++
++
x
x
0 ++
++
4 +
x
+
+
x -0.5
2 +
x
x
+
+
x
0
++
+++ + -1
+++
++
-2 -1.5
-20 -10 0 10 20 -20 -10 0 10 20
β (deg) β (deg)
1.2 10
1 x x
x 5 x
0.8 +
x
x +
+ +
0.6 +
103 M’
103 N’
x +
x ++
+ 0 +
+
+
0.4 + + +
x + +
+ +
+ +
x
+++++
0.2 ++
+ x
-5
x
0
-0.2 -10
-20 -10 0 10 20 -20 -10 0 10 20
β (deg) β (deg)
Figure 5.26: Comparison between configuration 2 experiments [34] and calculations, Walrus
(deep water, 4 × 106 cells)
5.6 – Conclusion 91
5.6 Conclusion
In this chapter, verification and validation of the predicted forces and moments on several
different ship hulls in manoeuvring motions have been presented using available validation
data from literature. It was demonstrated that for a wide range of ship types, good
predictions of the loads on the hull in manoeuvring motion can be obtained. The trends
in the forces and moments as a function of the drift angle or yaw rate are simulated well.
The results obtained with two different solvers and using different turbulence models were
compared. In general, it is concluded that the differences between the results obtained
with the two solvers are relatively small and within the numerical uncertainties, except
for cases where large differences in grid density were used. In those cases, better results
were obviously obtained on the finer grids. In general, it appears that the combination
of ReFRESCO with SST provides results that are slightly closer to the experimental
values than the combination of Parnassos with MNT. Considering the computational
effort, some differences exist between the two solvers. Parnassos is very fast when a
converged solution can be obtained and needs only two processors for the cases presented
in this section. For ReFRESCO a large number of processors is required to solve the
flow in a similar time-frame as Parnassos. However, for more complicated flows, such
as found at large drift angles or rotation rates, it becomes more difficult to converge a
solution with Parnassos, since this code is optimised for ships sailing in straight-ahead
conditions. The combination of ReFRESCO as a general code and the possibility to use
high-quality multi-block structured grids is more robust for these conditions and requires
much less manual interaction to arrive at a converged solution.
The verification studies provide useful insight into the influence of the grid density
on the predicted forces and moments. Summarizing, the numerical uncertainties in the
forces and moments obtained by the viscous-flow calculations were found to be about 10%
to 15% on average. The uncertainties in Y and K appear to be somewhat higher than in
X or N . In several cases, validation of the calculations failed, indicating modelling errors
in the numerical results. In these cases, it was generally seen that the magnitude of the
transverse force was under-predicted, while the magnitude of the yaw moment was over-
predicted. For manoeuvring studies in the early design, the comparison errors are within
acceptable levels. However, improvements remain desired and might be obtained using
finer grids, larger domain sizes, different grid topologies with refinement in the wake of
the ship, other turbulence models or incorporating free surface deformation. Furthermore,
unsteady phenomena in the flow have been ignored. In future studies, the influence of
instationary flow on the forces and moments needs to be investigated.
For one of the hull forms presented in this chapter, the DARPA SUBOFF, the influence
of a variation of the turbulence model was studied. By changing the turbulence model
from MNT to the more complex SST model, a considerable reduction of the comparison
error in X was found. Unfortunately, the changes in the other forces and moments were
small and therefore this did not lead to the desired overall improvements. Further study
is therefore required to determine the cause of the modelling error.
92 Chapter 5 – Verification and validation of steady motion calculations
Chapter 6
6.1 Introduction
Unfortunately, results of free sailing manoeuvres for the HTC (see 2.5.4) were not available
at the end of the VIRTUE project and validation of the predicted manoeuvres could not be
performed. Therefore, MARIN decided to perform such manoeuvres outside of the scope
of VIRTUE. The results of this manoeuvring test programme can be used for public
domain comparisons of simulation results and for development of procedural guidelines
for free running model tests when it concerns manoeuvring of single propeller ships and
engine control during manoeuvres.
The purpose of the manoeuvring tests was to determine the yaw checking and course
changing abilities and the turning ability of the ship and provide data for validation of
manoeuvring predictions. To determine the manoeuvring characteristics, standard zig-
zag and combined turning circle/pull-out experiments were conducted with the following
variations:
• different procedures for the rudder angle application during zig-zag manoeuvres with
respect to the neutral rudder angle
This chapter presents details of the model, the experimental facility and equipment,
the data reduction procedures and the test programme. In this thesis, only the tests
without bilge keels and with constant propeller RPM are considered. More details about
the tests and drawings of the propeller and bilge keels can be found in MARIN Report
No. 23277-3-SMB [105].
94 Chapter 6 – Free sailing manoeuvring tests
Tests have been performed with a constant RPM setting and with modelling of a
concise engine control. In the latter case, a maximum power level that could be delivered
to the propeller corresponding to 100% Maximum Continuous Rating (MCR) at full scale
was assumed. When the power absorbed by the propeller exceeds this value, the propeller
rate of revolutions is decreased, simulating full scale engine behaviour. This method has
been validated against full scale feedback and has been found to provide better agreement
between model tests and full scale trials. The power level absorbed by the propeller
was measured during a speed run corresponding to 18.0 kn at full scale. This power was
assumed to be equivalent to 85% MCR.
At two adjacent sides of the basin, segmented wave generators consisting of hinged
flaps are installed. Each flap is controlled separately by a driving motor and has a width
of 60 cm. This set-up makes it possible to generate waves in any direction. The waves can
be long and short crested and multi-directional. The wave generator system is equipped
with an active wave reflection compensation feature and higher order wave synthesis
techniques. Opposite the wave generators, passive sinkable wave absorbers are installed.
A Krypton contact-less optical measurement system (now part of Nikon, see e.g.
www.nikonmetrology.com/optical cmm) is used to determine the position of the model
in six DOF. A target consisting of several infra-red Light-Emitting Diodes (LEDs) glued
in fixed positions on a non-deformable plate is mounted on the model on a location such
that the target is in the Krypton observation area and the location and orientation of the
target can be determined. By specifying in Krypton the position of the target relative
to the centre of gravity, the system calculates the position of the centre of gravity of the
model relative to the sub-carriage based on the measured position and orientation of the
target. The Krypton camera is mounted on the subcarriage and consequently moves with
the carriage. To obtain the x and y position of the vessel in the basin, the position of the
sub-carriage and the relative distance measured with Krypton are combined. All motions
are defined in the basin-fixed system of axes, except roll and pitch which are defined in
the ship-fixed system of axes.
The rudder angle and propeller RPM are actively controlled by the steering system.
The data acquisition consists of recording analogue and digital signals. The analogue
signals (e.g. propeller and rudder forces) are sampled before being recorded by the mea-
surement system, while digital signals (e.g. steering system and Krypton output) are
recorded directly.
see ITTC Recommended Procedures and Guidelines on free sailing manoeuvring tests
[73]. Before commencing the manoeuvring tests, the relation between the propeller RPM
and the achieved speed was determined. During these tests, the model was steered on a
straight course with an autopilot. From the recordings, the average rudder angle required
for straight-ahead sailing was obtained. These angles were adopted as the neutral rud-
der angle to compensate the propeller wheel effect for that specific speed. Based on the
RPM-speed relationship, the propeller RPMs to sail at speeds corresponding to 10 kn and
18 kn were derived. These RPMs were used during the remainder of the test programme.
Standard zig-zag and combined turning circle/pull-out manoeuvres have been con-
ducted (see e.g. Bertram [13]). A photograph of the ship model during one of the tests is
shown in Figure 6.3. For the zig-zag tests, the rudder execute angle was given relative to
the neutral rudder angle. For a neutral angle of e.g. 1◦ to starboard, the actual mechanical
steering angles were 11◦ to starboard and 9◦ to port-side for a 10◦ /10◦ zig-zag. In the test
results, the presented rudder angle is however compensated for the neutral angle and will
therefore show rudder angles between 10◦ PS and 10◦ SB.
Figure 6.3: HTC ship model during free sailing manoeuvring test
For this thesis, only a subset of the test programme was used. This subset corresponds
to the conditions for which simulations were performed and is summarised in Table 7.5
on page 112, where also references are given to the table pages and figure pages with
the results. In MARIN Report No. 23277-3-SMB [105] a discussion of all tests is given.
98 Chapter 6 – Free sailing manoeuvring tests
The repeat tests with the ship model equipped with bilge keels were only used to provide
further estimates of the uncertainty in the experimental results.
20◦ /20◦ zig-zag test: at the higher approach speed, the higher turning rate during the
time required to reverse the rudder angle results in larger overshoot angles.
Interesting to see is the difference between the 18 kn results with and without bilge
keels. With bilge keels, the overshoot and initial turning ability values appear to in-
crease (i.e. a deterioration of the yaw checking and initial turning abilities) while also the
uncertainties in the results increase.
Comparing the uncertainty estimates based on the first test series (Table 6.2) and
those based on all tests (Table 6.1), it is observed that due to the larger number of
observations the uncertainty in the mean values of the zig-zag parameters reduced. This
is mainly caused by a reduction in the standard deviations sφ and a reduction of the
Student t coverage factor. From this, it can be concluded that the Student t coverage
factor results in a conservative uncertainty estimate.
6.8 Conclusion
Free sailing manoeuvring test on the HTC have been performed. This test campaign
resulted in a very valuable data set which can be used for public validation studies. Besides
obtaining general characteristics of the manoeuvrability of a single-screw container ship,
unique information has been obtained on the drift angles and rates of turn combined with
propeller and rudder forces. From this, important information for the development and
validation of manoeuvring prediction tools is obtained. Furthermore, repeat tests have
been conducted for selected manoeuvres and based on these tests, the uncertainty in the
characteristic manoeuvring properties has been estimated. Even when a small number
of observations is available, it is concluded that the verification procedure proposed in
100 Chapter 6 – Free sailing manoeuvring tests
Table 6.1: Experimental uncertainty estimate, zig-zag, 95% confidence interval, all tests
Table 6.2: Experimental uncertainty estimate, zig-zag, 95% confidence interval, set 1
Table 6.3: Experimental uncertainty estimate, turning circle, 95% confidence interval, all tests
Table 6.4: Experimental uncertainty estimate, turning circle, 95% confidence interval, set 1
this chapter provides good estimates of the uncertainty in the measurements, provided a
Student t coverage factor is used.
