Vous êtes sur la page 1sur 195

Justus Heimann

CFD Based Optimization of the


Wave-Making Characteristics of Ship Hulls

CFD Based Optimization of the


Wave-Making Characteristics of Ship Hulls

vorgelegt von
Diplom-Ingenieur
Justus Heimann
aus Berlin

von der Fakultt V Verkehrs- und Maschinensysteme


der Technischen Universitt Berlin
zur Erlangung des akademischen Grades
Doktor der Ingenieurwissenschaften
Dr.-Ing.
genehmigte Dissertation
Promotionsausschuss:
Vorsitzender:
Berichter:
Berichter:
Berichter:

Prof. Dr.-Ing. Klaus Brie


Prof. Dr.-Ing. Gnther F. Clauss
Prof. Dr.-Ing. Gerhard Jensen
Prof. Dr.-Ing. Lothar Birk

Tag der wissenschaftlichen Aussprache: 2. Mrz 2005


Berlin 2005
D 83

CFD Based Optimization of the


Wave-Making Characteristics of Ship Hulls

Dipl.-Ing. Justus Heimann

Approved by the
Faculty of Mechanical Engineering and Transport Systems
of the Technical University Berlin
in partial fulfillment of the requirements of the degree of
Doktor der Ingenieurwissenschaften
Dr.-Ing.

Doctoral committee:
Prof.
Prof.
Prof.
Prof.

Dr.-Ing.
Dr.-Ing.
Dr.-Ing.
Dr.-Ing.

Klaus Brie (Chair) Technical University Berlin


Gnther F. Clauss Technical University Berlin
Gerhard Jensen Technical University Hamburg-Harburg
Lothar Birk University of New Orleans

Day of the oral examination: 2nd March 2005


Berlin 2005
D 83

Bibliografische Information Der Deutschen Bibliothek


Die Deutsche Bibliothek verzeichnet diese Publikation in der
Deutschen Nationalbibliografie; detaillierte bibliografische Daten
sind im Internet ber <http://dnb.ddb.de> abrufbar.
ISBN 3-89820-445-6
Dissertation, Technische Universitt Berlin, 2005
Dieses Werk ist urheberrechtlich geschtzt.
Alle Rechte, auch die der bersetzung, des Nachdruckes und der Vervielfltigung
des Buches, oder Teilen daraus, vorbehalten. Kein Teil des Werkes darf ohne
schriftliche Genehmigung des Verlages in irgendeiner Form reproduziert oder
unter Verwendung elektronischer Systeme verarbeitet, vervielfltigt oder
verbreitet werden.
This document is protected by copyright.
No part of this document may be reproduced in any form by any means without
prior written authorization of Mensch & Buch Verlag.

Mensch & Buch Verlag

Nordendstr. 75 - 13156 Berlin 030-45494866


info@menschundbuch.de www.menschundbuch.de

Preface
Several years ago when I was an enthusiastic regatta sailor the foundation for the present research
work was laid. Many times during the races occurred the challenging question how to improve my
own performance and that of my Laser dinghy very likely, I wasnt the only one!
My tactical and sailing skills improved over the years. So I refined techniques like roll tacking and
jibing, sailing close-hauled and upright on upwind courses, steering through waves with minimal speed
loss, reducing the wetted hull surface on downwind courses or speeding up in planning conditions. But
how to improve the hull hydrodynamics? Since the Laser dinghy is the ultimate one design there is no
room left for hull modifications. Nevertheless, my friends and I did or at least we thought we did
our best to optimize the hull hydrodynamics, mostly within the rules. For instance we polished the
underwater hull with sandpaper to generate a micro air cushion, we shaped the leading and trailing
edges of the centreboard to increase the lift-drag ratio and we fitted a little (invisible) flapper to
the transom edge to improve the planning performance. However, all this hydrodynamic fine tuning
remained rather a matter of faith than of measurable benefit.
Nowadays, after several years of study and research think I have come a bit closer to the answer of how
to improve the hull hydrodynamics measurably at least the size of the dinghies has improved a lot.
The present research work was undertaken at the Division of Naval Architecture and Ocean Engineering of the Technical University Berlin. Preparation and finalization of the thesis were undertaken at the
premises and with the generous support of FRIENDSHIP SYSTEMS GmbH. For a long educational
period I was supervised by Prof. Dr.-Ing. Dr. h.c. Horst Nowacki in whose team I had the pleasure to
work as a research and teaching scientist from 1995 to 2000 in the field of ship geometric modelling,
numerical hydrodynamics and optimization. I wish to express my sincere thanks to Prof. Nowacki for
his valuable advice, inspiration, criticism and support and his companionship of the present work.
I like to express my sincere thanks to Prof. Dr.-Ing. Gnther F. Clauss for guiding my first scientific
steps and for his commitment in the doctoral examination to become thesis supervisor. I am very grateful to Prof. Dr.-Ing. Gerhard Jensen and Prof. Dr.-Ing. Lothar Birk for taking interest in my work and
for being thesis supervisors. Furthermore, I would like to thank Prof. Dr.-Ing. Klaus Brie for chairing
the doctoral examination.
I would like to say thank you to Prof. Dr. Carl-Erik Janson from Chalmers University of Technology and
FLOWTECH Int. AB, Gteborg, Sweden, for his continuous support, many fruitful discussions and for
customizing the CFD system SHIPFLOW according to my needs. Many thanks goes to Prof. Dr. Lars
Larsson and the whole FLOWTECH team for kindly making available the SHIPFLOW system.
Very special thanks I owe to my friend Dr.-Ing. Stefan Harries for being an excellent colleague for
many years and for his invaluable support and encouragement in finalizing this work.
A nice and valuable time I spent with my friends and colleagues, particularly Claus Abt, Frauke
Baumgrtel, zgr Baskaya, Dr.-Ing. Carl-Uwe Bttner, Jrn Hinnenthal, Dr.-Ing. Karsten Hochkirch,
Dr.-Ing. Katja Jacobsen, Bernd L. Kther, Bernhard Krger, Dr.-Ing. Yeon-Seung Lee, Prof. Dr.Ing. Detlef Schulze and, last but not least, Dr.-Ing. Geir Westgaard.
Thanks for being close friends in every respect to Dr. Uwe Laude, Stephan Meyer and Suse Pfeiffer.
I wish to express my deep gratitude to my parents Prof. Ingrid and Prof. Gerhard Heimann and to my
brother Felix Heimann. Finally, without the love, the steadfast support, endless patience and continuous
pushing of the woman I love, Antje Kmmerer, this work would not have come to its happy end.
Justus Heimann Potsdam, March 2005

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

Abstract
Ship optimization based on computer simulations has become a decisive factor in the development of
new, economically efficient and environmentally friendly ship hull forms. An important task at an early
design stage is the optimization of the wave-making characteristics of the ship hull since for fast ships a
considerable resistance component stems from the steady ship wave system. Moreover, the ship waves
cause adverse effects in the far field as the wash hits the shore line or other vessels.
A novel hull form optimization approach has been developed and implemented. It builds on a CFD
(Computational Fluid Dynamics) based evaluation of the nonlinear ship wave pattern, on realization of
the cause and effect relation of hull variations and their impact on wave formation, which is accessed
by a perturbation approach, and on wave cut analysis (WCA). WCA yields an excellent assessment
of the wave-making characteristics of a hull form in terms of its free wave spectrum and the wave
associated pattern resistance. These features are highly integrated and controlled by a fully automated
optimization scheme. The scheme is specific in the way it tackles the optimization problem:
The hull adaptation is driven directly by hydrodynamics, avoiding prerequisites to the shape
representation and to the hull modification method.
Wave cut analysis yields the objective function of optimization in terms of the free wave spectrum
and the wave pattern resistance. This considerably improves the system identification allowing
a focused optimization.
The hull optimization is carried out directly for the effective wetted hull portion of the advancing
ship including the effects of dynamic trim and sinkage and wave formation along the hull.
The optimization process is established in terms of an iterative marching scheme of successive
sub-optimization loops. In each loop a region of the solution space is mapped to a simplified
convex quadratic image which merely possesses a single minimum determined by the active
constraints. This enables a straightforward solution procedure and a simultaneous treatment of
a large number of locally acting optimization variables, introducing much freedom to the hull
variation.
All aspects of the proposed optimization approach are presented. Applications of the optimization
scheme to practical ship hull forms show the following:
The wave-making characteristics can be considerably improved with a tangible reduction of the
wave (pattern) resistance.
Hull variations are driven by the optimization to the expected optimum hull shapes with the hull
geometry fully self-adjusting according to the optimization requirements.
Optimal interferences of local wave trains are enforced both in amplitude and phase which results
in a beneficial cancelation of the wave trains.
The specific fingerprints of locally confined hull variations (e.g., at the bulbous bow) can be
effectively traced in the wave spectrum and in the wave pattern resistance to allow a focused
minimization of particular adverse wave components.
The system identification can be significantly improved by utilizing additional information in
terms of the wave spectrum and the distribution of the wave pattern resistance over the waves
range.

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

Kurzfassung
Die Entwicklung konomisch und kologisch effizienter Schiffe erfordert im groem Mae den Einsatz moderner, rechnergesttzter Verfahren. Eine ausschlaggebende Rolle hierbei spielt die auf Computersimulationen basierende hydrodynamische Optimierung des Schiffsrumpfes, die in einem frhen
Entwurfsstadium erfolgen sollte. Ein erheblicher Teil des Rumpfwiderstandes und damit des Leistungsbedarfs wird bei schnellen Schiffen durch Energieverluste infolge der Ausbildung des stationren
Schiffswellensystems verursacht. Zudem gefhrden Schiffswellen Kstenschutzmanahmen, Uferbebauungen, die Ufervegetation, Hafenanlagen und andere Schiffe sowie Badende.
Im Rahmen der Arbeit wurde ein neuartiger Optimierungsansatz entwickelt und implementiert, der
sich auf die Wellenbildungseigenschaften von Schiffsrmpfen konzentriert. Das Verfahren beruht auf
der rechnergesttzten Simulation des nichtlinearen, stationren Schiffswellenfeldes, auf der Betrachtung des Ursache-Wirkungs-Zusammenhanges von Rumpfvariationen und deren Auswirkung auf die
Wellenbildung sowie auf der Wellenschnittmethodik, die ein ntzliches Instrument fr die Bewertung
der Wellenbildungseigenschaften von Rumpfgeometrien in Form von freien Wellenspektren sowie des
Wellen(bild)widerstands liefert. Diese Bausteine wurden zu einem voll automatischen Optimierungssystem integriert. Das Problem der Wellenwiderstandsminimierung wird hierbei auf eine spezielle Weise angegangen:
Rumpfformadaptionen werden direkt ber Strmungsgren gesteuert, unabhngig von einer
speziellen Geometriedarstellung sowie einem speziellen Verfahren zur Geometrievariation.
Mit Hilfe der Wellenschnittmethodik lsst sich eine aussagekrftige Zielfunktion in Form von
freien Wellenspektren und des Wellen(bild)widerstands ber den Einzelwellenkomponenten definieren, was eine erheblich verbesserte Systemidentifikation und damit eine zielgerichtete Optimierung ermglicht.
Die Optimierung erfolgt direkt fr den benetzten Rumpf unterhalb der Wellenkontur des fahrenden Schiffes in seiner dynamischen Schwimmlage.
Der Optimierungsprozess baut auf aufeinander folgenden Sub-Optimierungsschleifen, in denen
jeweils ein Ausschnitt des Lsungsraums vereinfacht auf einen konvex, quadratischen Raum
abgebildet wird, von dem man wei, dass er ein einziges Minimum aufweist, das von den aktiven
Nebenbedingungen bestimmt wird. Dadurch wird das Lsungsverfahren beschleunigt und die
Verwendung einer groen Zahl von Optimierungsvariablen mit jeweils lokalem Einflussbereich
ermglicht. Dies generiert ein substantielles Ma an Freiheit fr die Formvariation.
Die vorliegende Arbeit stellt alle Aspekte des Optimierungsansatzes ausfhrlich dar. Die Anwendung
des Verfahrens auf praxisnahe Rumpfformen weist folgendes nach:
Die Wellenbildung lsst sich deutlich verringern. Dies ist mit einer erheblichen Reduktion des
Wellen(bild)widerstand verbunden.
Rumpfgeometrievariationen sind entsprechend den Erfordernissen der Optimierung so einstellbar, dass sich erwartete optimale Geometrienderungen ergeben. Freiheiten der Rumpfvariation
knnen dort in das System eingebracht werden, wo im spezifischen Entwurfszusammenhang
noch Freirume vorliegen.
Optimale berlagerungen verschiedener Einzelwellensysteme in Amplitude und Phase entlang
des Rumpfes stellen sich so ein, dass die Einzelwellensysteme sich gegenseitig auslschen.