In the following chapter, hydrodynamic coefficients will be derived and manoeuvring
simulations will be conducted for the HTC. The model tests results will be used to validate
these simulations.
103
Chapter 7
7.1 Introduction
When SurSim is used to predict the manoeuvrability of the HTC, results as shown
in Figure 7.1, Table 7.1 and Table 7.2 are obtained1 . It is clearly seen that in order to
reliably assess the manoeuvring behaviour of the ship an improvement of the mathematical
formulae is required: the comparison error for the first overshoot angle during the 20◦ /20◦
zig-zag manoeuvre is −72% of the experimental value and the comparison error in the
Tactical Diameter (TD) is +46%. In earlier studies, it was already seen that the forces
and moments on the bare hull predicted by SurSim were insufficiently accurate, see [148].
To improve the empirical formulations in the mathematical model, CFD calculations will
be used.
600
45
500
30
400
15
−δ [deg] ; ψ [deg]
x [m]
300
E
0
200
−15
100
−30
0
20◦ /20◦ PS 35◦ SB
−45
0 25 50 75 100 125 150 175 200 225 250 −100 0 100 200 300 400 500 600 700
108005 108013 205002 104028 104026
time 108009 zz09.26−20.00 yE [m] 104019 tc09.26−−35.00
Figure 7.1: Comparison between the original simulations and the free sailing experiments, 18 kn
(thick blue lines: simulation, others: experiments)
In this chapter, hydrodynamic coefficients for the modelling of the HTC bare hull
1
In this chapter, modifications to some of the empirical hull-propeller-rudder coefficients will be made.
In the original SurSim predictions here, the same modifications were made. This means that any differ-
ences between these simulations and the simulations presented later in this chapter are only caused by
changes in the mathematical model for the bare hull forces.
104 Chapter 7 – Simulation of ship manoeuvrability
forces and moments as presented in chapter 5 will be derived. Only the results computed
using Parnassos will be considered, to limit the scope of the work. The procedure pro-
posed in section 3.3 is followed. With the obtained coefficients, manoeuvring simulations
will be conducted with SurSim. The results of the simulations will be compared to the
free sailing manoeuvring experiments to demonstrate the feasibility of the approach and
the improvement of the simulations compared to the original fully empiric SurSim sim-
ulations. See section 2.5.4 and chapter 6 for more information on the HTC and the free
sailing manoeuvring experiments.
Table 7.1: Summary of zig-zag manoeuvre results, original simulations, 18 kn (average from
repeat tests)
Table 7.2: Summary of turning circle manoeuvre results, original simulations, 18 kn (average
from repeat tests)
200
1:24 (HSVA trimmed)
180 1:24 (est. even keel)
160 1:24 (EFD captive)
140 1:30.02 (est.)
120
Rm [N]
100
80
60
40
20
0
0 0.5 1 1.5 2 2.5
Vm [m/s]
The wake fraction w and thrust deduction fraction t for the captive condition are
estimated based on DESP predictions. The predicted values are: w = 0.38 and t = 0.22.
in Figure 7.3 are obtained. For all comparisons between the SurSim results and the
captive measurements propeller 2208 is modelled, while for the comparisons with the free
sailing tests propeller 5286 is modelled. The coefficients derived from the open water tests
with propeller No. 5286 are as given in Table 7.3.
0.5
KT (2208)
KQ (2208)
0.4 KT (5286)
KQ (5286)
0.3
KT , 10 · KQ
0.2
0.1
-0.1
0 0.2 0.4 0.6 0.8 1
J
Table 7.3: Propeller No. 5286, open water test No. 45127
Based on the HSVA captive experiments, the relation between the side force on the rudder
and its effect on the total side force on the ship was validated. The value calculated by
SurSim for (1 + aH ) (see Equation 3.37) appears to correctly model the rudder force on
the ship. For the HTC, a value of (1 + aH ) = 1.255 is found. For the ship without rudder,
a relation between the rudder angle and the force on the ship as shown in Figure 7.4 is
obtained. It is seen that this modelling closely approximates the experimental values,
except for the largest rudder angles, where stall appears to be present in the experiments.
7.2 – Deriving the hydrodynamic coefficients 107
However, during the manoeuvres studied in this thesis, the rudder operates in the propeller
race, and it is known that for those conditions, the stall angle increases considerably, see
Kracht [80] or Molland and Turnock [94], and a lift curve as modelled by SurSim is more
appropriate.
1 8
SurSim SurSim
exp 6 exp
0.5 4
N × 104 [kNm]
Y × 103 [kN]
2
0 0
-2
-0.5 -4
-6
-1 -8
-40-30-20-10 0 10 20 30 40 -40 -30 -20 -10 0 10 20 30 40
δ [deg] δ [deg]
Figure 7.4: Forces on the ship as function of rudder angle, HTC without propeller, 18 kn, β = 0◦
Propeller-to-rudder interaction
Based on the HSVA captive experiments, the relation between the propeller thrust and
induced velocity on the rudder was estimated (see Equation 3.28). To correlate the forces
predicted by SurSim to the measured forces on the ship, the following value is used:
Crue = 0.55. A relation between the rudder angle and the force on the ship for different
propeller revolutions n as given in Figure 7.5 is obtained. The revolutions at model self
propulsion are designated ns . It is seen that this modelling closely approximates the
experimental values, except for the largest rudder angles.
108 Chapter 7 – Simulation of ship manoeuvrability
3
SurSim, no prop
SurSim, ns
2 SurSim, 1.4 · ns
exp
Y × 103 [kN]
-1
-2
-3
-40 -30 -20 -10 0 10 20 30 40
δ [deg]
2
SurSim, no prop
1.5 SurSim, ns
SurSim, 1.4 · ns
1 exp
N × 105 [kNm]
0.5
0
-0.5
-1
-1.5
-2
-40 -30 -20 -10 0 10 20 30 40
δ [deg]
Figure 7.5: Forces on the ship as function of rudder angle and propeller revolutions,
18 kn, β = 0◦
7.2 – Deriving the hydrodynamic coefficients 109
Flow straightening
The relation between the drift angle β and rotation rate γ and the effective inflow at the
rudder was validated with the HSVA captive experiments. The last two free parameters
in the SurSim rudder model are the coefficients for the flow straightening for drift (Cdb )
and the flow straightening for rotation (Cdr ), see Equation 3.29. These needed to be
modified to obtain better agreement with the tests. For the HTC, the following values
were adopted: Cdb = 0.9, Cdr = 0.8. Figure 7.6 and Figure 7.7 show the relation between
the rudder angle and the force on the ship for a drift angle of β = −10◦ and for a yaw rate of
γ = 0.4 respectively. With these settings, the rudder forces are modelled reasonably well,
although some discrepancies still remain. Further improvement to the rudder modelling
is however judged to be outside of the scope of the present work, since the main focus is
to demonstrate the influence of the bare hull force model on the manoeuvrability of the
ship.
6 4.5
SurSim SurSim
exp 4 exp
5
3.5
N × 105 [kNm]
Y × 103 [kN]
4 3
2.5
3
2
2 1.5
1
1
0.5
0 0
-40 -30 -20 -10 0 10 20 30 40 -40-30-20-10 0 10 20 30 40
δ [deg] δ [deg]
Figure 7.6: Forces on the ship as function of rudder angle, 18 kn, β = −10◦ , n = ns
5 0
SurSim SurSim
exp -0.5 exp
4
-1
N × 105 [kNm]
Y × 103 [kN]
3 -1.5
-2
2
-2.5
1 -3
-3.5
0
-4
-1 -4.5
-40 -30 -20 -10 0 10 20 30 40 -40-30-20-10 0 10 20 30 40
δ [deg] δ [deg]
Figure 7.7: Forces on the ship as function of rudder angle, 18 kn, γ = 0.4, n = ns
Table 7.4: Hydrodynamic bare hull and added mass coefficients, HTC
0.4 0.2
β = −10◦
0.3 β = 0◦ 0.15
β = 6◦
0.2 β = 10◦ 0.1
β = 16◦
0.1 0.05
N [-]
Y [-]
0 0
-0.1 -0.05
β = −10◦
-0.2 -0.1 β = 0◦
β = 6◦
-0.3 -0.15 β = 10◦
β = 16◦
-0.4 -0.2
-0.8 -0.6 -0.4 -0.2 0 0.2 0.4 -0.8-0.6-0.4-0.2 0 0.2 0.4
γ γ
0.4 0.2
γ = −0.6
0.3 γ = −0.4 0.15
γ = −0.2
0.2 γ=0 0.1
γ = 0.2
0.1 0.05
N [-]
Y [-]
0 0
-0.1 -0.05
γ = −0.6
-0.2 -0.1 γ = −0.4
γ = −0.2
-0.3 -0.15 γ =0
γ = 0.2
-0.4 -0.2
-30 -20 -10 0 10 20 30 -30 -20 -10 0 10 20 30
β [deg] β [deg]
Figure 7.8: Forces on the bare hull as function of yaw rate (top) or drift angle (bottom)
(lines: fit, markers: cfd results)
112 Chapter 7 – Simulation of ship manoeuvrability
Manoeuvre Presentation
Id V0 δ ψ Table page Figure page
[kn] [deg] [deg] Timetrace Track
zz05.14-10.00 10 -10 161 177 178
zz05.14–10.00 10 -10 10 162 179 180
zz05.14-20.00 20 -20 163 181 182
zz05.14–20.00 -20 20 164 183 184
zz09.26-10.00 10 -10 165 185 186
zz09.26–10.00 18 -10 10 166 187 188
zz09.26-20.00 20 -20 167 189 190
zz09.26–20.00 -20 20 168 191 192
tc05.14-35.00 35 169 193 194
tc05.14–35.00 -35 195 196
tc05.14-25.00 10 25 - 170 197 198
tc05.14–25.00 -25 199 200
tc05.14-15.00 15 171 201 202
tc05.14–15.00 -15 203 204
tc09.26-35.00 35 172 205 206
tc09.26–35.00 -35 173 207 208
tc09.26-25.00 18 25 - 174 209 210
tc09.26–25.00 -25 211 212
tc09.26-15.00 15 175 213 214
tc09.26–15.00 -15 215 216
All calculations have been conducted without incorporating heel. The model tests
have been performed with a GM value that is relatively high for this type of ship and
therefore the influence of heel on the manoeuvres is expected to be small. It should be
noted that for speeds close to the design speed of the ship, assessment of the heel angle
for this type of ship will be important when the GM value is small.