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

Der Fingerabdruck lokal begrenzter Rumpfvariationen (z.B. im Bugwulstbereich) lsst sich przise bis in die Wellenspektren und bis zum Wellen(bild)widerstand verfolgen, was eine zielgerichtete Minimierung besonders schdlicher Wellenanteile ermglicht.
Mit den Wellenspektren und der Verteilung des Wellen(bild)widerstands ber den Einzelwellenkomponenten werden wertvolle Informationen bereit gestellt, die eine erheblich verbesserte
Systemidentifikation und damit eine zielgerichtete Optimierung erlauben.

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

Contents

1
2

Introduction
Hydrodynamic optimization approach

11

2.1
2.2
2.3

11
13
14

3.2

CFD simulation . . . . . . . . . . . .
3.1.1 Boundary value problem . . .
3.1.2 Discretization . . . . . . . . .
3.1.3 Numerical solution . . . . . .
3.1.4 The CFD system SHIPFLOW
3.1.5 Application aspects . . . . . .
Wave cut analysis . . . . . . . . . . .
3.2.1 Boundary value problem . . .
3.2.2 Longitudinal wave cut analysis
3.2.3 Truncation correction . . . . .
3.2.4 Application aspects . . . . . .

19

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

Sensitivity analysis

4.1
4.2

4.3

4.4
4.5
4.6
4.7
4.8
5

Motivation and goals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .


Cause and effect chain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Process flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Hydrodynamic analysis

3.1

Design variables . . . . . . . . . . . . . . . . . . . . . . . . .
Perturbation approach . . . . . . . . . . . . . . . . . . . . . .
4.2.1 Perturbation of the discretized boundary value problem
4.2.2 Perturbation magnitudes . . . . . . . . . . . . . . . .
Panel relocation scheme . . . . . . . . . . . . . . . . . . . .
4.3.1 Joint relocation of hull panels . . . . . . . . . . . . .
4.3.2 Relocation range . . . . . . . . . . . . . . . . . . . .
Process flow of sensitivity analysis . . . . . . . . . . . . . . .
Impact on hull form and main form parameters . . . . . . . .
Impact on wave elevation and flow quantities . . . . . . . . .
Impact on free wave spectra and wave pattern resistance . . .
Evaluation of gradients . . . . . . . . . . . . . . . . . . . . .

43

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

Improvement of the wave-making characteristics

5.1
5.2

19
20
24
25
26
28
28
30
32
35
36

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

43
45
45
51
53
56
59
70
71
79
84
89
95

Constrained minimization problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96


Optimization functional . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
5.2.1 Objective function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

Contents

5.3
5.4
6

Applications

6.1
6.2
6.3
7

5.2.2 Penalty functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108


Solution scheme . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
Hull form adaptation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
117

Introductory comments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117


Wigley hull . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
FantaRoRo ferry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134

Conclusion

7.1
7.2

161

Main achievements and contributions . . . . . . . . . . . . . . . . . . .


Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.2.1 Improvement and validation . . . . . . . . . . . . . . . . . . . .
7.2.2 Coupling to high level geometric modelling and variation schemes
7.2.3 Coupling to seakeeping optimization methods . . . . . . . . . . .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

Lists

165

Bibliography . . . . .
Figures . . . . . . . .
Tables . . . . . . . . .
Symbols, Abbreviations

161
162
162
163
164

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

165
173
179
181

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

1 Introduction
In recent time ship optimization based on computer simulations has become a decisive factor in the
development of new, economically efficient and environmentally friendly ship hull forms. An important task at an early design stage is the optimization of the hydrodynamic properties of the ship hull.
For instance, the reduction of the power consumption of a medium-sized ferry typical of fast short-sea
shipping by 1% already reduces the annual fuel costs by about $50.000 and the annual greenhouse gas
emissions by about 350 t.
A considerable resistance component stems from the steady ship wave system even when sailing in
relatively calm water. Fast displacement type ships consume 50% and above of their installed power
to overcome the wave resistance. The waves generated also cause adverse effects in the far field as the
wash hits the shore line or other vessels. It therefore is mandatory to reduce wave-making as much as
possible.
In order to do so early in the design process a novel hull form optimization approach has been developed
and implemented. The optimization process is fully automated requiring no user interaction, however,
permitting it. The steady wave system of a ship moving through otherwise calm water is approximated
by means of CFD (Computational Fluid Dynamics) simulation applying the state of the art nonlinear
free surface Rankine panel module of the SHIPFLOW system. The proposed optimization method
aims at an improvement of the wave-making characteristics, i.e., the minimization of the energy losses
associated with the wave formation along the hull. A valuable measure of the energy losses is attained
by wave cut analysis (WCA) which allows to determine the free wave spectrum and the wave pattern
resistance. Wave cut analysis yields the wave pattern resistance while by convention an integration
of the longitudinal pressure components over the hull provides the wave resistance.
The present optimization approach substantially differs from other approaches to wave resistance minimization in the way it tackles this objective. The substantial differences are:
Hull variations are driven directly by hydrodynamics, i.e., the hull shape is controlled via a
selected flow quantity along the hull surface. Neither, are prerequisites to the hull geometry
representation and its characteristic features nor to the hull variation method imposed. The optimization variables are flow related, constituting a link between the hull geometry and their
hydrodynamic properties.
The objective function of optimization is directly related to the steady ship wave systems in
terms of the free wave spectrum and the wave pattern resistance as determined by a longitudinal
wave cut analysis. This improves the system identification by tracing hull variations up to their
respective fingerprints in the wave spectral distribution and in the wave pattern resistance. Thus,
waves can be influenced and reduced more systematically and with higher control.
The hull optimization is carried out for the effective floating position of the advancing ship
including trim and sinkage. Variations are directly applied to the wetted hull portion under (and
slightly above) the actual wavy free surface level.

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

1 Introduction
The optimization process is established in terms of an iterative marching scheme of successive
sub-optimization loops which built on top of each other. In each sub-optimization the boundary
value problem is accessed by a linear perturbation approach around the current base point yielding a convex quadratic image of the solution space in a trust-region, a sub-optimization space
around the current base point. This simplified image of the solution space is known to possess a
single minimum determined by the active constraints. Thus, fast and straightforward evaluation
of the gradients of the objective function and the constraints is enabled, allowing a simultaneous
treatment of numerous, locally acting optimization variables which introduces great freedom to
the hull variation.
This particular optimization concept was conceived to pursue the following goals:
Optimization of the wave-making characteristics, i.e., tangible reduction of the wave pattern
resistance of the hull at its dynamic floating position.
Self-adjusting hull geometry variation, driven directly and solely by the hydrodynamic optimization.
Concerted reduction of particular adverse wave components.
Improvement of the system identification, i.e., correlation of cause and effect of local hull variations and their respective fingerprints in the wave spectrum and in the wave pattern resistance.
All aspects of the proposed optimization approach are presented. Chapter 2 introduces the optimization
approach and outlines the process flow. Chapter 3 elaborates the hydrodynamic analysis in terms of
CFD simulation of the nonlinear free surface flow and it represents the wave cut analysis technique
as applied to the computed wave pattern. Chapter 4 substantiates the particular choice of flow related
optimization variables. A perturbation approach is derived to access the boundary value problem in
a straightforward manner. The impact of variations on the hull form, on the flow field, the free surface wave elevation and finally the free wave spectrum and the wave pattern resistance is presented.
The compilation of gradients matrices for the objective function and the constraints is discussed. In
Chapter 5 the constrained minimization problem is introduced, the optimization functional is set-up,
the necessary conditions for a minimum are applied, and the solution by an iterative weighting scheme
is outlined. The chapter closes with a discussion of the hull form adaptation scheme. In chapter 6
the capabilities of the present optimization approach are illustrated by means of one academic and one
practical ship optimization application. The former refers to the standard Wigley hull, the latter to a
state of the art twin screw ferry featuring a bulbous bow and tunnelled transom stern. Finally, in chapter 7 the main achievements and contributions of the present work are summarized and an outlook to
further research work is given.

10

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

2 Hydrodynamic optimization approach


The basic ideas underlying the presented hydrodynamic ship hull optimization approach are
outlined. The pursued goals are stated. The optimization concept and the process flow is
described.
Over the last decade ship optimization based on computer simulations has become an indispensable
factor in the development of new hull forms. Optimization applications comprise all engineering tasks
involved in the design and construction process of a ship. However, an important task at an early design
stage is the optimization of the hydrodynamic properties of the ship hull. In this respect the reduction
of the ship resistance and, hence, the power consumption pays off economically for the shipowner
and/or the operator in terms of savings of operating costs. An additional benefit is less environmental
impact due to reduced emissions of exhaust-gas and noise, less wastage of fossil fuel and a protection
of the shore lines as a consequence of a reduction of the waves emitted by the ship.
A considerable resistance component stems from the steady ship wave system. This effect amplifies 1
as the ship speed increases. Fast displacement type ships consume 50% and above of their installed
power to overcome the wave resistance. It therefore is mandatory to reduce wave-making as much
as possible. In order to do so early in the design process a focused hull form optimization approach
has been developed and implemented in terms of a MATHEMATICA (Wolfram (2003)) based optimization environment. It controls the optimization process flow, it comprises a sensitivity analysis, the
optimization kernel and a longitudinal wave cut analysis (WCA) scheme, and it establishes the connection to the CFD (Computational Fluid Dynamics) system SHIPFLOW. The entire optimization process
is fully automated. Nevertheless, user interaction is supported if need be.

2.1 Motivation and goals


An extensively studied field of ship hydrodynamics is the determination of the wave pattern and the
wave resistance of a ship moving with constant speed through otherwise calm water. This is explained
by the importance of the wave flow for the design of the entire ship hull. Hence, hydrodynamic ship hull
optimization has been challenging for numerous researchers since decades. A comprehensive compilation of literature on geometric ship hull modelling and CFD based hull optimization since the outgoing
19th century until the state of the art in 1998 is given by Harries (1998). The reader is referred to this
publication for additional references. Milestones in the development of wave resistance optimization
methods are in chronological order: 2 Michell (1898), Taylor (1915), Weinblum et al. (1957), Inui
(1962), Pien and Moore (1963), Lin et al. (1963), Wigley (1963) (the collected papers of Sir Thomas
1

This holds true for displacement type ships; planning or semi-planning hulls experience a resistance reduction after transition to the planning condition.
2 this list neither claims to be complete nor representative

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

11

2 Hydrodynamic optimization approach


Method

Flow
analysis

Objective
function

Evaluation of
objective function

Free
variables (N)

Constraints

Experimental
study

EFD

Residuary
resistance,
wave height

Froudes hypothesis
(k-factor method),
wave probes testing

Form parameters,
max. dimensions,
stability...

Interactive /
Automatic
optimization

CFD
inviscid

Wave
(pattern)
resistance

Hull pressure
integration, wave
pattern analysis

Distinctive
hull features
(N 101 )

Present
automatic
optimization

CFD
inviscid

Wave
pattern
resistance

Wave pattern
analysis
(wave spectrum)

Flow related
quantities
(N > 103 )

- dito (at dynamic


floating position)

Hull shape
parameters
(N 102 )

- dito -

Tab. 2.1: Wave resistance optimization techniques.