7.4 – Sensitivity study 113
1000
CFD 1:30.02
Exp. (bilgekeels)
800 Exp. (no bilgekeels)
RPM (model)
600
400
200
0
0 0.5 1 1.5 2
Vm [m/s]
Figure 7.9: RPM-Speed relation for HTC, scale 1:30.02, model scale values
HTC HTC
40 40
X’u’|u’|×1.1
30 30 N’β×1.1
X’βγ×1.1
N’γ×1.1
20 Y’β×1.1 20
N’γ|γ|×1.1
10 Y’γ×1.1 10
Change %
Change % N’β|β|×1.1
0 Y’β|β|×1.1 0
N’ab×1.1
Y’γ|γ|×1.1
−10 −10
N’|u’|γc×1.1
Y’ab×1.1
−20 −20 N’ββγ×1.1
Y’β|γ|×1.1
−30 −30 N’βγγ×1.1
Y’|β|γ×1.1
−40 −40
osa1 osa2 ita osa1 osa1 osa2 ita osa1
10/10 10/10 10/10 20/20 10/10 10/10 10/10 20/20
HTC
40
30
20 Crue×1.1
10 Chru×1.1
Change %
0 Cdbluff×1.1
−10 Cdblee×1.1
Cdr×1.1
−20
−30
−40
osa1 osa2 ita osa1
10/10 10/10 10/10 20/20
Based on the sensitivity study, the results as collected in Figure 7.10, Figure 7.11 and
Table 7.6 were obtained for an approach speed of 10 kn. Another sensitivity study that
was conducted for an approach speed of 18 kn shows similar results. It is clear that for
the HTC deviations in Nβ0 have the largest impact on the accuracy of the prediction of
the yaw checking and course keeping ability, while of all linear coefficients it also has the
largest influence on the turning ability. Nγ0 is also an important coefficient. Yγ0 is the least
important linear coefficient for accurate predictions. Furthermore, it is seen that for the
zig-zag manoeuvres, the linear derivatives are more important compared to the non-linear
derivatives than during the turning circle manoeuvres. It is also found that the 10◦ /10◦
zig-zag manoeuvre is more sensitive to changes in the hydrodynamic derivatives than the
20◦ /20◦ zig-zag manoeuvre.
0
The turning ability is most sensitive to changes in the non-linear coefficients Nββγ
7.4 – Sensitivity study 115
HTC HTC
40 40
X’u’|u’|×1.1 X’u’|u’|×1.1
30
30
X’βγ×1.1 X’βγ×1.1
20 Y’β×1.1
20 Y’β×1.1
Y’γ×1.1 10 Y’γ×1.1
Change %
10
Change %
Y’β|β|×1.1 0 Y’β|β|×1.1
0
Y’γ|γ|×1.1 Y’γ|γ|×1.1
−10 −10
Y’ab×1.1 Y’ab×1.1
−20 −20
Y’β|γ|×1.1 Y’β|γ|×1.1
−30 −30
Y’|β|γ×1.1 Y’|β|γ×1.1
−40 −40
AD TD
rstc Vstc βstc
HTC HTC
40 40
N’β×1.1 30 N’β×1.1
30
N’γ×1.1 N’γ×1.1
20 20
N’γ|γ|×1.1 N’γ|γ|×1.1
10
Change %
10
Change %
N’β|β|×1.1 N’β|β|×1.1
0 0
N’ab×1.1 N’ab×1.1
−10
−10
N’|u’|γc×1.1 N’|u’|γc×1.1
−20 N’ββγ×1.1
−20 N’ββγ×1.1
−40 −40
AD TD
rstc Vstc βstc
HTC HTC
40 40
30 30
20 20 Crue×1.1
Crue×1.1
Chru×1.1 10 Chru×1.1
Change %
10
Change %
0 Cdbluff×1.1 0 Cdbluff×1.1
−30 −30
−40 −40
AD TD
rstc Vstc βstc
0
and Nβγγ . The sensitivity of the steady turning circle results to changes in the non-
0
linear coefficient Nβγγ is large: when this coefficient is increased by 10%, the rate of
turn increases and the speed drops such that an unrealistic situation is reached and the
simulation is aborted (indicated by a change of 100%).
Similar conclusions were found by Lee and Shin [84] who studied zig-zag manoeuvres
for a chemical carrier and two oil tankers and Bulian et al. [18] who conducted a sensitivity
study for the Esso Osaka.
The hull-propeller-rudder interaction coefficients Crue , Chru (= 1 + aH ), Cdbluf f (= Cdb )
and Cdr also have a relatively large influence on the yaw checking and course keeping
ability of the ship. From these coefficients, the results are most sensitive to changes in
Chru .
116 Chapter 7 – Simulation of ship manoeuvrability
The sensitivity study demonstrates that for accurate predictions of the manoeuvra-
bility using coefficients derived from CFD calculations, accurate predictions of especially
the yawing moment must be made. It should be noted however, that the sensitivity of the
results depends on the individual ship, due to different balancing between coefficients.
Changes in coefficients are not necessarily independent of changes in other coefficients.
For example, in the procedure used to derive the coefficients, a change in the linear
0
coefficient Yγ0 will lead to changes in the non-linear coefficients Yβ|γ| 0
and Y|β|γ , since these
0
are derived after subtracting the linear contribution Yγ ·cos β·γ and non-linear contribution
0
Yγ|γ| · γ · |γ| from the total force Y 0 .
Table 7.6: Sensitivity study, HTC, 10 kn, changes in percentages of the original values
(Blue values indicate UMFs larger than 50%, while red values indicate UMFs larger
than 100%)
Table 7.7: Summary of zig-zag manoeuvre results (average from repeat tests)
7.5 Validation
A comparison between the simulations based on the improved hydrodynamic derivatives
and the free sailing experiments is made in Table 7.7, Figure 7.12 and Figure 7.13 for the
zig-zag manoeuvres and in Table 7.8 and Figure 7.14 for the turning circles.
Compared to the original simulations with the original bare hull mathematical model
of SurSim, see Table 7.1, a considerable improvement is obtained: all comparison errors
reduce in magnitude and now the predictions of the overshoot angles are conservative
instead of too optimistic in case of the original SurSim predictions.
45 45
◦ ◦
30
10 /10 PS 30
10◦ /10◦ SB
15 15
−δ [deg] ; ψ [deg]
−δ [deg] ; ψ [deg]
0 0
−15 −15
−30 −30
−45 −45
0 50 100 150 200 250 300 350 400 0 50 100 150 200 250 300 350 400
203013 203008 203004 106011 106002 203011 203007 203003 106010 106005
203010 203006 time 106007 zz05.14−10.00 203009 203005 time 106006 zz05.14−−10.00
45 45
30 30
20◦ /20◦ PS
15 15
−δ [deg] ; ψ [deg]
−δ [deg] ; ψ [deg]
0 0
−15 −15
−30 −30
20◦ /20◦ SB
−45 −45
0 50 100 150 200 250 300 350 400 0 50 100 150 200 250 300 350 400
204003 106013 106004 204002 106012 106003
time 106009 zz05.14−20.00 time 106008 zz05.14−−20.00
Figure 7.12: Comparison between the simulations and the free sailing zig-zag experiments, 10 kn
(thick blue lines: simulation, others: experiments)
45 45
30 30
20◦ /20◦ PS
15 15
−δ [deg] ; ψ [deg]
−δ [deg] ; ψ [deg]
0 0
−15 −15
−30 −30
20◦ /20◦ SB
−45 −45
0 25 50 75 100 125 150 175 200 225 250 0 25 50 75 100 125 150 175 200 225 250
108005 108013 108012 108003
time 108009 zz09.26−20.00 time 108008 zz09.26−−20.00
Figure 7.13: Comparison between the simulations and the free sailing 20◦ /20◦ zig-zag experi-
ments, 18 kn (thick blue lines: simulation, others: experiments)
7.6 – Conclusion 119
7.6 Conclusion
Using hydrodynamic manoeuvring coefficients derived from CFD calculations of the forces
on the bare hull, it has been shown that it is possible to improve the prediction of ship ma-
noeuvres compared to predictions using coefficients based on empirical equations, which
was the objective of the present study. In this chapter, it has been demonstrated that
a considerable improvement of the turning circle predictions was obtained. The predic-
tion of the yaw checking and course keeping and initial turning abilities based on zig-zag
simulations improved as well, but further improvements are required for more reliable
assessment of the manoeuvring performance.
The sensitivity of the manoeuvring predictions on changes in the hydrodynamic co-
efficients was studied. It was found that the linear coefficients mostly determine the
sensitivity of the results of the zig-zag manoeuvres, while non-linear coefficients affect
mostly the turning circle results. Hull-propeller-rudder coefficients were also found to
be important in the sensitivity study. The study demonstrates that for accurate predic-
tions of the manoeuvrability using coefficients derived from CFD calculations, accurate
predictions of especially the yawing moment must be made.
120 Chapter 7 – Simulation of ship manoeuvrability
Table 7.8: Summary of turning circle manoeuvre results (average from repeat tests)
450
400
400
350
350
300
300
250
250
xE [m]
x [m]
200 200
E
150 150
100 100
50 50
0 0
◦
−50 35 PS −50 35◦ SB
−500 −400 −300 −200 −100 0 100 −100 0 100 200 300 400 500
205003 104021 104029 205002 104028 104026
y [m] 104027 tc09.26−35.00 y [m] 104019 tc09.26−−35.00
E E
500
450
400
400
350
300
300
250
xE [m]
x [m]
200
E
200
150
100 100
50
0 0
−50 25◦ PS 25◦ SB
−600 −500 −400 −300 −200 −100 0 100 −100 0 100 200 300 400 500 600
104023 104022
y [m] tc09.26−25.00 y [m] tc09.26−−25.00
E E
600
600
500
500
400
400
300
xE [m]
x [m]
300
E
200
200
100
100
0 0
15◦ PS 15◦ SB
−700 −600 −500 −400 −300 −200 −100 0 100 −200 −100 0 100 200 300 400 500 600 700 800
104025 104024
yE [m] tc09.26−15.00 yE [m] tc09.26−−15.00
Figure 7.14: Comparison between the simulations and the free sailing turning circle experiments,
18 kn (thick blue lines: simulation, others: experiments)
122 Chapter 7 – Simulation of ship manoeuvrability
Chapter 8
8.1 Conclusions
The present work was initiated in order to improve traditional manoeuvring simulations
based on empirical mathematical equations to model the forces and moments on the ship.