Havelock on hydrodynamics), Sharma (1966), Kracht (1978), Hsiung (1981), Janson and Larsson
(1996), Nowacki (1997), Sding (1997), and Harries (1998).
Due to the availability of fast and reliable CFD codes, in particular free surface Rankine panel codes,
advanced CASHD (Computer Aided Ship Hull Design) tools for the geometric hull modelling and
a wide range of suitable optimization methods, hydrodynamic ship hull optimization has become
a rapidly developing field of both research and practical application. Tools range from fully automatic implementations over semi-interactive environments to highly interactive optimization processes.
The latter are still favoured by many shipyards and model basins and often involve EFD (Experimental Fluid Dynamics). Recent CFD based wave resistance optimizations are reported, e.g. by
Baskaya (1997), Hirayama et al. (1998), Huan and Huang (1998), Huang et al. (1998), Harries and Abt
(1999), Birk and Harries (2000), Harries et al. (2001), Hendrix et al. (2001), Percival et al. (2001),
Peri et al. (2001), Sding (2001a), Tuck et al. (2002), Birk and Harries (2003), Heimann and Harries
(2003), Maisonneuve et al. (2003), Valdenazzi et al. (2003), Valorani et al. (2003), and Grigoropoulos
(2004).
The formation of waves generated by a ship moving with constant speed through unrestricted calm
water is determined by its hull shape and speed. 3 The design speed, normally, is fixed very early by
the prospective shipowner and/or the operator. A fleet, service and/or route optimization might have
been conducted to determine an appropriate service speed. Hence, a minimum requirement to ship
design is that the ship reaches the demanded service speed. However, more critical is the requirement
of an optimal hydrodynamic performance at (or around) the service speed. In terms of resistance and
power consumption the latter requirement is accomplished solely by a favourable hull design.
The present optimization approach aims at an improvement of the wave-making characteristics, i.e.,
the minimization of the energy losses associated with the wave formation along the hull which can be
measured in terms of the free wave spectrum and the wave pattern resistance. This improves the system
identification by tracing hull variations up to their respective fingerprints in the spectral distribution
and in the wave pattern resistance. Hull variations are driven directly by hydrodynamics, i.e., the hull
shape is controlled via a selected flow quantity along the hull. The motivation of addressing the wave
resistance problem by an optimization approach which focuses on wave cut analysis and, hence, on the
3

disregarding the influence of different loading conditions

12

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

2.2 Cause and effect chain


free wave spectrum and the wave pattern resistance may be best given by citation of selected historical
references:
"[...] The above important equation [editorially: integral equation for the wave pattern resistance determined from the free wave spectrum] not only gives a method of calculating wave-making resistance
but also clearly indicates that the wave-making resistance can be reduced only by reducing the amplitude of each individual elementary wave. Hence, the main effort of our research work in this field is to
derive a method of reducing these amplitudes. [...]"
Pien and Moore (1963)
"[...] Since the free-wave spectrum suffices to describe the dominant part of the entire wave field,
we can approximately predict the wave flow in a major part of the fluid, once we have measured it
at a few places. Further, it should be noted that within the linearized approach [...] the free-wave
spectrum is a linear function of the generating source distribution whereas the wave resistance is not.
Hence we can predict the wave-making properties of certain linear combinations of systems which have
been analyzed individually. For example, if we only know the wave resistance of a certain hull form,
we cannot estimate rationally the wave-resistance of a catamaran, consisting of a pair of such hulls.
However, if we know the free-wave spectrum of the single hull, we can reasonably predict the wave
resistance of any parallel configuration of two or more such hulls in the horizontal plane. Evidently,
this principle has potential applications to design work. [...]"
Sharma in Eggers et al. (1967)
"[...] The use of the wave-resistance integral [editorially: integral equation for the wave pattern resistance determined from the free wave spectrum] with experimental measurements of the amplitude
function A() to determine the total wave resistance provides a direct measurement of the wave resistance without recourse to Froudes hypothesis. This approach is known as wave pattern analysis.
A significant feature of this technique is that the measured quantity A() is linear in the ships disturbance, as compared to the quadratic wave resistance. By relating changes in the wave amplitude
to changes in hull form, a linear optimization of the hull shape can be achieved more directly than is
possible by studying the effects of shape on the total wave resistance. [...]"
Newman (1977)
The present optimization approach substantially differs from other approaches to wave resistance minimization in the way it tackles this objective. Table 2.1 gives an outline of the present optimization
approach in comparison to EFD and other CFD based wave resistance optimization techniques. Differences to related methods comprise the
Objective function: Wave pattern resistance (free wave spectrum)
Evaluation of the objective function: Wave pattern analysis of the computed ship wave pattern
Free variables: Flow related quantities along the hull surface (N > 103 )
Scope: Optimization of confined hull regions or the entire hull in its dynamic floating position

2.2 Cause and effect chain


The cause and effect chain from the variation of an optimization variable up to its impact in the wave
spectrum and the wave pattern resistance is, in its proper sense, of complicated and nonlinear nature.

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

13

2 Hydrodynamic optimization approach


In the proposed method the whole chain is simplified, i.e., linearized by linear perturbation of the
boundary value problem around its actual state. The boundary value problem is that of potential theory dealing with the nonlinear free surface conditions. By application of the perturbation approach
the nonlinear relations are transferred to explicit equations depending in a linear and straightforward
manner on the optimization variables.
As a consequence, the whole cause and effect chain starting with the variation of the optimization
variables, resulting in a hull form variation, affecting the flow quantities like the free surface velocities
and the wave formation, and finally bringing about a variation of the free wave spectrum and the wave
pattern resistance, as the latter determined by longitudinal wave cut analysis, can be consistently traced.
This opens the track to an improved system identification and optimization since hull variations can be
explicitly traced all the way to their respective fingerprints in the spectral distribution and in the wave
pattern resistance. In the present scheme the cause and effect relations can be simply represented by
sensitivity terms or gradients.

2.3 Process flow


The optimization process flow is shown in figure 2.1. A full sub-optimization task is performed by
processing the flow chart once from the initial quality assessment down to the improvement assessment.
Normally, a series of sub-optimization tasks is conducted by progressively reentering the loop.
The flow simulations are conducted by means of the nonlinear free surface potential flow module of
the CFD system SHIPFLOW. SHIPFLOW is a Rankine panel solver requiring a panelization of the
hull and the free surface. On the basis of the waves computed with SHIPFLOW a detailed performance
assessment is carried out in terms of the free wave spectrum and the wave pattern resistance. A longitudinal wave cut analysis utilizing a novel implementation of the longitudinal wave cut method in the
code SWASH, see chapter 3, is undertaken.
Within the optimization environment five stages are highly integrated setting up a complete suboptimization task:
Initial quality assessment
Initialization
Sensitivity analysis
Hull improvement
Improvement assessment
Initial quality assessment

The optimization process commences with the initial 4 quality assessment. At this stage the process
control parameters and the hull geometry enter the process. Control parameters concern the global
process flow, the perturbation and sensitivity analysis, the free surface flow simulation, the wave cut
analysis and the set-up and solution of the minimization problem. The hull geometry is given in terms
4

or an intermediate quality assessment when the process progresses

14

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

Parent / interim
hull form

Free surface flow simulation


(CFD)

Control parameter set


(process flow)

Detailed shape evaluation

Performance assessment
(WCA)

Modification of
control parameter set
(Restart)

Initialization

(Re-)Initialization of
optimization task
(variables, constr., bounds)

Systematic hull
perturbation analysis

Transfer matrices

Initial quality
assessment

2.3 Process flow

Systematic variation of
design variables
(flow quantity related)

Sensitivity analysis

Impact on hull form


(inverse perturb. solution)

Impact on flow field


(wave elevat., flow quant.)

Impact on
wave energy distribution
(free wave spectrum, RWP)
Evaluation of gradients
(MoM, constraints, bounds)

Minimization of objective
function implying constraints
(iterative scheme)
Generation of
improved hull form &
prediction of gains / changes

Improved hull form

Free surface flow simulation


(CFD)

Detailed shape evaluation &


further processing

Performance assessment
(WCA)

Automatic adaptation of
optimization task

Stop

Improvement assessment Hull improvement

Gradients matrices

Fig. 2.1: Optimization process flow.

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

15

2 Hydrodynamic optimization approach


of an offset or panel input point mesh. The offset geometry is used to perform an initial nonlinear
free surface CFD computation with free trim and sinkage which is forced to firm convergence. The
computed free surface elevations and the panel mesh geometry are then passed to a longitudinal wave
cut method which provides the wave spectrum and the wave pattern resistance. 5 Details are given in
chapter 3. The flow solution from SHIPFLOW along with the wave spectrum and the wave pattern
resistance from the wave cut analysis using SWASH serve as the base point (base solution) for the
sub-optimization task.
Initialization

At the second stage the optimization variables (flow related quantities in terms of the hull panel source
densities) are set according to the control parameters. The constraints (bounds) are determined. A
systematic hull perturbation analysis is conducted which correlates the optimization variables, the (discretized) hull geometry, the flow velocities, the wave formation, the wave spectrum and the wave
pattern resistance. This yields transfer matrices which serve as the basic instrument for the sensitivity
analysis. Details are given in chapter 4.
Sensitivity analysis

Once the transfer matrices are assembled the systematic variation of the optimization variables (i.e., the
hull panel source densities) starts. The variation of a single optimization variable has an impact on the
hull form, on the flow field, on the wave formation and, finally, on the free wave spectrum and the wave
pattern resistance. All these influences are determined within the present environment for all variables
(by parallel solving) and are compiled in terms of gradients matrices for the objective function and the
constraints (bounds). Details are given in chapter 4.
Hull improvement

The gradients matrices are directly utilized by the hull improvement instance. The optimum the
minimum of the objective function or measure of merit (MoM) is determined within the limits of
the sub-optimization task. The governing linear equation system is repeatedly solved for the seeked
optimization variables in terms of an iteration scheme which is used to isolate the active from the
passive constraints. Finally, an improved hull form is derived applying the gradients matrices. The hull
form is obtained in terms of a relocated hull panel input point mesh which possesses the same topology
as the initial mesh. Details are given in chapter 5.
Improvement assessment

The final stage of the sub-optimization task serves as the base point (base solution) for the successive
sub-optimization loop. The improved hull geometry is sent again to SHIPFLOW to compute the fully
nonlinear free surface flow. The longitudinal wave cut method then provides the wave spectrum and
the wave pattern resistance.
5

The evaluation of the free surface elevations and their bidirectional interpolation which yields the longitudinal wave cut is
performed outside SHIPFLOW by the wave cut method.

16

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

2.3 Process flow


The optimization process is terminated by three alternative stopping criteria. Details are given in
section 6.1. If none of the stopping criteria apply the optimization enters the next sub-optimization
loop. The optimization process is fully automated requiring no user interaction. However, the user
may manually interrupt the process at certain points. This allows, for instance, to restart from any
previous solution with a modified or adapted set of control parameters.
Formally, each sub-optimization task is composed of a prediction and a correction step. The prediction
step, essentially, comprises the optimization kernel, i.e., the initialization, the sensitivity analysis and
the hull improvement stage. In the correction step the predicted improvements are assessed by a full
flow and wave cut analysis. Here, the predicted optimization results are corrected for nonlinearities, for
the new dynamic floating position and for effects caused by the reconstruction of a coherent hull surface. The solution of the correction step then serves as the base point (base solution) for the subsequent
sub-optimization task.

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

17

2 Hydrodynamic optimization approach

18

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

3 Hydrodynamic analysis
The hydrodynamic flow analysis comprises the evaluation of the nonlinear free surface wave
flow by a CFD simulation and the analysis of the energy components contained within the
wave pattern. The first task is accomplished by the free surface potential flow module XPAN
of the CFD system SHIPFLOW. The second task is performed by longitudinal wave cut analysis utilizing a new tool called SWASH, developed by the author. The potential theory based
CFD simulation as well as the wave cut analysis method are presented in this chapter.
An extensively studied field of ship hydrodynamics is the determination of the wave pattern and the
wave resistance of a ship moving with constant speed through otherwise calm water. This is explained
by the importance of the wave flow for the design of the entire ship hull. Nowadays, advanced, reliable
and fast CFD tools are available for the numerical evaluation of the free surface wave flow. Methods
based on inviscid potential theory are widespread and are the first choice in practical wave resistance
computations for a wide range of ship types and flow cases. Moreover, potential flow methods, or
rather free surface Rankine panel methods, are extremely useful in the CFD based wave resistance
optimization, for they are comparatively straightforward in handling, they are fast and they generate
reliable results in terms of detailed flow information for a variety of applications.
Most of the CFD codes utilize integral resistance values derived from hull pressure integration. In this
way, however, information is lost on where beneficial and adverse effects originate. A more detailed
examination is attainable on the basis of the spectral distribution of wave energy along the components
of the steady ship wave system. Hence, the wave pattern evaluated by means of a free surface Rankine
panel method are thoroughly analysed by a longitudinal wave cut analysis. In the present context wave
cut analysis is used both in the sense of system identification and to provide the objective function of
optimization. Wave cut analysis yields the wave pattern resistance RW P while by convention an
integration of the longitudinal pressure components over the hull provides the wave resistance RW .

3.1 CFD simulation


The steady ship wave system is determined by means of a CFD simulation. In the present work the
nonlinear free surface Rankine panel module of the SHIPFLOW system is applied, see section 3.1.4.
The boundary value problem of potential theory is solved numerically. Boundary value problems are
addressed by boundary element methods (BEM) which require the discretization of all boundaries of
the flow domain, here the hull and the wavy free surface. The boundaries are discretized either by
first order flat or higher order panels. Potential functions are introduced by distributing singularities
(usually sources and/or dipoles) along each panel. The boundary conditions on the hull or the wavy
free surface are enforced for each panel at its control point. The solution of the boundary value problem
leads to a linear(ized) system of equations for the unknown panel source densities. Once the source

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

19

3 Hydrodynamic analysis
densities are known the velocities and the pressure forces are evaluated. Either linear or nonlinear free
surface conditions are applied. Nonlinear methods have been found superior to linear methods with
respect to predicting the wave amplitudes, the phases and the wave resistance, see Jensen (1988), Raven
(1996) and Janson (1997). For a numerical treatment the nonlinear free surface condition is linearized
around a known base solution. Since the free surface condition is to be applied at the a priori unknown
free surface an iterative solution scheme is used. In the iteration scheme the flow quantities and the
free surface position are updated in an alternating manner. The dynamic floating position is iteratively
updated by a force and momentum balance.
At present it is not possible to compute the ship resistance by CFD methods, whether viscous or inviscid, to the same accuracy as in a towing tank test by means of EFD (Experimental Fluid Dynamics).
CFD tools still cannot compete with EFD when it comes to power prediction. However, even if CFD
cannot provide the absolute power consumption to a high level of accuracy, CFD methods are well
applicable in comparing different hull shapes or variants in terms of their hydrodynamic performance.
This particularly holds true for nonlinear free surface Rankine panel solvers. Compared to EFD, application of CFD is
fast,
relatively inexpensive,
not sensitive to measurement uncertainties,
reproducible and
able to provide detailed flow information.
These properties are especially valuable for CFD based hull optimization. Hence, nonlinear free surface
Rankine panel solvers are frequently used by ship model basins, shipyards and consultants for manual
and sometimes semi-automated or even fully automated hull shape improvement.
Solvers that fall in the category of nonlinear free surface Rankine panel methods and which have
been successfully applied in ship design and optimization are, for instance, KELVIN by Sding
(2000), RAPID by MARIN, see Raven (1996) and Raven (2004), SHALLO/-SHALLO by HSVA,
see Jensen (1988) and Marzi (2004) and the potential flow module XPAN of the SHIPFLOW system
by FLOWTECH, see section 3.1.4.

3.1.1 Boundary value problem

The boundary value problem of potential free surface flow is thoroughly discussed and documented in
the literature. For instance see Dawson (1977), Ni (1987), Jensen (1988), Jensen et al. (1989), Nowacki
(1995), Raven (1996) and Janson (1997) for literature on this topic. In this section only an outline of
the boundary value problem is given. For a detailed discussion the reader is referred to the given
references.
Reference coordinate system and normalization

The flow is described in an Eulerian sense, i.e., the reference coordinate system is fixed to the ship
having the same speed but does not follow its dynamic trim and sinkage. A right-handed Cartesian

20

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

3.1 CFD simulation


system is defined with the origin located at the fore perpendicular in the undisturbed free surface level,
x pointing downstream and z pointing upward, see figure 3.2.
~ | and/or the length
All quantities are used in non-dimensional form, normalized by the ship speed |U
between perpendiculars LPP .
Laplace equation

The governing equation of potential free surface flow is the Laplace equation which is a linear, homogeneous partial differential equation derived from mass continuity
~2 = 0 .

(3.1)

Velocity potential

Assuming an inviscid (ideal) incompressible fluid with irrotational flow, the velocity components can
be determined from the gradient of a scalar potential function
~ = |U
~ | ~ .
U

(3.2)

The total velocity potential comprises the potential of the undisturbed onset flow and the potential

for the disturbance velocity caused by the ship

= + .

(3.3)

The velocity potential has to satisfy the Laplace equation (3.1) and it has to comply with the boundary
conditions.
Kinematic and dynamic hull boundary conditions

For a well posed boundary value problem boundary conditions need to be defined. The kinematic hull
boundary condition implies that the flow velocity must have a known component in the hull normal
direction 1

=F ,
n

(3.4)

where, usually F = 0, since no fluid particle is supposed to pass through the surface of a rigid body. In
terms of the disturbance potential the kinematic hull boundary condition reads

b ~n + F ,
= U
n

(3.5)

b = U
~ /|U
~ | indicates the unit velocity vector of the undisturbed onset flow. An additional
where U
dynamic hull boundary condition may be imposed, claiming equilibrium between the hydrodynamic
and hydrostatic pressure forces on the wetted hull portion and the ships weight distribution. This
integral condition primarily determines the dynamic trim and sinkage.
1

which is directed into the fluid

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

21

3 Hydrodynamic analysis
Kinematic and dynamic free surface boundary conditions

At the free surface two more boundary conditions are applied, both of which need to be fulfilled at the
initially unknown wavy free surface. The kinematic free surface condition implies that the flow must
be tangential to the free surface, i.e., no fluid particle is supposed to leave the surface = f (x, y)

+

=0,
x x y y z

(3.6)

where refers to the distance (normalized) of the free surface from the undisturbed initial plane.
therefore describes the wave elevation (normalized). The second condition, the dynamic free surface
boundary condition, states that the static pressure, expressed through Bernoullis equation, must be
constant (atmospheric) at the free surface
#
"     

1
2
2
2
+
+
+
1 = 0 .
(3.7)
Fn2 2
x
y
z
The kinematic (3.6) and dynamic (3.7) free surface boundary conditions are given in normalized form.
Fn in (3.7) indicates the Froude number as defined by
VS
Fn =
.
gL
where VS the ship speed and g the acceleration due to gravity. Here the characteristic length L is taken
as the length between perpendiculars LPP .
Linearization of the free surface boundary condition

The free surface problem is nonlinear since the free surface boundary conditions, (3.6) and (3.7), are
nonlinear and are to be imposed at the initially unknown free surface. Hence, the flow quantities depend
nonlinearly on the location of the free surface. Solution methods for the fully nonlinear free surface
problem have been proposed since the 1980s. Present methods linearize the free surface boundary
conditions around a known base solution and solve the problem in an iterative manner. In the iteration scheme the flow quantities and the free surface elevation are updated in alternating steps, see
section 3.1.3. A common approach is to linearize the free surface boundary conditions in a first order
Taylor series expansion around the known base solution and to introduce perturbations due to the hull
and/or the waves. Thus, higher order contributions are assumed to be small. Reviews of different
linearizations are presented, e.g., in Newman (1976) and Raven (1996, 1997).
The majority of the presently available nonlinear free surface Rankine panel solvers adopt the so called
Dawson linearization of the free surface boundary conditions, as originally suggested by Dawson
(1977). In Dawsons method the linearization is conducted around the so called double-body flow 2
as base solution. The kinematic and dynamic boundary conditions are linearized in a perturbation
sense. The perturbation is due to the waves. The free surface boundary conditions, actually, are to be
satisfied at the initially unknown free surface. But, so as to allow a direct solution the conditions are
transferred to and solved at the known free surface. Consistency would require to incorporate transfer
terms to account for this transfer of the boundary conditions. Dawsons linearization is known to be an
2

In a double-body flow the water surface is treated as a symmetry plane at which the underwater portion of the hull is
mirrored.

22

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

3.1 CFD simulation


inconsistent linearization since it neglects these transfer terms. However, practical applications have
shown that Dawsons method performs equally well or even more stable than other more consistent
linearizations, see Raven (1996) and Janson (1997) for a thorough investigation. At convergence the
inconsistency of the linearization is no longer relevant.
The free surface elevation and the velocity potential are expanded to first order in Taylor series
around the known base solution . The expansions are then introduced to the free surface boundary
conditions neglecting higher order contributions in and or mixed higher order terms. Finally, the
linearized kinematic and dynamic boundary conditions are combined to give a formulation in known
and unknown velocities and their derivatives. The linearization of the free surface boundary conditions
is described in detail by Janson (1997).
Once the new velocity potential is determined the free surface elevation is obtained from the linearized free surface boundary condition 3
"
 2  2  2

#
1 2


(x, y) = Fn 1 +
+
+
2
+
+
, (3.8)
2
x
y
z
x x y y z z
with being the velocity potential of the known base solution.
Radiation and transom conditions

Further to the above conditions a radiation condition is required which implies that the disturbance
velocity due to the presence of the hull vanishes as the distance from the hull r approaches infinity

lim |~ | = 0 .

(3.9)

In the steady state solution a further radiation condition is needed to prevent waves from propagating
upstream of the hull. However, this condition is not formulated explicitly in Rankine panel methods.
It is enforced implicitly by the numerical method, for instance by applying an upstream differencing
scheme and/or an upstream shift of the free surface collocation (control) points in conjunction with the
so called raised panel method. 4
If a ship possesses a transom stern a special transom condition is required to model the flow separation
at the transom edge (according to the Kutta condition) and the flow behaviour aft of the transom stern.
Further boundary conditions are to be applied in case of restricted waters at the bottom and (if present)
at canal-walls. In inviscid flows typically a Neumann condition, i.e., a zero normal flow condition, is
applied at these boundaries.
Singularity distribution

Potential functions are introduced by distributing singularities on the boundaries of the flow domain,
essentially the hull and the wavy free surface. 5 The singularities (e.g., sources or dipoles) themselves
3

linear(ized) in the new velocity potential


In the SHIPFLOW code both an upstream differencing operator is applied longitudinally for the evaluation of the numerical
derivatives and, in addition, the free surface collocation points are shifted upstream relative to the corresponding free
surface panels, see Janson (1997) for details.
5 Unless the wave Green function is applied, known as Havelock singularity, which inherently fulfills the linearized free
surface boundary condition and only requires a singularity distribution on the hull.