With the evolution of viscous-flow solvers and their promising results in predicting the
forces and moments on ships, it was decided to develop a practical method to simulate the
manoeuvrability of ships in which viscous-flow solvers are utilised and investigate whether
this improves the accuracy of the predicted simulations.
Several different methods are given in literature to simulate ship manoeuvres using
viscous-flow calculations, see chapter 2. In this thesis, the virtual captive test approach is
adopted, because of the efficient use of computational resources compared to other meth-
ods. Furthermore, this approach can directly be used to improve mathematical models
for manoeuvring simulators. The present study extends the work of other researchers by
providing extensive verification and validation of the predicted forces and moments on
the hull and a detailed study of the sensitivity of the manoeuvring characteristics of the
ship to changes in the hydrodynamic coefficients in the simulation model.
Concerning the forces and moments acting on a ship hull in manoeuvring motions, it is
noteworthy that validation data for captive steady drift motions are much more abundant
than for steady rotation or combined motion. Especially for more extreme conditions with
large turning rates and drift angles, such as occur during tight turning circles, not much
validation data can be found in literature. Furthermore, cases in which both extensive
captive test data and free sailing manoeuvring test data are available are scarce.
In the work leading to this thesis, much effort was spent on simulating the flow around
the HTC for captive conditions and simulations of standard free sailing manoeuvres were
conducted within the VIRTUE project. To generate validation data for these manoeu-
vres, MARIN decided to perform free sailing manoeuvring tests for the HTC. This test
campaign resulted in a very valuable data set which can be used for public validation
studies, see chapter 6. Besides obtaining general characteristics of the manoeuvrability of
a single-screw container ship, unique information has been obtained on the drift angles
124 Chapter 8 – Conclusions and recommendations
and rates of turn in combination with propeller and rudder forces. From this, impor-
tant information for the development and validation of manoeuvring prediction tools is
obtained. Furthermore, repeat tests have been conducted for selected manoeuvres and
based on these tests, the uncertainty in the characteristic manoeuvring properties has
been estimated. Even when a small number of observations is available, it is concluded
that the verification procedure proposed in this thesis provides good estimates of the
uncertainty in the measurements, provided a Student t coverage factor is used.
The manoeuvring prediction program SurSim has been used to simulate the manoeu-
vrability of the HTC, see chapter 3. In the program, it is possible to provide user-defined
hydrodynamic coefficients for the bare hull forces and moments. This makes the program
well suited for the present work to investigate whether the use of viscous-flow calcula-
tions can help in improving the prediction accuracy of simulations. A procedure to derive
the coefficients is proposed. This procedure is chosen to enable accurate modelling of
the linearised behaviour for course-keeping as well as realistic modelling of the harbour
manoeuvring characteristics, and to enable accurate modelling of non-linear manoeuvres.
For ship manoeuvres, not only the flow around the ship in oblique motion is of interest,
but also the flow around the ship when it performs a rotational motion. To compute the
flow around the ship for such conditions, the flow solvers used in the present study had
to be modified. For this work, the rotational motion was incorporated by using a non-
inertial reference system and supplementing the equations of motions with body forces
representing the centrifugal and Coriolis contributions on the flow.
To generate the grids for a range of drift angles and yaw rates, different approaches
were adopted depending on the flow solver used. For Parnassos, automated scripts
were developed with which the desired grids could be generated rapidly. For the more
generic solver ReFRESCO, all calculations are conducted using one grid, but changing
the inflow angles and boundary conditions depending on the desired manoeuvring motion.
To facilitate this, a new boundary condition was developed, which removes the need to
pre-process each grid for each new manoeuvring condition.
In chapter 5, it was demonstrated that for a wide range of ship types, good predictions
of the loads on the hull in manoeuvring motion can be obtained. The trends in the forces
and moments as a function of the drift angle or yaw rate are simulated well. Two different
solvers and using different turbulence models were applied. In general, it is concluded
that the differences between the results obtained with the two solvers are relatively small
and within the numerical uncertainties, except for cases where large differences in grid
density were used. In those cases, better results were obviously obtained on the finer
grids.
The verification studies provide useful insight into the influence of the grid density on
the predicted forces and moments. In several cases, validation of the calculations failed,
indicating modelling errors in the numerical results. In these cases, it was generally seen
that the magnitude of the transverse force was under-predicted, while the magnitude of
the yaw moment was over-predicted. For manoeuvring studies in the early design, the
8.2 – Recommendations 125
comparison errors are within acceptable levels. However, improvements remain desired
and may be obtained using finer grids, larger domain sizes, different grid topologies with
refinement in the wake of the ship, other turbulence models or incorporating free surface
deformation.
The verification and validation studies show that the drift-sweep procedure proposed
in this thesis and implemented in ReFRESCO provides good results in comparison
with the Parnassos calculations and with the experiments. By using the procedure,
the amount of manual interaction for the user decreases considerably. Furthermore, the
convergence for the different drift angles is faster than when each drift angle would be
calculated separately.
8.2 Recommendations
To improve the accuracy of manoeuvring simulations and to reduce the uncertainty in the
simulation results due to small changes in the predicted forces and moments on the hull,
the following recommendations are made:
• One of the most promising prospects of the use of viscous-flow solvers is the ability
to estimate the hull forces for full scale conditions. This will eliminate possible
scale effects and improve the correspondence between the predictions and the actual
prototype results. Therefore, calculations for prototype Reynolds numbers should
be made.
• The overall accuracy of force and moments predictions on ship hulls should be im-
proved by investigating in more detail the influence on the predictions of turbulence
models, domain size, grid density and topology.
126 Chapter 8 – Conclusions and recommendations
• In this thesis, only the bare hull is considered in the viscous-flow calculations. How-
ever, it is possible to include the propeller influence by e.g. using an actuator disc
model to model the propeller thrust, or by calculating the flow around the propeller
with a potential flow code and to introduce the calculated forces on the propeller
as a force field in the RANS calculations. Furthermore, it is possible to include the
rudder in the grid, such that the forces due to the rudder can be obtained as well.
With such computations, the empirical modules for the propeller and rudder can be
substituted by results obtained with the viscous-flow solvers and probably improve
the manoeuvring predictions.
• To prepare for future developments (i.e. increase in computing power), the coupling
of the RANS equations and the equations of motions should be implemented in
ReFRESCO. This will avoid the simplifications made in the virtual captive test
approach, such as quasi-steadiness, and therefore result in a better reliability of
the simulations. In a first stage, this can be done for the bare hull only, with the
appendage forces predicted using e.g. a coupling with SurSim, but fully appended
in a later stage.
127
References
[1] I. H. Abbott and A. E. von Doenhoff. Theory of Wing Sections: Including a Sum-
mary of Airfoil Data. Dover Publications, Inc., New York, 1959.
[3] B. Alessandrini and G. Delhommeau. Viscous free surface flow past a ship in drift
and in rotating motion. In 22nd Symposium on Naval Hydrodynamics, pages 491–
507, Fukuoka, Japan, August 1998.
[5] N. Alin, C. Fureby, and S. U. Svennberg. LES of the flow past simplified submarine
hulls. In 8th International Conference on Numerical Ship Hydrodynamics, Busan,
Korea, September 2003.
[7] R. Aris. Vectors, Tensors and the Basic Equations of Fluid Mechanics. Prentice-
Hall, Inc., Englewood Cliffs, N.J., 1962.
[9] ASME. Standard for verification and validation in computational fluid dynamics
and heat transfer. American Society of Mechanical Engineers (ASME) V&V 20
Committee, November 2009.
128 References
[10] P. Atsavapranee, T. Forlini, D. Furey, J. Hamilton, S. Percival, and C.-H. Sung. Ex-
perimental measurements for CFD validation of the flow about a submarine model
(ONR Body-1). In 25th Symposium on Naval Hydrodynamics, St. Johns, Newfound-
land and Labrador, Canada, August 2004.
[14] M. C. Bettle, A. G. Gerber, and G. D. Watt. Unsteady analysis of the six DOF
motion of a buoyantly rising submarine. Computers & Fluids, 38(9):1833–1849,
October 2009.
[16] S. Bhushan, T. Xing, P. Carrica, and F. Stern. Model- and full-scale URANS simu-
lations of Athena resistance, powering, seakeeping, and 5415 maneuvering. Journal
of Ship Research, 53(4):179–198, December 2009.
[19] P. W. Bull. The validation of CFD predictions of nominal wake for the SUBOFF
fully appended geometry. In 21st Symposium on Naval Hydrodynamics, pages 1061–
1076, Trondheim, Norway, June 1996.
[20] P. W. Bull and S. Watson. The scaling of high Reynolds number viscous flow
predictions for appended submarine geometries. In 22nd Symposium on Naval Hy-
drodynamics, pages 1000–1014, Fukuoka, Japan, August 1998.
[21] E. F. Campana, A. Di Mascio, and R. Penna. CFD analysis of the flow past a ship
in steady drift motion. In 3rd Osaka Colloquium on advanced CFD applications to
ship flow and hull form design, pages 151–159, Osaka, Japan, May 1998.
References 129
[22] P. M. Carrica, F. Ismail, M. Hyman, S. Bhushan, and F. Stern. Turn and zigzag
maneuvers of a surface combatant using a URANS approach with dynamic overset
grids. In SIMMAN Workshop on Verification and Validation of Ship Manoeuvring
Simulation Methods, pages F17–22, Copenhagen, Denmark, April 2008.
[23] P. M. Carrica and F. Stern. DES simulations of KVLCC1 in turn and zigzag maneu-
vers with moving propeller and rudder. In SIMMAN Workshop on Verification and
Validation of Ship Manoeuvring Simulation Methods, pages F11–16, Copenhagen,
Denmark, April 2008.
[25] D. Clarke and A. P. Mesbahi, editors. 24th International Towing Tank Conference,
Edinburgh, United Kingdom, September 2005.
[26] C. L. Crane. Maneuvering trials of the 278,000 DWT Esso Osaka in shallow and
deep water. In Transactions of the SNAME, volume 87, pages 251–283, 1979.
[27] A. Cura Hochbaum. Computation of the turbulent flow around a ship model in
steady turn and in steady oblique motion. In 22nd Symposium on Naval Hydrody-
namics, pages 550–567, Fukuoka, Japan, August 1998.
[28] A. Cura Hochbaum. Virtual PMM tests for manoeuvring prediction. In 26th Sym-
posium on Naval Hydrodynamics, Rome, Italy, September 17-22 2006.
[30] A. Cura Hochbaum, M. Vogt, and S. Gatchell. Manoeuvring prediction for two
tankers based on RANS simulations. In SIMMAN Workshop on Verification and
Validation of Ship Manoeuvring Simulation Methods, pages F23–28, Copenhagen,
Denmark, April 2008.
[36] A. Di Mascio and E. F. Campana. The numerical simulation of the yaw flow of a free
surface ship. In 7th International Conference on Numerical Ship Hydrodynamics,
pages 6.1–1–10, Nantes, France, July 1999.
[38] L. Eça and M. Hoekstra, editors. Workshop on CFD Uncertainty Analysis, Lisbon,
Portugal, October 2004. Instituto Superior Técnico.
[39] L. Eça and M. Hoekstra. On the influence of grid topology on the accuracy of ship
viscous flow calculations. In 5th Osaka Colloquium on Advanced CFD Applications
to Ship Flow and Hull Form Design, pages 1–10, March 2005.
[40] L. Eça and M. Hoekstra. The numerical friction line. Journal of Marine Science
and Technology, 13(4):328–345, November 2008.
[41] L. Eça and M. Hoekstra. Evaluation of numerical error estimation based on grid
refinement studies with the method of the manufactured solutions. Computers &
Fluids, 38(8):1580–1591, September 2009.
[42] L. Eça and M. Hoekstra. On the numerical accuracy of the prediction of resis-
tance coefficients in ship stern flow calculations. Journal of Marine Science and
Technology, 14(1):2–18, March 2009.
[43] L. Eça, M. Hoekstra, and J. Windt. Practical grid generation tools with applications
to ship hydrodynamics. In 7th International Conference on Grid Generation and
Computational Field Simulations, Hawaii, February 2002.
[48] K. Eloot and M. Vantorre. Development of a tabular manoeuvring model for hull
forces applied to full and slender ships in shallow water. In MARSIM Interna-
tional Conference on Marine Simulation and Ship Manoeuvring, pages RC–18–1–9,
Kanazawa, Japan, August 2003. The Society of Naval Architects of Japan and Japan
Institute of Navigation.
[51] F. Fathi, C. M. Klaij, and A. H. Koop. Predicting loads on a LNG carrier with
CFD. In 29th International Conference on Ocean, Offshore and Arctic Engineering
OMAE, number OMAE2010-20122, Shanghai, China, June 2010.
[52] T. I. Fossen. Guidance and Control of Ocean Vehicles. John Wiley and Sons, 1994.
[53] C. Fureby. Towards the use of large eddy simulation in engineering. In 46th AIAA
Aerospace Sciences Meeting and Exhibit, number AIAA 2008-605, Reno, Nevada,
January 2008. American Institute of Aeronautics and Astronautics.
[55] K. Granlund. Steady and Unsteady Maneuvering Forces and Moments on Slender
Bodies. PhD thesis, Virginia Polytechnic Institute and State University, February
2009.
[58] T. Hino, editor. CFD Workshop Tokyo, Tokyo, Japan, March 2005. National Mar-
itime Research Institute.
[60] M. Hoekstra and L. Eça. PARNASSOS: an efficient method for ship stern flow
calculation. In 3rd Osaka Colloquium on advanced CFD applications to ship flow
and hull form design, pages 331–357, Osaka, May 1998.
[61] M. Hoekstra, L. Eça, J. Windt, and H. C. Raven. Viscous flow calculations for
KVLCC2 and KCS models using the PARNASSOS code. Gothenburg 2000 Work-
shop on Numerical Ship Hydrodynamics, 2000.
[63] J. P. Hooft. The cross flow drag on a manoeuvring ship. Ocean Engineering,
21(3):329–342, April 1994.
[64] J. P. Hooft and U. Nienhuis. The prediction of the ship’s manoeuvrability in the
design stage. SNAME Annual Meeting, 102:1–24, November 1994.
[68] T. T. Huang, H.-L. Liu, N. Groves, T. Forlini, J. Blanton, and S. Gowing. Mea-
surements of flows over an axisymmetric body with various appendages in a wind
tunnel: the DARPA SUBOFF experimental program. In 19th Symposium on Naval
Hydrodynamics, pages 312–346, Seoul, South Korea, August 1992.
[69] IMO. Resolution MSC. 137(76): Standards for ship manoeuvrability. International
Maritime Organisation, London, 2002.
References 133
[71] T. Ishiguro, S. Tanaka, and Y. Yoshimura. A study on the accuracy of the recent
prediction technique of ship’s manoeuvrability at early design stage. In MARSIM
International Conference on Marine Simulation and Ship Manoeuvring, pages 547–
561, Copenhagen, Denmark, September 1996.
[73] ITTC. Testing and extrapolation methods manoeuvrability - free running model
tests. Recommended Procedures and Guidelines, 7.5-02, 06-01, September 2008.
[74] G. Jensen, M. Klemt, and Y. Xing-Kaeding. On the way to the numerical basin for
seakeeping and manoeuvring. In 9th Symposium on Practical Design of Ships and
Other Floating Structures (PRADS), number 74, Lübeck-Travemünde, Germany,
September 2004.
[78] J. C. Kok. Resolving the dependence on free-stream values for the k − ω turbulence
model. National Aerospace Laboratory, The Netherlands, Report NLR-TP-99295,
July 1999.
[79] A. H. Koop, C. M. Klaij, and G. N. V. B. Vaz. Predicting wind shielding for FPSO
tandem offloading using CFD. In 29th International Conference on Ocean, Offshore
and Arctic Engineering OMAE, number OMAE2010-20284, Shanghai, China, June
2010.
[81] G. Kuiper. Resistance and propulsion of ships. Delft University of Technology, July
1991.
134 References
[84] H.-Y. Lee and S.-S. Shin. The prediction of ship’s manoeuvring performance in
initial design stage. In Practical Design of Ships and Mobile Units (PRADS), pages
633–639, The Hague, The Netherlands, September 1998.
[85] S.-J. Lee, H.-R. Kim, W.-J. Kim, and S.-H. Van. Wind tunnel tests on flow charac-
teristics of the KRISO 3,600 TEU containership and 300k VLCC double-deck ship
models. Journal of Ship Research, 47(1):24–38, March 2003.
[86] S. K. Lee and M. Fujino. Assessment of a mathematical model for the manoeuvring
motion of a twin-propeller twin-rudder ship. International Shipbuilding Progress,
50(1/2):109–123, 2003.
[87] S. W. Lee. Free sailing manoeuvring tests on KVLCC1 and KVLCC2. MARIN
Report No. 21571-1-SMB, Wageningen, The Netherlands, December 2007.
[88] H.-L. Liu and T. T. Huang. Summary of DARPA SUBOFF experimental program
data. Report No. CRDKNSWC/HD-1298-11, June 1998.
[89] J. F. Longo. Effects of Yaw on Model-Scale Ship Flows. PhD thesis, University of
Iowa, May 1996.
[90] S. C. McParlin and R. W. Tramel. Taxonomy of flight mechanics issues for aircraft,
and underlying fluid dynamics phenomena. In 47th AIAA Aerospace Sciences Meet-
ing Including The New Horizons Forum and Aerospace Exposition, number AIAA
2009-744, Orlando, Florida, January 2009. American Institute of Aeronautics and
Astronautics.
[92] F. R. Menter. Eddy viscosity transport equations and their relation to the k −
model. Journal of Fluids Engineering, 119(4):876–884, December 1997.
[93] R. W. Miller. PMM calculations for the bare and appended DTMB 5415 using the
RANS solver CFDShip-Iowa. In SIMMAN Workshop on Verification and Validation
of Ship Manoeuvring Simulation Methods, pages F41–49, Copenhagen, Denmark,
April 2008.
References 135
[96] NATO. ANEP-70 Guidance for naval surface ships mission oriented manoeuvring re-
quirements. Naval Group 6 Specialist Team on Naval Ship Manoeuvrability, Septem-
ber 2003.
[97] N. H. Norrbin. Theory and observations on the use of a mathematical model for
ship maneuvering in deep and confined waters. In 8th Symposium on Naval Hydro-
dynamics, pages 807–904, Pasadena, California, August 1970.
[100] T. Ohmori, M. Fujino, and H. Miyata. A study on flow field around full ship forms
in manoeuvring motion. Journal of Marine Science and Technology, 3(1):22–29,
March 1998.
[101] P. Oltmann and S. D. Sharma. Simulation of combined engine and rudder ma-
noeuvres using an improved model of hull-propeller-rudder interactions. In 15th
Symposium on Naval Hydrodynamics, pages 83–108, Hamburg, Germany, Septem-
ber 1984.
[102] M. W. C. Oosterveld and P. van Oossanen. Further computer analyzed data of the
Wageningen B-screw series. International Shipbuilding Progress, 22(251):251–262,
July 1975.
[103] M. Örnfelt. Naval mission and task driven manoeuvrability requirements for naval
ships. In 10th International Conference on Fast Sea Transportation (FAST), pages
505–518, Athens, Greece, October 2009.
[104] B. Overpelt. Manoeuvring tests on KCS. MARIN Report No. 23991-1-SMB, Wa-
geningen, The Netherlands, September 2010.
[105] B. Overpelt and S. L. Toxopeus. Manoeuvring tests on HTC. MARIN Report No.
23277-3-SMB, Wageningen, The Netherlands, January 2011.
136 References
[111] F. H. H. A. Quadvlieg and P. van Coevorden. Manoeuvring criteria: more than IMO
A751 requirements alone! In MARSIM International Conference on Marine Sim-
ulation and Ship Manoeuvring, pages RB–1–1–8, Kanazawa, Japan, August 2003.
The Society of Naval Architects of Japan and Japan Institute of Navigation.
[113] H. C. Raven, A. van der Ploeg, and L. Eça. Extending the benefit of CFD tools in
ship design and performance prediction. In 7th ICHD International Conference on
Hydrodynamics, pages 573–580, October 2006.
[114] H. C. Raven, A. van der Ploeg, and A. R. Starke. Computation of free-surface vis-
cous flows at model and full scale by a steady iterative approach. In 25th Symposium
on Naval Hydrodynamics, August 2004.
[118] N. Sakamoto. URANS and DES Simulations Of Static And Dynamic Maneuvering
For Surface Combatant. PhD thesis, University of Iowa, Iowa City, Iowa, May 2009.