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

23

3 Hydrodynamic analysis
already satisfy the Laplace equation (3.1), see e.g. Newman (1977) and Wendt (1996). Usually, source
singularities are distributed along the boundaries of the flow domain and are integrated to give the
disturbance potential
ZZ
q

dS ,
(3.10)
r pq
SD

where q is the source density per unit area at point q on the boundary surface SD and r pq is the distance
from point q where the source is located to an arbitrary point p where the potential is to be computed.
Introducing (3.10) to the kinematic hull boundary condition (3.5) yields the Fredholm integral equation
of the second kind which serves as an explicit relation for the unknown hull source densities. Introducing (3.10) to the combined, linearized free surface condition (if the free surface is present) yields
additional relations at the unknown free surface. A simultaneous solution of both relations then yields
the seeked a priori unknown source densities on the hull and the free surface.
3.1.2 Discretization

However, a solution of the boundary value problem in closed form for practical ship application cases
is infeasible. Hence, the boundary value problem is solved numerically. A numerical solution requires
the discretization of the boundaries of the flow domain, essentially the hull and a reasonable portion of
the wavy free surface. Panel methods discretize the boundaries by quadrilateral or triangular panels.
Panel methods for computing the potential flow around arbitrary three dimensional bodies were invented by Hess and Smith (1962). They introduced flat quadrilateral panels with a constant singularity
distribution to approximate the boundary value problem. In their notation the flat quadrilateral panels
are generated from so called input points which constitute a coherent point mesh along the hull. From
a topological point of view input points are equivalent to offset points which often are arranged along
hull sections and profile curves and are located directly on the hull surface. However, it is impossible
to connect the input points by flat quadrilateral panels along a curved hull surface without introducing
gaps between the panel edges. In the original Hess & Smith method the flat quadrilateral panels are
determined from the input points so that the panel corner points are at least a close approximation of the
guiding input points and that the panel normal direction is a good approximation of the actual surface
normal. The boundary conditions are enforced for each panel at a panel control point. 6 Later Hess
(1972, 1979) extended the method to higher order curved panels with a linear singularity distribution
across the panels and to lifting surfaces.
As a consequence of the discretization of the boundary value problem the integral equations are transformed to sum equations, which yield an explicit relation for the unknown panel source densities j :
NF S

j Ai j = Bi

j=1

i = 1, , NFS ,

(3.11)

where Ai j indicates the influence coefficient from panel j to the velocity at panel i, Bi is the inhomogeneous term at panel i. NFS accounts for the total number of panels, including the hull and free surface
6

In the original Hess & Smith method the so called null point is suggested where the panel itself induces no velocity in
its own plane. However, the determination of the null point requires the solution of additional simultaneous nonlinear
equations, so, usually, it is simply approximated by the centroid of the panel.

24

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

3.1 CFD simulation


panels. The influence coefficients Ai j are determined from the induced velocities 7 Xi j , Yi j , Zi j in the
respective directions and, in case of the free surface portion, from the velocities and velocity derivatives
of the base solution. Details are given in Janson (1997).
The free surface domain is discretized by so called Rankine panels. The linearized boundary condition
is satisfied in a discrete sense at the free surface panel collocation points. Accordingly, the free surface
elevation is evaluated in a discrete sense with the elevation at collocation point i given by virtue of the
linearized free surface boundary condition (3.8):
i =


1 2
Fn 1 + 2x i + 2y i + 2z i 2 (x i x i + y i y i + z i z i )
2

i = NH + 1, , NFS , (3.12)

with x = /x etc. and NH < i NFS for the free surface panels.
3.1.3 Numerical solution

The numerical solution of the nonlinear free surface boundary value problem is thoroughly described
by Jensen (1988), Raven (1996) and Janson (1997).
State of the art CFD systems make use of the so called raised panel method. The free surface panels are
raised by a certain 8 distance above the free surface collocation points possibly in conjunction with an
upstream collocation point shift, Jensen et al. (1986). In this way the radiation condition is implicitly
enforced and convergence is improved. As a further numerical feature the velocity derivatives at the
free surface are determined along longitudinal panel rows by an upstream differencing scheme and
transversally along rows with constant x by a central difference scheme, as originally suggested by
Dawson (1977).
Equation (3.11) provides a linear system of equations for the unknown hull and free surface panel
source densities. But, since the free surface condition is to be applied at the a priori unknown free
surface an iterative solution scheme is used where (3.11) is solved successively. In each iteration the
boundary value problem is linearized with respect to the solution of the previous iteration. The setup
is as follows:
1. Compute a base solution, e.g., by adopting a Neumann or a double-body condition
undisturbed free surface.

at the

2. Use the velocities from the previous step to compute the free surface elevation from the linearized
dynamic free surface condition (3.8) at the undisturbed free surface.
3. Move the free surface panels to the new free surface computed at the previous step.
4. Move the latest known flow solution to the adjusted free surface panelization of the previous
step, without being recomputed.
5. Solve the linearized boundary value problem according to (3.11) at the actual free surface.
7

Xi j , Yi j , Zi j are pure geometrical quantities and represent the induced portion from panel j to the velocity at panel i in the
x-, y- and z-direction respectively.
8 carefully selected
9 Applying the double-body condition yields the Bernoulli wave.

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

25

3 Hydrodynamic analysis

Fig. 3.1: Zonal approach of the CFD system SHIPFLOW (courtesy of FLOWTECH Int. AB.)
6. Apply the solution of the previous step to adjust the free surface elevation according to the
linearized dynamic free surface condition (3.8).
7. Continue with step 3 until convergence.
Firm convergence is requested for the maximum wave change, for the maximum change of the source
densities and, if allowed to freely adjust, for the trim and sinkage of the hull in successive iterations.
The dynamic floating position is updated via a force and momentum balance at each iteration.
The wave resistance is computed from the converged solution by integration of the longitudinal pressure components over the hull. For further details be referred to Janson (1997).
3.1.4 The CFD system SHIPFLOW

In the present work the CFD simulation of the steady ship wave system is conducted by means of the
free surface Rankine panel module XPAN of the CFD system SHIPFLOW 10 , see e.g. Larsson et al.
(1989, 1990, 2004a,b), Larsson (1993) and Janson (1997).
SHIPFLOW solves the boundary value problem of potential theory numerically. The boundaries of the
flow domain, essentially the hull and the wavy free surface, are discretized either by flat or, as in the
higher order method, by bi-parabolic quadrilateral panels. Source singularities 11 are distributed along
each panel either with a constant source density or, as in the higher order method, with a linear distribution of the source density along the panel. The boundary conditions on the hull and the wavy free
surface are for each panel enforced at a control point defined as the centroid of the panel. Raised panels
are utilized at the free surface in conjunction with an upstream collocation point shift. Tangential flow
separation at the edge of a transom stern (if present) is modelled by an additional free surface transom
group which extents downstream of the transom. So called extra hull panels above the free surface level
might be used in case of ships with transom sterns or complicated geometries to stabilize the convergence history. The solution of the boundary value problem leads to a linear system of equations for the
unknown panel source densities. Once the source densities are known, the velocities and the pressure
forces are evaluated. SHIPFLOW adopts either a linear or the nonlinear free surface condition. For a
10
11

FLOWTECH Int. AB, Gteborg, Sweden, http://www.flowtech.se


and dipoles on lifting surfaces and on the wake sheet

26

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

3.1 CFD simulation


numerical treatment the nonlinear free surface condition is linearized around a known base solution.
Since the free surface condition is to be applied at the a priori unknown free surface contour an iterative solution scheme is used. In the iteration scheme the flow quantities, i.e., the source densities and
velocities, and the free surface position are updated in an alternating manner, see section 3.1.3. The
dynamic floating position is iteratively updated by a force and momentum balance.
Prominent features of SHIPFLOW/XPAN are:
Wave resistance from pressure integration over the wetted hull surface
Wave pattern resistance from a transverse wave cut method 12
Linear and nonlinear wave pattern
Longitudinal and transverse wave cuts
Flow velocities
Pressure distribution
Dynamic trim and sinkage
Shallow water effects
Multihull option
Automated mesh generation
Propeller representation by an actuator disk model
Lifting surfaces
Induced resistance and lift
The SHIPFLOW code is specifically designed for steady ship hydrodynamics. In addition to the inviscid potential free surface flow, SHIPFLOW also offers modules to compute boundary layers, viscous
flow effects and drag. The way SHIPFLOW handles these unique problems is by splitting the flow
into three zones, which allows an efficient simulation of the governing flow phenomena. The zonal
approach is outlined in figure 3.1. A full investigation comprises the following three stages which are
processed successively:
Zone 1 Potential flow calculation. Rankine panel method with linear or nonlinear free surface bound-

ary conditions and adjustable dynamic floating position.


Zone 2 Boundary layer calculation. Momentum integral method for laminar and turbulent boundary

layers (computation of transition point).


Zone 3 RANSE calculation. Finite difference based solution of the Reynolds-Average Navier-Stokes

Equations (RANSE) applying a k- turbulence model on a cylindrical structured grid.


12

SHIPFLOW Release 2.8 or higher

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

27

3 Hydrodynamic analysis
SHIPFLOW is robust in application and extensively validated. It comes with an automated mesh
generation utility, it provides a variety of control mechanism and I/O capabilities to the user, supplying
detailed results. SHIPFLOW Release 2.8.10 and higher provides additional, specifically customized 13
interface capabilities like the I/O of the panel geometry, of selected influence coefficients and certain
mechanism for the SHIPFLOW execution control. Moreover, the SHIPFLOW system is widely used
by ship model basins, shipyards and consultants both for hydrodynamic analysis in the preliminary
design stage and for hull shape improvement.
The SHIPFLOW system was therefore selected for the CFD computations in the present work.
3.1.5 Application aspects

The inherent neglect of the viscosity by potential flow methods implies that wave-viscous interactions
cannot be captured. A thorough investigation, both numerically and experimentally, of wave-viscous
interactions including propeller effects is reported by Nowacki and Sharma (1972). Wave-viscous interactions occur mainly in the stern region where wave amplitudes tend to be overpredicted by CFD.
However, the differences are small for the ship types and speed range particularly addressed by the
present optimization approach: relatively fast and slender hulls like ferries, RoPax, container vessels
etc.. Potential flow codes also assume the boundary layer to be sufficiently thin. This is a valid assumption for the forebody of a displacement ship and for the entire hull of a fast ship where the boundary
layer detaches the transom edge without considerable thickening. CFD validation studies comparing
computed with measured wave pattern showed the excellent ability of modern nonlinear free surface
Rankine panel methods to model the steady wave systems of the ship types and speed range 14 relevant
for the present work, see e.g. Raven (1996), Janson (1997), Raven and Prins (1998a,b), Nowacki et al.
(1999), Heimann (2000), Valdenazzi et al. (2003), Heimann and Harries (2003). Furthermore, Rankine
panel methods proved to be sensitive even to small and locally confined hull form changes.
A well known and extensively studied feature of CFD methods, whether viscous or inviscid, is their
dependence upon the grid or mesh on which the discretized flow equations are solved numerically.
This property also applies to Rankine panel methods. Further uncertainties are introduced by the
raised panel method. These effects are attributed to numerical dispersion (wavelength error) and in
particular to numerical damping (amplitude error). Within the scope of the SHIPFLOW solver griddependence and CFD validation studies were reported, e.g., by Janson (1997), Harries and Schulze
(1997), Schulze and Harries (1997) and Lee (2003). Janson (1997) provides a valuable guideline for
the panel mesh generation and the selection of certain control parameters.
Raven (1996) and Janson (1997) report comprehensive studies to the application of the nonlinear versus
the linear free surface conditions. In general, nonlinear free surface Rankine panel codes have been
found superior in predicting the wave amplitudes, the phases and the wave resistance.