[119] N. Sakamoto, P. M. Carrica, and F. Stern. URANS simulations of static and dy-
namic maneuvering for surface combatant. In SIMMAN Workshop on Verification
and Validation of Ship Manoeuvring Simulation Methods, pages F50–55, Copen-
hagen, Denmark, April 2008.
[120] M. D. Salas. Digital flight: The last CFD aeronautical grand challenge. Journal of
Scientific Computing, 28(213):479–505, September 2006.
[122] C. D. Simonsen. Rudder, Propeller and Hull Interaction by RANS. PhD thesis,
Technical University of Denmark, Department of Naval Architecture and Offshore
Engineering, May 2000.
[124] C. D. Simonsen and F. Stern. Flow pattern around an appended tanker hull form
in simple maneuvering conditions. Computers & Fluids, 34(2):169–198, February
2005.
[125] C. D. Simonsen and F. Stern. RANS maneuvering simulation of esso osaka with
rudder and a body-force propeller. Journal of Ship Research, 49(2):98–120, June
2005.
[126] C. D. Simonsen and F. Stern. Flow structure around maneuvering tanker in deep
and shallow water. In 26th Symposium on Naval Hydrodynamics, number M1, Rome,
Italy, September 17–22 2006.
[127] C. D. Simonsen, F. Stern, and K. Agdrup. CFD with PMM test validation for
manoeuvring VLCC2 tanker in deep and shallow water. In MARSIM Interna-
tional Conference on Marine Simulation and Ship Manoeuvring, number M04, Ter-
schelling, The Netherlands, June 2006.
[130] F. Stern and K. Agdrup, editors. SIMMAN Workshop on Verification and Validation
of Ship Manoeuvring Simulation Methods, Copenhagen, Denmark, April 2008.
[132] F. Stern, F. Ismail, T. Xing, and P. Carrica. Vortical and turbulent structures
using various convection schemes with algebraic reynolds stress-DES model for the
KVLCC2 at large drift angles. In 27th Symposium on Naval Hydrodynamics, volume
2, B34, Seoul, Korea, October 5–10 2008.
[135] C.-H. Sung, T.-C. Fu, and M. Griffin. Validation of incompressible flow computation
of forces and moments on axisymmetric bodies undergoing constant radius turning.
In 21st Symposium on Naval Hydrodynamics, pages 1048–1060, Trondheim, Norway,
June 1996.
[137] C.-H. Sung, M. J. Griffin, J. F. Tsai, and T. T. Huang. Incompressible flow computa-
tion of forces and moments on bodies of revolution at incidence. In 31st Aerospace
Sciences Meeting and Exhibit, number AIAA 1993-787, Reno, Nevada, January
1993.
[138] C.-H. Sung, M.-Y. Jiang, B. Rhee, S. Percival, P. Atsavapranee, and I.-Y. Koh.
Validation of the flow around a turning submarine. In 24th Symposium on Naval
Hydrodynamics, Fukuoka, Japan, July 2002.
[139] C.-H. Sung, B. Rhee, and E. Ammeen. SHWG CFD benchmark study on DARPA
SUBOFF. Part 1. Bare hull. SHWG Correspondence, pages 1–26, November 2005.
References 139
[140] Y. Tahara, J. F. Longo, and F. Stern. Comparison of CFD and EFD for the Series
60 CB =0.6 in steady drift motion. Journal of Marine Science and Technology,
7(1):17–30, August 2002.
[141] S. L. Toxopeus. Simulation and validation of the viscous flow around the Series 60
hull form at 10◦ drift angle. In 7th NuTTS Numerical Towing Tank Symposium,
Hamburg, Germany, October 2004.
[143] S. L. Toxopeus. Verification and validation of calculations of the viscous flow around
KVLCC2M in oblique motion. In 5th Osaka Colloquium on Advanced CFD Appli-
cations to Ship Flow and Hull Form Design, pages 200–209, Osaka, Japan, March
2005.
[146] S. L. Toxopeus, editor. SHWG collaborative CFD excercise - Bare hull DARPA
SUBOFF submarine at straight flight and drift angle. MARIN Report 21668-1-CPM,
November 2007.
[147] S. L. Toxopeus. Viscous-flow calculations for bare hull DARPA SUBOFF submarine
at incidence. International Shipbuilding Progress, 55(3):227–251, December 2008.
[148] S. L. Toxopeus. Deriving mathematical manoeuvring models for bare ship hulls using
viscous flow calculations. Journal of Marine Science and Technology, 14(1):30–38,
March 2009.
[150] S. L. Toxopeus and S. W. Lee. Free sailing tests for European Destroyer. MARIN
Report No. 22367-5-CPM, Wageningen, The Netherlands, February 2009.
140 References
[154] A. van der Ploeg and H. C. Raven. CFD-based optimization for minimal power and
wake field quality. In 11th International Symposium on Practical Design of Ships
and Other Floating Structures (PRADS), Rio de Janeiro, Brazil, September 2010.
[155] B. J. van Oers and S. L. Toxopeus. On the relation between flow behaviour and the
lateral force distribution acting on a ship in oblique motion. In 10th International
Cooperation on Marine Engineering Systems ICMES, London, UK, March 2006.
[161] M. Vogt, A. Cura Hochbaum, M. Schneider, and J. Dautel. Report on model tests.
Virtue Deliverable D3.1.3, December 2007.
[164] C. C. Whitfield. Steady and unsteady force and moment data on a DARPA2 sub-
marine. Master’s thesis, Virginia Polytechnic Institute and State University, August
1999.
[165] B.-S. Wu, F. Xing, X.-F. Kuang, and Q.-M. Miao. Investigation of hydrodynamic
characteristics of submarine moving close to the sea bottom with CFD methods.
Journal of Ship Mechanics, 9(3):19–28, June 2005.
[166] C. Yang and R. Löhner. Prediction of flows over an axisymmetric body with ap-
pendages. In 8th International Conference on Numerical Ship Hydrodynamics, Bu-
san, Korea, September 2003.
Samenvatting
Praktische toepassing van viskeuze omstromingsberekeningen voor
het simuleren van manoeuvrerende schepen
Acknowledgements
The work for this thesis was sponsored by the background research and development of
MARIN. I thank MARIN and in particular my former manager Frans Quadvlieg and
current manager Henk Prins for their support. I also appreciate the time and effort spent
by my supervisor René Huijsmans from Delft University of Technology during this work.
Thanks to him I had the opportunity to finalise this work with a PhD degree.
The work presented in this thesis regarding the DARPA SUBOFF and the Walrus
submarine was funded through TNO Defence, Security and Safety within the framework
of Programma V705 carried out for DMO of the Royal Netherlands Navy. Their support
is greatly acknowledged.
Another part (the work on the HTC) was funded by the Commission of the Euro-
pean Communities through the Integrated Project VIRTUE under grant 516201 in the
sixth Research and Technological Development Framework Programme (Surface Trans-
port Call). I thank the members of the manoeuvring Work Package of VIRTUE for their
discussions and insights during the project.
When starting work at MARIN, I was introduced to the subject of ship manoeu-
vring by my former colleague Jan Hooft. Thanks to him, I became interested in ship
manoeuvring simulations.
Furthermore, this work could not have been done without the support of various
colleagues at MARIN. I therefore acknowledge the help of (in random order): Jaap Windt
and Chris Klaij for their help in grid generation; Auke van der Ploeg for his efforts
in getting difficult Parnassos jobs to converge; Martin Hoekstra and Luı́s Eça (from
Instituto Superior Técnico, Portugal (IST)) for their advice regarding uncertainty ana-
lysis; Hoyte Raven for his insight into hydrodynamics in general and discussions about the
scope of work; Guilherme Vaz for his development and support of ReFRESCO; Bram
Starke for discussions on viscous-flow and help with getting started with Parnassos;
Erik van Wijngaarden for his textual suggestions about my thesis and our discussions
about being a PhD student; my colleagues at the Helpdesk for keeping the computers and
cluster alive. And all others who put up with me when I enthusiastically wanted to show
them my ColourFul Diagrams.
Also the work of Mark Bettle from University of New Brunswick is appreciated. He
conducted the calculations for the Walrus submarine presented in this thesis. It was a
pleasure to have you stay at MARIN and thanks for the great job, Mark!
146 Acknowledgements
Bas Overpelt managed and analysed the free sailing experiments for the HTC. Kees
Boers supervised also some of the tests. Thanks to both for their efforts and hard work.
Experimental data existing in the public domain has been used to validate the viscous-
flow calculations presented in this thesis. This contribution of valuable data of INSEAN,
NMRI, HSVA and Naval Surface Warfare Center, Carderock Division (NSWCCD) is
greatly appreciated.
Finally, I dearly thank my wife Aisah for her continuous encouragement and support
of my PhD work and her valuable advice on planning. Without you, I would still be
writing this thesis!
147
Curriculum Vitae
Serge Toxopeus was born on April 23rd , 1972 in Alkmaar, The Netherlands. He finished
high school at the C.S.G. Jan Arentsz in Alkmaar in 1990 and subsequently started to
study Maritime Engineering at Delft University of Technology, Faculty of Mechanical En-
gineering and Marine Technology (currently part of the Faculty of Mechanical, Maritime
and Materials Engineering). Serge graduated in 1996 at the Ship Hydromechanics De-
partment with a Master’s thesis on the dynamic stability and manoeuvring of planing
ships.
In the same year, he became an employee of the Maritime Research Institute Nether-
lands (MARIN) in Wageningen as Project Leader, Consultant, Project Manager and
Senior Project Manager (depending on the year) in the Manoeuvring Department. His
main tasks were the development and improvement of mathematical manoeuvring models
for surface ships and submarines and the assessment of the manoeuvrability of ships using
simulations and model tests. He was actively involved in commercial projects for about
eight years.
Around 2002, the desire arose to improve the manoeuvring simulations using advanced
techniques and therefore Serge started a PhD study regarding the practical application of
viscous-flow calculations for the simulation of manoeuvring ships. Since 2005, most of his
time was spent on this PhD study. The work was partly funded as background research
by MARIN and partly through the EU project VIRTUE. This thesis is the consolidation
of the work done in the past eight years.
To be able to fully concentrate on his research, Serge moved to the Research and
Development department in 2007. His present position is Senior Researcher.
148 Curriculum Vitae
Appendices
150 Appendices
25
20
15
10
δ [deg]
5
0
−5
−10
−15
−20
−25
0 100 200 300 400 500 600 700 800 900 1000
45
30
15
ψ [deg]
0
−15
−30
−45
0 100 200 300 400 500 600 700 800 900 1000
0.9
0.6
0.3
r [deg/s]
0
−0.3
−0.6
−0.9
0 100 200 300 400 500 600 700 800 900 1000
2.5
2
1.5
1
φ [deg]
0.5
0
−0.5
−1
−1.5
−2
−2.5
0 100 200 300 400 500 600 700 800 900 1000
6
5
4
V [m/s]
3
2
1
0
0 100 200 300 400 500 600 700 800 900 1000
15
10
5
βG [deg]
0
−5
−10
−15
0 100 200 300 400 500 600 700 800 900 1000
203010 106011
time 203008 106007
203006 106002
203013 203004 zz05.14−10.00
thick blue lines: simulation, others: experiments
178 Appendices
4000
3500
3000
2500
x [m]
2000
E
1500
1000
500
25
20
15
10
δ [deg]
5
0
−5
−10
−15
−20
−25
0 100 200 300 400 500 600 700 800 900 1000
45
30
15
ψ [deg]
0
−15
−30
−45
0 100 200 300 400 500 600 700 800 900 1000
0.9
0.6
0.3
r [deg/s]
0
−0.3
−0.6
−0.9
0 100 200 300 400 500 600 700 800 900 1000
2.5
2
1.5
1
φ [deg]
0.5
0
−0.5
−1
−1.5
−2
−2.5
0 100 200 300 400 500 600 700 800 900 1000
6
5
4
V [m/s]
3
2
1
0
0 100 200 300 400 500 600 700 800 900 1000
15
10
5
βG [deg]
0
−5
−10
−15
0 100 200 300 400 500 600 700 800 900 1000
203009 106010
time 203007 106006
203005 106005
203011 203003 zz05.14−−10.00
thick blue lines: simulation, others: experiments
180 Appendices
4000
3500
3000
2500
x [m]
2000
E
1500
1000
500
25
20
15
10
δ [deg]
5
0
−5
−10
−15
−20
−25
0 100 200 300 400 500 600 700 800 900 1000
45
30
15
ψ [deg]
0
−15
−30
−45
0 100 200 300 400 500 600 700 800 900 1000
0.9
0.6
0.3
r [deg/s]
0
−0.3
−0.6
−0.9
0 100 200 300 400 500 600 700 800 900 1000
2.5
2
1.5
1
φ [deg]
0.5
0
−0.5
−1
−1.5
−2
−2.5
0 100 200 300 400 500 600 700 800 900 1000
6
5
4
V [m/s]
3
2
1
0
0 100 200 300 400 500 600 700 800 900 1000
15
10
5
βG [deg]
0
−5
−10
−15
0 100 200 300 400 500 600 700 800 900 1000
106013
time 106009
106004
204003 zz05.14−20.00
thick blue lines: simulation, others: experiments
182 Appendices
3000
2500
2000
x [m]
E
1500
1000
500
25
20
15
10
δ [deg]
5
0
−5
−10
−15
−20
−25
0 100 200 300 400 500 600 700 800 900 1000
45
30
15
ψ [deg]
0
−15
−30
−45
0 100 200 300 400 500 600 700 800 900 1000
0.9
0.6
0.3
r [deg/s]
0
−0.3
−0.6
−0.9
0 100 200 300 400 500 600 700 800 900 1000
2.5
2
1.5
1
φ [deg]
0.5
0
−0.5
−1
−1.5
−2
−2.5
0 100 200 300 400 500 600 700 800 900 1000
6
5
4
V [m/s]
3
2
1
0
0 100 200 300 400 500 600 700 800 900 1000
15
10
5
βG [deg]
0
−5
−10
−15
0 100 200 300 400 500 600 700 800 900 1000
106012
time 106008
106003
204002 zz05.14−−20.00
thick blue lines: simulation, others: experiments
184 Appendices
3000
2500
2000
x [m]
E
1500
1000
500
25
20
15
10
δ [deg]
5
0
−5
−10
−15
−20
−25
0 50 100 150 200 250 300 350 400 450 500
45
30
15
ψ [deg]
0
−15
−30
−45
0 50 100 150 200 250 300 350 400 450 500
2
1.5
1
r [deg/s]
0.5
0
−0.5
−1
−1.5
−2
0 50 100 150 200 250 300 350 400 450 500
6
4.5
3
φ [deg]
1.5
0
−1.5
−3
−4.5
−6
0 50 100 150 200 250 300 350 400 450 500
10
8
V [m/s]
6
4
2
0
0 50 100 150 200 250 300 350 400 450 500
20
15
10
βG [deg]
5
0
−5
−10
−15
−20
0 50 100 150 200 250 300 350 400 450 500
108011
time 108007
108002
zz09.26−10.00
thick blue lines: simulation, others: experiments
186 Appendices
4000
3500
3000
2500
x [m]
2000
E
1500
1000
500
25
20
15
10
δ [deg]
5
0
−5
−10
−15
−20
−25
0 50 100 150 200 250 300 350 400 450 500
45
30
15
ψ [deg]
0
−15
−30
−45
0 50 100 150 200 250 300 350 400 450 500
2
1.5
1
r [deg/s]
0.5
0
−0.5
−1
−1.5
−2
0 50 100 150 200 250 300 350 400 450 500
6
4.5
3
φ [deg]
1.5
0
−1.5
−3
−4.5
−6
0 50 100 150 200 250 300 350 400 450 500
10
8
V [m/s]
6
4
2
0
0 50 100 150 200 250 300 350 400 450 500
20
15
10
βG [deg]
5
0
−5
−10
−15
−20
0 50 100 150 200 250 300 350 400 450 500
108010
time 108006
108001
zz09.26−−10.00
thick blue lines: simulation, others: experiments
188 Appendices
4000
3500
3000
2500
x [m]
2000
E
1500
1000
500
25
20
15
10
δ [deg]
5
0
−5
−10
−15
−20
−25
0 50 100 150 200 250 300 350 400 450 500
45
30
15
ψ [deg]
0
−15
−30
−45
0 50 100 150 200 250 300 350 400 450 500
2
1.5
1
r [deg/s]
0.5
0
−0.5
−1
−1.5
−2
0 50 100 150 200 250 300 350 400 450 500
6
4.5
3
φ [deg]
1.5
0
−1.5
−3
−4.5
−6
0 50 100 150 200 250 300 350 400 450 500
10
8
V [m/s]
6
4
2
0
0 50 100 150 200 250 300 350 400 450 500
20
15
10
βG [deg]
5
0
−5
−10
−15
−20
0 50 100 150 200 250 300 350 400 450 500
108005
time 108009
108013
zz09.26−20.00
thick blue lines: simulation, others: experiments
190 Appendices
3000
2500
2000
x [m]
E
1500
1000
500
−1200 −1000 −800 −600 −400 −200 0 200 400 600 800
yE [m]
108005
108009
108013
zz09.26−20.00
thick blue lines: simulation, others: experiments
Figure pages – zig-zag 191
25
20
15
10
δ [deg]
5
0
−5
−10
−15
−20
−25
0 50 100 150 200 250 300 350 400 450 500
45
30
15
ψ [deg]
0
−15
−30
−45
0 50 100 150 200 250 300 350 400 450 500
2
1.5
1
r [deg/s]
0.5
0
−0.5
−1
−1.5
−2
0 50 100 150 200 250 300 350 400 450 500
6
4.5
3
φ [deg]
1.5
0
−1.5
−3
−4.5
−6
0 50 100 150 200 250 300 350 400 450 500
10
8
V [m/s]
6
4
2
0
0 50 100 150 200 250 300 350 400 450 500
20
15
10
βG [deg]
5
0
−5
−10
−15
−20
0 50 100 150 200 250 300 350 400 450 500
108012
time 108008
108003
zz09.26−−20.00
thick blue lines: simulation, others: experiments
192 Appendices
3000
2500
2000
x [m]
E
1500
1000
500
−800 −600 −400 −200 0 200 400 600 800 1000 1200
yE [m]
108012
108008
108003
zz09.26−−20.00
thick blue lines: simulation, others: experiments
Figure pages – turning-circle 193
40
30
δ [deg]
20
10
0
0 100 200 300 400 500 600 700 800 900 1000 1100 1200 1300 1400 1500
0
−90
−180
−270
ψ [deg]
−360
−450
−540
−630
−720
−810
0 100 200 300 400 500 600 700 800 900 1000 1100 1200 1300 1400 1500
0
−0.2
−0.4
r [deg/s]
−0.6
−0.8
−1
−1.2
0 100 200 300 400 500 600 700 800 900 1000 1100 1200 1300 1400 1500
1.2
0.8
0.4
φ [deg]
0
−0.4
−0.8
−1.2
0 100 200 300 400 500 600 700 800 900 1000 1100 1200 1300 1400 1500
6
5
4
V [m/s]
3
2
1
0
0 100 200 300 400 500 600 700 800 900 1000 1100 1200 1300 1400 1500
30
25
20
βG [deg]
15
10
5
0
0 100 200 300 400 500 600 700 800 900 1000 1100 1200 1300 1400 1500
time
107002
tc05.14−35.00
thick blue lines: simulation, others: experiments
194 Appendices
2500
2000
1500
x [m]
E
1000
500
107002
tc05.14−35.00
thick blue lines: simulation, others: experiments
Figure pages – turning-circle 195
−10
δ [deg]
−20
−30
−40
0 100 200 300 400 500 600 700 800 900 1000 1100 1200 1300 1400 1500
810
720
630
540
ψ [deg]
450
360
270
180
90
0
0 100 200 300 400 500 600 700 800 900 1000 1100 1200 1300 1400 1500
1.2
1
0.8
r [deg/s]
0.6
0.4
0.2
0
0 100 200 300 400 500 600 700 800 900 1000 1100 1200 1300 1400 1500
1.2
0.8
0.4
φ [deg]
0
−0.4
−0.8
−1.2
0 100 200 300 400 500 600 700 800 900 1000 1100 1200 1300 1400 1500
6
5
4
V [m/s]
3
2
1
0
0 100 200 300 400 500 600 700 800 900 1000 1100 1200 1300 1400 1500
0
−5
−10
βG [deg]
−15
−20
−25
−30
0 100 200 300 400 500 600 700 800 900 1000 1100 1200 1300 1400 1500
time
107001
tc05.14−−35.00
thick blue lines: simulation, others: experiments
196 Appendices
1200
1000