3.2 Wave cut analysis


The steady ship wave system, as determined by means of a CFD simulation, is further assessed by a
wave cut analysis. In the present scheme the longitudinal wave cut method is applied by means of
13
14

Professor Dr. Carl-Erik Janson, Chalmers University of Technology and FLOWTECH Int. AB, Gteborg, Sweden
0.2 Fn 0.4 (potentially higher Froude numbers)

28

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

3.2 Wave cut analysis

Q=0

Q>0

z, z

x
yWC

VS

longitudinal wave cut

xE

Fig. 3.2: Schematic sketch of the ship wave system (only one symmetric half is considered).
the MATHEMATICA based tool called SWASH (Ship Waves Analysis Light) developed by the author.
SWASH is an implementation of the wave cut method following Sharma (1963, 1966) and Eggers et al.
(1967).
Wave cut analysis (WCA) in the sense of system identification has shown to be a powerful tool
in CFD based optimization of the wave-making characteristics of ships, e.g. Heimann (2000) and
Heimann and Harries (2003). Most of the CFD codes utilize integral resistance values derived from
pressure integration. In this way, however, information is lost on where beneficial and adverse effects
originate. A more detailed examination is attainable on the basis of the spectral distribution of wave
energy along the components of the steady ship wave system by utilizing the so called free wave spectrum. This improves the system identification, i.e., the energy loss associated with the transverse and
diverging waves can be studied more clearly. Furthermore, the impact of local hull variations on the
individual components of wave formation can be traced by investigating the changes in the wave spectrum with respect to the associated hull variations. This then allows a reduction of particular adverse
wave components. However, the primary aim of the present optimization approach is to minimize the
total energy losses associated with the wave formation along the hull which can be measured in terms
of the wave pattern resistance RW P as determined by the wave cut analysis.
As an additional benefit computed wave pattern and hence wave cut analysis based on them prove
to be less sensitive to the flow field discretization than hull pressure integration. This is a tentative

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

29

3 Hydrodynamic analysis
result of related studies concerning that topic, e.g. Janson (1997), Raven and Prins (1998a,b) and
Heimann and Harries (2003).
3.2.1 Boundary value problem

The boundary value problem of the steady linear free surface potential flow is thoroughly discussed and
documented in the literature, see Michell (1898), Lord Kelvin (1906), Wehausen and Laitone (1960),
Eggers (1962), Inui (1962), Newman (1963), Wigley (1963) (the collected papers of Sir Thomas Havelock on hydrodynamics) and Sharma (1965) for a condensed selection of literature on this topic. In this
section only a brief outline of the problem is given. For an introduction see also Newman (1977).
Reference coordinate system and normalization

The flow is described in an Eulerian sense, i.e., the reference coordinate system is fixed to the ship.
A right-handed Cartesian system is defined with the origin located at the fore perpendicular in the
undisturbed free surface level, x pointing downstream and z pointing upward, see figure 3.2. Infinite
water depth is assumed, i.e., the water depth h is greater than half the fundamental wavelength 0 :
h > 0 /2.
All quantities are used in non-dimensional form. Length related measures are normalized by the basic
wavenumber
k0 =

g
,
VS2

(3.13)

where g is the acceleration due to gravity and VS the ship speed. VS is assumed constant. In other
words: The problem described is of steady-state nature and all dependence on time is blended out.
The wave pattern resistance RW P is obtained in non-dimensional form by
k03
,
g

RW P = RWP

(3.14)

where the bold expression RWP stands for the dimensional resistance component and is the water
density.
Dispersion relation, wavelength and wavenumbers

Prior to the further derivations some important quantities should be defined. The dispersion relation
for infinite water depth connects the wavenumber k, the wavelength of the plane progressive wave
moving at an oblique angle with respect to the x-axis and the speed of the steady wave system. The
dispersion relation is given in non-dimensional form (normalized by k0 ) by
k() =

2
= sec2 () .
()

(3.15)

From (3.15) the wavelength directly follows as


() =

30

2
k()

(3.16)

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

3.2 Wave cut analysis


and the wavenumber k is defined according to (3.15) in non-dimensional form (normalized by k0 ) by
k() = sec2 () .

(3.17)

Based on (3.17) the so called longitudinal


s() = k() cos() = sec()

(3.18)

and transverse wavenumber


u() = k() sin() = sec() tan()

(3.19)

are derived. The component wave direction , the longitudinal s and transverse u wavenumbers are
defined within the integration bounds 15
0 /2 ,

1s<,

0u<.

, s and u are fully convertible into one another, and so are the relations based on them. 16
Linearized free surface boundary condition

A non-viscous (ideal) incompressible fluid with irrotational flow is assumed. The kinematic (3.6) and
dynamic (3.7) free surface boundary conditions are linearized around the undisturbed free surface level
with respect to the undisturbed onset flow. Their combination yields the well known Neumann-Kelvin
boundary condition which is given in normalized form by
2
x

=0.
z

(3.20)

Far field velocity potential

A velocity potential which satisfies the Laplace equation (3.1) and the linearized free surface boundary
condition (3.20) can be derived following Havelock (1934a) at a large distance from the disturbance
(the ship) as

Z/2
=

e(z sec

())

cos() [AS () cos(s x + u y) + AC () sin(s x + u y)] d ,

(3.21)

/2

where AS and AC denote the sine (odd) and cosine (even) component, respectively, of the so called free
wave spectrum. The velocity potential (3.21) satisfies the linearized boundary condition at the free
surface in the far field of the ship.
For modelling the whole wave system, including the local, bounded waves in the vicinity of the hull,
a special form of the velocity potential can be utilized which is made up by another Greens function,
as given in Havelock (1932). This inherently satisfies the linearized free surface boundary condition
(3.20) in the whole flow domain, see e.g. Eggers et al. (1967) for details.
The wave pattern resistance can be determined from the far field waves represented by the velocity
potential (3.21) alone, as shown below.
15
16

for ship wave systems symmetrical to the midships plane


The lower integration bound of s = 1 conforms by virtue of (3.18) with = 0 , i.e., no waves are longer than those
determined by the fundamental wavelength.

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

31

3 Hydrodynamic analysis
3.2.2 Longitudinal wave cut analysis

The present optimization approach aims at the minimization of the wave pattern resistance which is
determined by a longitudinal wave cut analysis. In case of flows symmetric to the ship center plane, a
single infinitely long wave cut parallel to the ship suffices to perform the whole analysis.
The longitudinal wave cut method is preferred to other wave analysis methods since it is comparatively fast, relatively robust in application and it has a broad application range. This points are further
discussed in section 3.2.4.
The used notation and axes orientation follows to a large extent Sharma (1963). Complementary material is selected from Sharma (1966) and Eggers et al. (1967).
Free wave system

Out of the double infinite manifold of all possible plane progressive waves, varying both in wavenumber k and in component wave direction , it is only a single infinite set of so called free waves which
contribute to the wave pattern resistance, see e.g. Eggers et al. (1967). These free waves satisfy condition (3.15), i.e., wavenumber and wavelength of each component of the free wave system are determined one-to-one by its wave direction. Furthermore, outside a confined region around the disturbance
by the ship it is the free waves which constitute the wave pattern. Whereas, the local disturbance
arising from the presence of the nonfree waves vanish with the square of the reciprocal distance from
the disturbance.
Assuming a control surface located parallel to the course of the ship at infinite distance 17 to the ship,
the free wave system can be derived from the velocity potential (3.21) and from the Neumann-Kelvin
linearization of the dynamic free surface condition. The elevation of the free waves expressed in terms
of the component wave direction is given by

Z/2
(x, y) = [AS () sin(s x + u y) AC () cos(s x + u y)] d ,

(3.22)

or as a function of the longitudinal wavenumber s by

Z
(x, y) =
1

[AS () sin(s x + u y) AC () cos(s x + u y)]

d
ds .
ds

(3.23)

After rearranging in terms of s, relation (3.23) results in a form suitable for further treatment

Z
(x, y) =
1

{[AS () cos(u y) + AC () sin(u y)] sin(s x)+


[AS () sin(u y) AC () cos(u y)] cos(s x)}

17

d
ds .
ds

(3.24)

Here the control surface at y is adopted for practical reasons in order to conform with the SHIPFLOW internal
representation; an analogous derivation can be conducted for the control surface at y +, see Sharma (1963).

32

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

3.2 Wave cut analysis


Relations (3.22) to (3.24) can be interpreted as a superposition of elementary plane progressive waves
of all feasible wave directions and varying wave amplitudes in compliance with the steadiness condition
for the phase velocity, see relation (3.17) and compare e.g. Havelock (1934b). This results in the well
known Kelvin wave pattern which is confined in a bisector to the x-axis of 19 28 on deep water.
The waves follow the ship. The transverse wave component according to the basic wavenumber and
the basic wavelength travels in the ships advance direction, i.e., with = 0 . The transverse waves
possess transversally aligned wave crests and are composed of component waves up to = 35 16 .
Larger values of correspond to the so called diverging waves which have shorter wavelength and
converge toward the origin, compare e.g. Newman (1977).
Fourier transformation

The continuous Fourier integral is applied to transform between the free waves and the related free
wave spectrum, as it is represented further below. The Fourier transformation is given by
1
(x, y) =

Z
[S(s) sin(s x) +C(s) cos(s x)] ds ,

(3.25)

with the Fourier transforms

Z
(x, y) sin(s x) dx ,

S(s) =

(x, y) cos(s x) dx ,

C(s) =

(3.26)

where S and C denote the sine (odd) and cosine (even) components, respectively, of the Fourier transforms. In terms of the longitudinal wave cut analysis (x, y = const.) in the Fourier transforms (3.26)
represents a longitudinal wave cut either determined from a CFD simulation or from wave probe measurements in a towing tank. The longitudinal wave cut is given as wave elevations along varying
x-positions for a constant transverse cutting plane or offtrack position y = const..
Free wave spectrum

The Fourier transforms and the components of the free wave spectrum can be correlated by equating
(3.25) with the expression for the free wave pattern (3.24) followed by a separation of variables. By
virtue of the Fourier transforms (3.26) this yields

Z
S(s) =

(x, y) sin(s x) dx = [AS () cos(u y) + AC () sin(u y)]

d
,
ds

(x, y) cos(s x) dx = [AS () sin(u y) AC () cos(u y)]

d
.
ds

C(s) =

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

(3.27)

33

3 Hydrodynamic analysis
Resolving (3.27) for AS and AC finally results in
1
[S(s) cos(u y) +C(s) sin(u y)]

1
AC () = [S(s) sin(u y) C(s) cos(u y)]

AS () =

ds
,
d
ds
,
d

(3.28)

with AS and AC denoting the sine (odd) and cosine (even) components, respectively, of the free wave
spectrum. For the sake of simplicity AS and AC are referred to as free wave spectra from now on.
The notation of the free wave spectra in terms of A suggests a wave amplitude spectrum. Hence,
AS and AC are frequently called wave amplitude functions, e.g. Newman (1977). However, this terminology is slightly misleading, since AS and AC do not actually have the physical meaning of wave
amplitudes. In fact, the free wave spectra are better denoted as wave energy equivalents, i.e., terms
which express the energy losses associated with the formation of the individual components of the free
wave system. The wave spectra correlate with the real and imaginary part of the degenerated form of
the Kochin function. A detailed derivation and review is given by Eggers et al. (1967).
The free wave spectra are pure functions of the hull shape and the ship speed assuming infinite water depth. They yield a direct measure of the wave energy distribution along the components of the
ship wave system. This is why they are extremely valuable in ship hull optimization. In the present
optimization scheme the free wave spectra are directly utilized for decision making in terms of the objective function. However, the free wave spectra ought to be applied with some care since the Fourier
transforms and the free wave spectra are not invariant against coordinate transformations, like a shift
of the origin, whereas the integrand of the wave pattern resistance, equations (3.30) and (3.32) below,
is invariant against coordinate transformations.
Wave pattern resistance

Formally, the wave pattern resistance is deduced from energy considerations, i.e., the conservation of
momentum in a control volume which contains the steady ship wave system composed of a superposition of plane progressive waves of all feasible wave directions and varying wave amplitudes in
compliance with the steadiness condition for the phase velocity. The wave pattern resistance can be
evaluated directly from the Fourier transforms without reference to the free wave spectra which yields
(in non-dimensional form)
r
Z

1  2
1
1
RW P =
S (s) +C2 (s)
1 2 ds .
(3.29)

s
s
1

A different notation, going back to Havelock, adopts the free wave spectra and is given (in nondimensional form) by
RW P =

Z/2
0


A2S () + AC2 () cos3 () d .

(3.30)

Both expressions for the wave pattern resistance are equivalent and fully convertible into one another.
Expression (3.30) is favoured in the present work because the notation in terms of the free wave spectra and the component wave direction is more intuitive both for the optimization application (objective

34

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

3.2 Wave cut analysis


function) and for the interpretation of the results. For instance, application of (3.30) enables the designer to trace the cause and effect of hull variations and wave pattern changes. The factor cos3 () implies that the predominant portion of the wave pattern resistance originates from the transverse waves,
which explains why ship hull optimizations often focus on the transverse waves regime. However, due
to the positive definite nature of expression (3.30) it is impossible to fully neutralize the wave pattern
resistance, e.g., by favourable interference effects. 18
From the wave pattern resistance coefficient 19
RW P
CW P =
=
0.5 Sre f k02

Z/2
CW P () d ,

(3.31)

an extremely useful and descriptive chart, as will become evident in chapter 6, is obtained by plotting
the integrand of (3.31) as a function of the wave direction
CW P () =




AS ()2 + AC ()2 cos3 () ,


2
0.5 Sre f k0

(3.32)

with Sre f denoting the wetted hull surface area for normalization. Expression (3.32) yields a resistance
function which measures the wave energy distribution along the components of the ship wave system.
SWASH conducts the numerical integrations by means of the trapezoidal rule assuming a dense sampling which proved to be robust and straightforward.

3.2.3 Truncation correction

In practical applications due to limited computer resources (CFD) or restrictions due to tank wall reflections in experiments (EFD) longitudinal wave cuts need to be truncated at a finite distance downstream
of the hull. In order to retain the wave energy contained within the wave pattern downstream of the
truncation point SWASH treats wave cuts of finite length with a special truncation correction. At a sufficiently large distance aft of the stern the wave cut is usually dominated by transverse waves following
the ship. This gives rise to utilizing an analytic truncation function composed of harmonic waves decaying to infinity. A truncation correction based on an analytical asymptotic extension of the wave cut
is applied here.
Sharma (1966) and Eggers et al. (1967) suggested the following approach for an analytical asymptotic
extension of the wave cut infinitely downstream


e
(x, y)

xE+

c1 cos(x) c2 sin(x)

c3 + x

x + ,

(3.33)

where the coefficients c1 , c2 and c3 are determined by fitting (3.33) to the tail end of the wave cut
signal up to the truncation point xE . The unknown coefficients are determined by regression analysis. 20
Because of the strong asymptotic decay of both the wave elevation and the slope at x , according
to Sharma, practically no correction is needed to account for the truncation of the wave cut upstream.
18

except in the trivial case of a zero displacement ship


note the non-dimensional form of RW P in (3.14)
20 In the present implementation the Levenberg-Marquardt method is applied.
19

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

35

3 Hydrodynamic analysis
Sharma suggested to drop the coefficient c3 from (3.33) for the sake of simplicity if the wave cut signal
extends far downstream xE , see e.g. Eggers et al. (1967). Lalli et al. (1998, 2000) suggested to drop
c3 not only for the sake of simplicity, but also because it stabilized their results. Lalli systematically
applied longitudinal wave cut analysis to measured wave signals of a Series 60 model and a fast hardchine catamaran. As a consequence of dropping c3 the singularity of the correction function (3.33) is
inevitably fixed to the origin. Otherwise, retaining c3 in the regression analysis causes the singularity
to self-adjusting, promising a better fit due to the additional degree of freedom. The latter conforms
with the experiences gained by the author.
In order to evaluate the analytical asymptotic extension of the longitudinal wave cut weighted Fourier
transforms are introduced. The weighted Fourier transforms are given by virtue of (3.26) as

SW FT (s) =

ZxE p

s2 1

s2 1

ZxE

CW FT (s) =

(x, y) sin(s x) dx +

Z p

xE

(x, y) cos(s x) dx +

s2 1 e
(x, y) sin(s x) dx ,

Z p

xE

s2 1 e
(x, y) cos(s x) dx ,

(3.34)

where SW FT and CW FT denote the sine (odd) and cosine (even) component, respectively, of the
weighted Fourier transforms, the weight being s2 1. The respective first integrals over (, xE ]
in (3.34) are evaluated by numerical integration as before with the longitudinal wave cut (x, y) truncated at xE . The respective second integrals over [xE , ) in (3.34) are evaluated in closed analytical
form by means of the complementary Fresnel integrals, e.g. Press et al. (1992) and Bronstein et al.
(1993), which arise in the description of diffraction phenomena. The closed analytical evaluation is
elaborated in detail in Sharma (1966) and Eggers et al. (1967) and not repeated here.
Introducing the weighted Fourier transforms (3.34) to the free wave spectra (3.28) yields
s
[SW FT (s) cos(u y) +CW FT (s) sin(u y)] ,

s
AC () = [SW FT (s) sin(u y) CW FT (s) cos(u y)] .

AS () =

(3.35)

Finally the wave pattern resistance is determined by (3.30) with the free wave spectra from (3.35).
3.2.4 Application aspects

The longitudinal wave cut method is favoured over other wave analysis methods. Advantages of the
longitudinal cut method are:
In case of flows symmetric to the ship center plane a single wave cut parallel to the ship suffices
to perform the whole analysis. No wave slopes need to be determined.
Asymmetric flows can be treated smoothly.
The tendency of over-predicting the stern waves in non-viscous CFD simulations or the viscous
wake flow phenomena in EFD does not affect the analysis as much as in case of transverse wave
cuts.

36

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

3.2 Wave cut analysis


The longitudinal wave cut method is comparatively less affected by lateral stretching of the free
surface panelization in CFD, provided the longitudinal wave cut stays in the neighbourhood of
the ship.
The method is well applicable to wave probe measurements in a towing tank which allows easy
comparison between CFD and EFD.
Disadvantages of the longitudinal cut method are:
The method is error-prone in case of an improper application of the truncation correction which,
usually, is avoidable by monitoring the fitting process.
Proper application of the truncation correction demands for comparatively large free surface
domains in CFD, extended fairly downstream, which is costly both in terms of computation time
and memory consumption.
A bidirectional interpolation of the free surface mesh is required, first transversally then longitudinally, in order to determine the longitudinal wave cut from computed discrete free surface
elevations. This introduce further inaccuracies.
Basically, inverting the advantages of the longitudinal wave cut method to disadvantages and vice
versa yields the equivalents for the transverse cut method. A comprehensive overview of the underlying theory of the different methods is given in Eggers (1976). Advocates of the transverse wave
cut method are for instance Nakos (1991) and Raven and Prins (1998a,b). Whereas, recent studies to
the applicability of the longitudinal wave cut method, especially focusing on measured wave cuts of
fast and high-speed crafts, were reported by Dumez and Cordier (1996), Brizzolara et al. (1998) and
Lalli et al. (1998, 2000). Spinney (2002) compared four different longitudinal and transverse wave cut
techniques in a thorough numerical study applying the linear and nonlinear Rankine panel module of
the SHIPFLOW code. In terms of the longitudinal wave cut method he utilized an implementation of
the method following Michelsen and Uberoi (1971) and the SWASH code. Moreover, implementations
of the single and multiple transverse wave cut method following Eggers et al. (1967) were utilized. He
found that the multiple transverse wave cut method is robust in application, however, some care is
required in selecting appropriate transverse cutting planes, see also Raven and Prins (1998a,b). In his
study the longitudinal wave cut method, particularly the truncation correction, appeared to be sensitive
to the signal length and the lateral position of the cutting plane.
According to the experience of the author, improper application of the truncation correction in the
longitudinal wave cut method can be avoided by a proper setup of the fitting process. An alternative
approach of dealing with truncated wave cuts is the application of the so called Fourier Transformation
Proper (FTP) as proposed by Schmiechen (1999). He claims that his method permits precisely to
identify spectra from truncated records, i.e., from a finite set of sampled data of transient signals and
that it overcomes the problem of classical uncertainty.
Sding (2001b) proposed a different strategy of wave pattern analysis which is independent of longitudinal and transverse wave cuts. In his method the coefficients of the harmonic wave components of all
feasible directions which constitute the free wave system are determined directly by a least-square fitting procedure. The fitting is conducted with respect to the computed (by CFD) wave elevations at the
discrete free surface collocation points sampled in a rectangular region aft of the ship. Sdings wave
analysis method requires special care in the setup of the free surface discretization and the selection of
the fitting region.

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

37

3 Hydrodynamic analysis
In the present work the longitudinal wave cut method with truncation correction following Sharma
(1966) and Eggers et al. (1967) is applied for the wave pattern analysis. Practical application of longitudinal wave cut analysis mainly depends on the ships speed, on the lateral position of the cutting plane
yWC and on the utilizable length of the wave signal up to the truncation point xE , i.e., the downstream
edge of the computational free surface domain. The latter two influences are in opposite direction. In
order to avoid serious analysis errors and to fulfill the requirement of a free wave system the cut is to
stay outside a region close to the hull where local flow effects are predominant. However, the greater
the lateral distance, the shorter is the utilizable wave signal until truncation, which is a detrimental
effect of the Kelvin angle and the limited extension of the computational domain. Various studies
indicate that satisfactory results are obtained for medium speed and fast monohulls at lateral cut positions in the range 0.2 y/LPP 0.3 provided the cut extends far enough downstream to allow for
a reasonable truncation correction. The cut should feature several times the fundamental wavelength.
For details see Lalli et al. (1998, 2000), Nowacki et al. (1999), Heimann (2000) and Valdenazzi et al.
(2003).
The effects of the lateral position of the cutting plane yWC and the signal length up to the truncation
point xE on the wave pattern resistance from longitudinal wave cut analysis are subsequently demonstrated by means of a theoretical and a practical example.
In the theoretical example the linearized Neumann-Kelvin wave pattern of an even submerged dipole
was investigated by wave cut analysis. The dipole moment was adjusted so that the problem of a flow
about a sphere of unit radius, located at the coordinate origin directly beneath the free surface level, was
simulated. A basic wavenumber of k0 = 1 was adopted which conforms to a fundamental wavelength
of 0 = 2 . The exact theoretical, non-dimensional wave resistance is RW = 1.528, see e.g. Sharma
(1966). The linearized wave pattern was determined from the expressions in the original publication by
Havelock (1928). A representation of the wave system, computed with the aid of the MATHEMATICA
package, is given in figure 3.3. Longitudinal wave cuts were evaluated for the center plane (including
the local wave system) and for three transversal positions at yWC /0 = 0.5, 1, 2, with a signal length of
up to xE /0 = 10. The wave cuts off the center plane were analysed with SWASH with and without
correction of the truncation error as well as for three different signal lengths xE /0 = 5, 7, 10. A
comparison of the wave pattern resistance from wave cut analysis with the exact theoretical value
is shown in figure 3.4 and 3.5. Figure 3.4 reveals that the wave pattern resistance approaches the
exact theoretical value with increasing signal length. Whereas figure 3.5 illustrates the immanent
effect of shortening the utilizable wave signal length by shifting the wave cut transversally. A large
error is introduced to RW P at the outmost wave cut. The truncation correction proves to be capable
of recovering the wave energy contained within the wave pattern downstream of the truncation point
fairly well. Generally, in order to allow a truncation correction a minimum fitting length (tail end of the
wave cut up to the truncation point xE ) of 0.5 0 is required. However, in terms of a stable regression
analysis a larger fitting length of about one fundamental wavelength is advisable. The total effect of
the fitting length on the wave pattern resistance is negligible as various tests have shown.
In the second example the wave pattern of the well-known Hamburg test case (alias Ville de Mercure) was studied. According to Bertram et al. (1992) the Fn = 0.24, LPP = 153.7 m, TF = 9.2 m
and TA = 10.3 m, CB = 0.652. The single-screw container vessel features a bulbous bow, a stern
bulb and a transom stern. Longitudinal wave cuts for varying transversal positions in the range
0.262 yWC /LPP 0.574 and signal lengths in the range 2 xE /LPP 4 were determined by a
nonlinear free surface CFD simulation with SHIPFLOW allowing trim and sinkage to adjust freely.
A comparatively dense hull and free surface panelization was adopted, extending 4 LPP downstream,

38

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

3.2 Wave cut analysis

Fig. 3.3: Linearized wave pattern of an even submerged dipole beneath the origin, k0 = 1 and 0 = 2 .
comprising the Kelvin wedge, with a longitudinal and lateral panel stretching towards the hull. The
computed wave cuts were analysed with SWASH and the results in terms of the wave pattern resistance coefficient 21 CW P were compared to the corresponding results from measurements. The latter
were attained from wave probe measurements conducted by the Hamburg Ship Model Basin (HSVA),
Bertram et al. (1992). No tank wall reflections were present in the measured wave cuts up to a signal
length of xE /LPP = 4. Both the computed and the measured wave cuts were corrected for truncation
errors. Figure 3.6 shows that already a signal length of xE /LPP = 2.5 yields almost the same value for
CW P as the longer wave cuts for the innermost position. The wave pattern resistance from the com21

applying the same reference surface for normalization

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

39

Wave pattern resist. ratio RWP / RW exact [%]

3 Hydrodynamic analysis

100

90

80
wave cut position : yWC / 0 = 0.5
truncation correction

70

no truncation correction
60
5

7
8
9
Normalized signal length xE / 0 [-]

10

Wave pattern resist. ratio RWP / RW exact [%]

Fig. 3.4: Related wave pattern resistance of the even submerged dipole, influence of the signal length
xE and the truncation correction.