800
x [m]
600
E
400
200
107001
tc05.14−−35.00
thick blue lines: simulation, others: experiments
Figure pages – turning-circle 197
40
30
δ [deg]
20
10
0
0 200 400 600 800 1000 1200 1400 1600 1800 2000
0
−90
−180
−270
ψ [deg]
−360
−450
−540
−630
−720
−810
0 200 400 600 800 1000 1200 1400 1600 1800 2000
0
−0.2
−0.4
r [deg/s]
−0.6
−0.8
−1
−1.2
0 200 400 600 800 1000 1200 1400 1600 1800 2000
1.2
0.8
0.4
φ [deg]
0
−0.4
−0.8
−1.2
−1.6
−2
0 200 400 600 800 1000 1200 1400 1600 1800 2000
6
5
4
V [m/s]
3
2
1
0
0 200 400 600 800 1000 1200 1400 1600 1800 2000
30
25
20
βG [deg]
15
10
5
0
0 200 400 600 800 1000 1200 1400 1600 1800 2000
time
107005
tc05.14−25.00
thick blue lines: simulation, others: experiments
198 Appendices
1600
1400
1200
1000
x [m]
E
800
600
400
200
107005
tc05.14−25.00
thick blue lines: simulation, others: experiments
Figure pages – turning-circle 199
−10
δ [deg]
−20
−30
−40
0 200 400 600 800 1000 1200 1400 1600 1800 2000
810
720
630
540
ψ [deg]
450
360
270
180
90
0
0 200 400 600 800 1000 1200 1400 1600 1800 2000
1.2
1
0.8
r [deg/s]
0.6
0.4
0.2
0
0 200 400 600 800 1000 1200 1400 1600 1800 2000
1.2
0.9
0.6
0.3
0
φ [deg]
−0.3
−0.6
−0.9
−1.2
−1.5
−1.8
−2.1
0 200 400 600 800 1000 1200 1400 1600 1800 2000
6
5
4
V [m/s]
3
2
1
0
0 200 400 600 800 1000 1200 1400 1600 1800 2000
0
−5
−10
βG [deg]
−15
−20
−25
−30
0 200 400 600 800 1000 1200 1400 1600 1800 2000
time
107004
tc05.14−−25.00
thick blue lines: simulation, others: experiments
200 Appendices
1200
1000
800
600
x [m]
E
400
200
107004
tc05.14−−25.00
thick blue lines: simulation, others: experiments
Figure pages – turning-circle 201
40
30
δ [deg]
20
10
0
0 200 400 600 800 1000 1200 1400 1600 1800 2000
0
−90
−180
−270
ψ [deg]
−360
−450
−540
−630
−720
−810
0 200 400 600 800 1000 1200 1400 1600 1800 2000
0
−0.2
−0.4
r [deg/s]
−0.6
−0.8
−1
−1.2
0 200 400 600 800 1000 1200 1400 1600 1800 2000
1.2
0.9
0.6
0.3
0
φ [deg]
−0.3
−0.6
−0.9
−1.2
−1.5
−1.8
−2.1
0 200 400 600 800 1000 1200 1400 1600 1800 2000
6
5
4
V [m/s]
3
2
1
0
0 200 400 600 800 1000 1200 1400 1600 1800 2000
30
25
20
βG [deg]
15
10
5
0
0 200 400 600 800 1000 1200 1400 1600 1800 2000
time
107007
tc05.14−15.00
thick blue lines: simulation, others: experiments
202 Appendices
1200
1000
800
600
x [m]
E
400
200
107007
tc05.14−15.00
thick blue lines: simulation, others: experiments
Figure pages – turning-circle 203
−10
δ [deg]
−20
−30
−40
0 200 400 600 800 1000 1200 1400 1600 1800 2000
810
720
630
540
ψ [deg]
450
360
270
180
90
0
0 200 400 600 800 1000 1200 1400 1600 1800 2000
1.2
1
0.8
r [deg/s]
0.6
0.4
0.2
0
0 200 400 600 800 1000 1200 1400 1600 1800 2000
1.2
0.9
0.6
0.3
0
φ [deg]
−0.3
−0.6
−0.9
−1.2
−1.5
−1.8
−2.1
0 200 400 600 800 1000 1200 1400 1600 1800 2000
6
5
4
V [m/s]
3
2
1
0
0 200 400 600 800 1000 1200 1400 1600 1800 2000
0
−5
−10
βG [deg]
−15
−20
−25
−30
0 200 400 600 800 1000 1200 1400 1600 1800 2000
time
107008
tc05.14−−15.00
thick blue lines: simulation, others: experiments
204 Appendices
1000
800
600
x [m]
E
400
200
−200
−100 0 100 200 300 400 500 600 700 800
yE [m]
107008
tc05.14−−15.00
thick blue lines: simulation, others: experiments
Figure pages – turning-circle 205
40
30
δ [deg]
20
10
0
0 100 200 300 400 500 600 700
0
−90
−180
−270
ψ [deg]
−360
−450
−540
−630
−720
−810
0 100 200 300 400 500 600 700
0
−0.5
r [deg/s]
−1
−1.5
−2
−2.5
0 100 200 300 400 500 600 700
2
φ [deg]
−2
−4
0 100 200 300 400 500 600 700
10
8
V [m/s]
6
4
2
0
0 100 200 300 400 500 600 700
30
25
20
βG [deg]
15
10
5
0
0 100 200 300 400 500 600 700
104021
time 104027
104029
205003 tc09.26−35.00
thick blue lines: simulation, others: experiments
206 Appendices
1200
1000
800
x [m]
600
E
400
200
−10
δ [deg]
−20
−30
−40
0 100 200 300 400 500 600 700
810
720
630
540
ψ [deg]
450
360
270
180
90
0
0 100 200 300 400 500 600 700
2.5
2
r [deg/s]
1.5
1
0.5
0
0 100 200 300 400 500 600 700
3
2
1
φ [deg]
0
−1
−2
−3
−4
0 100 200 300 400 500 600 700
10
8
V [m/s]
6
4
2
0
0 100 200 300 400 500 600 700
0
−5
−10
βG [deg]
−15
−20
−25
−30
0 100 200 300 400 500 600 700
104028
time 104019
104026
205002 tc09.26−−35.00
thick blue lines: simulation, others: experiments
208 Appendices
1200
1000
800
x [m]
600
E
400
200
40
30
δ [deg]
20
10
0
0 100 200 300 400 500 600 700 800 900 1000
0
−90
−180
−270
ψ [deg]
−360
−450
−540
−630
−720
−810
0 100 200 300 400 500 600 700 800 900 1000
0
−0.5
r [deg/s]
−1
−1.5
−2
−2.5
0 100 200 300 400 500 600 700 800 900 1000
4.5
3.6
2.7
1.8
φ [deg]
0.9
0
−0.9
−1.8
−2.7
−3.6
0 100 200 300 400 500 600 700 800 900 1000
10
8
V [m/s]
6
4
2
0
0 100 200 300 400 500 600 700 800 900 1000
30
25
20
βG [deg]
15
10
5
0
0 100 200 300 400 500 600 700 800 900 1000
time
104023
tc09.26−25.00
thick blue lines: simulation, others: experiments
210 Appendices
1000
800
600
x [m]
E
400
200
104023
tc09.26−25.00
thick blue lines: simulation, others: experiments
Figure pages – turning-circle 211
−10
δ [deg]
−20
−30
−40
0 100 200 300 400 500 600 700 800 900 1000
810
720
630
540
ψ [deg]
450
360
270
180
90
0
0 100 200 300 400 500 600 700 800 900 1000
2.5
2
r [deg/s]
1.5
1
0.5
0
0 100 200 300 400 500 600 700 800 900 1000
3
2
1
φ [deg]
0
−1
−2
−3
−4
0 100 200 300 400 500 600 700 800 900 1000
10
8
V [m/s]
6
4
2
0
0 100 200 300 400 500 600 700 800 900 1000
0
−5
−10
βG [deg]
−15
−20
−25
−30
0 100 200 300 400 500 600 700 800 900 1000
time
104022
tc09.26−−25.00
thick blue lines: simulation, others: experiments
212 Appendices
1000
800
600
x [m]
E
400
200
104022
tc09.26−−25.00
thick blue lines: simulation, others: experiments
Figure pages – turning-circle 213
40
30
δ [deg]
20
10
0
0 100 200 300 400 500 600 700 800 900 1000
0
−90
−180
−270
ψ [deg]
−360
−450
−540
−630
−720
−810
0 100 200 300 400 500 600 700 800 900 1000
0
−0.5
r [deg/s]
−1
−1.5
−2
−2.5
0 100 200 300 400 500 600 700 800 900 1000
4
3.2
2.4
φ [deg]
1.6
0.8
0
−0.8
−1.6
−2.4
0 100 200 300 400 500 600 700 800 900 1000
10
8
V [m/s]
6
4
2
0
0 100 200 300 400 500 600 700 800 900 1000
30
25
20
βG [deg]
15
10
5
0
0 100 200 300 400 500 600 700 800 900 1000
time
104025
tc09.26−15.00
thick blue lines: simulation, others: experiments
214 Appendices
1000
900
800
700
600
x [m]
500
E
400
300
200
100
104025
tc09.26−15.00
thick blue lines: simulation, others: experiments
Figure pages – turning-circle 215
−10
δ [deg]
−20
−30
−40
0 100 200 300 400 500 600 700 800 900 1000
810
720
630
540
ψ [deg]
450
360
270
180
90
0
0 100 200 300 400 500 600 700 800 900 1000
2.5
2
r [deg/s]
1.5
1
0.5
0
0 100 200 300 400 500 600 700 800 900 1000
2.4
1.6
0.8
φ [deg]
0
−0.8
−1.6
−2.4
−3.2
−4
0 100 200 300 400 500 600 700 800 900 1000
10
8
V [m/s]
6
4
2
0
0 100 200 300 400 500 600 700 800 900 1000
0
−5
−10
βG [deg]
−15
−20
−25
−30
0 100 200 300 400 500 600 700 800 900 1000
time
104024
tc09.26−−15.00
thick blue lines: simulation, others: experiments
216 Appendices
1000
800
600
x [m]
E
400
200
−200
−100 0 100 200 300 400 500 600 700 800
yE [m]
104024
tc09.26−−15.00
thick blue lines: simulation, others: experiments