100

90

80
signal length : xE / 0 = 10
truncation correction

70

no truncation correction
60
0.5

0.75
1
1.25
1.5
1.75
Normalized wave cut position yWC / 0 [-]

Fig. 3.5: Related wave pattern resistance of the even submerged dipole, influence of the wave cut position yWC and the truncation correction.

puted wave pattern exceeds those of the measured wave cuts. This phenomenon is attributed to a slight
over-prediction of the stern waves in the CFD simulation which in turn is due to the absence of viscous
damping. However, more significant is the decrease of the resistance values with increasing transverse
cut position in figure 3.7. This effect is considerably amplified in case of the computed wave pattern.
But, it is not fully explained by the shortening of the utilizable wave signal length as the comparatively
gentle decrease of the experimental curve suggests.

40

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

Wave pattern resist. coeff. CWP 1000 [-]

3.2 Wave cut analysis

0.2

0.15

0.1

wave cut position : yWC / LPP = 0.262


SHIPFLOW

0.05

Experiment
0
2

2.5
3
3.5
Normalized signal length xE / LPP [-]

Wave pattern resist. coeff. CWP 1000 [-]

Fig. 3.6: Wave pattern resistance of the Hamburg test case determined from computed (SHIPFLOW)
wave cuts and wave probe measurements (HSVA), influence of the signal length xE .

0.2

0.15

0.1

signal length : xE / LPP = 4


SHIPFLOW

0.05

Experiment
0
0.25

0.3

0.35
0.4
0.45
0.5
Normalized wave cut position yWC / LPP [-]

0.55

Fig. 3.7: Wave pattern resistance of the Hamburg test case determined from computed (SHIPFLOW)
wave cuts and wave probe measurements (HSVA), influence of the wave cut position yWC .
This effect is attributed to numerical damping (amplitude error) inherent in the CFD simulation. The
effects of numerical damping are thoroughly studied by Raven (1996) and by Janson (1997), the latter
by means of the SHIPFLOW code. The main error sources are the discretization of the continuous free
surface source distribution and the difference scheme used for evaluating the velocities and their derivatives in the free surface boundary condition. Tests suggest that the lateral panel stretching, inevitable
in the majority of cases, is further impairing the numerical damping. Hence, lateral stretching of the
free surface panelization ought to be avoided or at least kept small. Moreover, a sufficient number
of panels per fundamental wavelength ought to be placed uniformly in longitudinal direction. Studies

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

41

3 Hydrodynamic analysis
suggest the number of panels per fundamental wavelength 0 should be at least 25 in order to capture
most of the shorter wave components. The highest wave frequency that can be resolved by the free
surface panelization is defined by the Nyquist cut-off frequency which is proportional to the width of
the outermost lateral panel strip, if lateral panel stretching is used.
While computed wave pattern, and hence wave cut analysis based on them, prove to be less sensitive to
flow field discretization than hull pressure integration, the wave pattern resistance is affected by cumulative numerical damping effects. Consequently the wave pattern resistance deduced from computed
wave pattern typically is smaller than the wave resistance from pressure integration. Nevertheless,
in terms of the optimization application the error due to numerical damping in the CFD simulation
is of minor concern since only relative differences of hull variants are compared without explicitly
addressing the absolute resistance value.
It should be emphasized that the combination of a nonlinear free surface CFD simulation with a wave
cut analysis approach, which itself is based upon linearized theory assuming a vanishing effect of
nonlinear contributions, is inconsistent from a mathematical point of view. However, this approach
seems justified by its successes in optimization applications and in the application of wave cut analysis to measured wave cuts, as the pertinent literature sources given throughout this chapter suggest.
Moreover, a numerical superposition study conducted by Heimann and Harries (2003) indicates that
the linear wave cut analysis is capable of accurately tracing, on a relative basis, the nonlinear impact
of locally confined hull form variations 22 in the nonlinear wave pattern.

22

(a) a bulbous bow variation, (b) an entrance variation

42

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

4 Sensitivity analysis
The proposed hydrodynamic optimization scheme utilizes gradients or sensitivity terms for
the optimizations. The impact of hull variation on the system state quantities is expressed
in terms of gradients. Special flow related optimization variables are introduced. The nonlinear boundary value problem is accessed by a linear perturbation approach allowing a
straightforward and fast computation of the gradients. The process flow of the sensitivity
analysis is outlined.
The present chapter focuses on the evaluation of the gradients, i.e., the sensitivity terms 1 of the objective function and the constraints as needed for the optimization scheme. Nonlinear relations between
the system state quantities and the optimization variables are accessed by a linear perturbation approach. This allows a straightforward and fast computation of the gradients. Moreover, it enables a
consistent tracing of the impact of the variation of a single optimization variable on the system state
quantities like, e.g., the free wave spectra and the wave pattern resistance. In addition, the variation
impact on certain hull form parameters, some of which are constrained within the optimization, is
explained.
The main part of this chapter focuses on the theoretical background of the gradients evaluation. In
section 4.8 practical aspects of the gradients evaluation are discussed and an illustration of the gradients
with respect to local hull variations is given.
The derivation of the gradients is based on discrete relations as predetermined by the discrete view of
the underlying boundary element flow solver as outlined in section 3.1. Hence, the reader is recommended to first catch the basic concept of the underlying hydrodynamic analysis which is outlined in
chapter 3 before continuing with the present chapter.
All quantities defined or derived in this chapter are given in non-dimensional form, normalized by the
ship speed and/or the length between perpendiculars.

4.1 Design variables


Normally, the design variables in ship hull form optimization are directly related to the hull geometry.
For instance, nearly all recent optimization approaches attain control of the hull form variation by
utilizing either hull offsets, explicit polynomial functions, B-Spline basis functions or even higher level
geometric quantities like hull form parameters. Other methods adopt shift functions in order to modify
basic ship curves like sections and/or waterlines. A comprehensive overview of hull form variation
approaches can be found in Nowacki et al. (1995).
1

The terms gradients and sensitivities are used synonymously.

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

43

4 Sensitivity analysis
Hull form variation by shifting hull offsets in general gives the highest degree of freedom for form
variation, but increases the computational costs since a vast amount of free variables, i.e., offset coordinates, need to be controlled. In contrast the design variables are effectively condensed and thus
reduced in higher level hull modelling and variation approaches like in form parameter based schemes,
see Birk (1998) and Harries (1998). Nevertheless, the more general the design variables the more
global is their influence on the hull form, thus restraining the ability of locally confined hull variations. In either case, the individual choice of the hull form variation method and the particular design
variables very much depends on the underlying hull geometry representation.
The basic concept of the present approach particularly differs in the sense that the design variables
ought to be independent of the hull geometry representation and its characteristic features. Within
the current context the variation of the hull geometry should be understood as an implicit and derived
property in the optimization since actual control of hull variation is achieved by a non-geometric quantity.
The present choice of the design variables, more precisely called optimization variables, is based on
the following ideas:
The optimization variables ought to be self-explanatory and self-consistent.
They ought to be independent of the hull geometry representation and its characteristic features.
The optimization variables constitute a link between the hull geometry and their hydrodynamic
properties in terms of wave formation and wave pattern resistance.
Their influence should be locally confined.
As outlined in chapter 3 the computation of the steady ship wave system is carried out with the aid of the
nonlinear free surface Rankine panel module XPAN of the CFD system SHIPFLOW. The underlying
boundary value problem is solved by discretization of the boundaries of the flow domain, essentially,
the hull and the wavy free surface. In the first order panel method the boundaries are approximated
by flat quadrilateral panels. Potential functions are introduced by distributing source singularities of
constant density over each panel. 2 The boundary conditions on the hull and the wavy free surface are
enforced for each panel at a control point defined as the panels centroid. The direct solution of the
boundary value problem leads to a linear system of equations for the unknown panel source densities.
See section 3.1 for a closer description of the panel method.
In view of the above considerations the author concluded that the hull panel source densities ideally
meet the demands on the optimization variables. On the one hand the source densities are the principal
parameters in the hydrodynamic analysis which fulfill the governing Laplace equation and lead to the
free surface elevation and, thus, to the free wave spectra and to the wave pattern resistance. On the other
hand they are linked to the hull geometry, even if only in discretized form, since the hull panels are
employed to satisfy 3 the kinematic hull boundary condition. However, the hull panel source density
distribution is independent of the hull geometry representation and its characteristic features.
In this chapter perturbation relations are introduced which substitute the essentially nonlinear boundary
value problem by explicit, linear relations. In fact, provided the variations are small the perturbation
2

Alternatively a higher order method can be selected which applies bi-parabolic panels with a linear source density distribution along each panel.
3 in a discrete sense

44

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

4.2 Perturbation approach


relations establish an explicit and linear link between the source density distribution and the crossconnected boundaries of the flow domain in terms of the location and orientation of the hull and free
surface panels. The perturbation approach is presented in section 4.2.
From a theoretical point of view all hull panel source densities may serve as free optimization variables.
However, in practical applications, as presented in chapter 6, only specific hull regions are usually subject to an optimization like the bulbous bow, the entrance and/or run while other regions are required to
stay untouched. The latter is accomplished in the present scheme by freezing whole hull panel regions
throughout the optimization or even by activating or deactivating single hull panels. The influence of a
single hull panel source density variation is considered a locally traceable effect.

4.2 Perturbation approach


In the present optimization approach hull variation is controlled by a large number of variables in terms
of hull panel source densities. The optimization scheme is based on gradients. Two problems arise:
At first, an explicit relation between the optimization variables and the hull geometry itself is required.
Furthermore, the evaluation of the gradients by means of direct CFD simulations rapidly generates
high computational costs in view of the large number of optimization variables. In order to overcome
these problems the present optimization method is based on a perturbation approach. The nonlinear
boundary conditions are transferred to explicit equations depending in a linear and straightforward
way on the optimization variables. In other words, the nonlinear boundary value problem is linearized
around the actual state of the system. Thus, the whole cause and effect chain starting with the variation
of the optimization variables, resulting in a hull geometry 4 variation, finally affecting the wave pattern,
the free wave spectra and the wave pattern resistance can be consistently traced.
4.2.1 Perturbation of the discretized boundary value problem

The unknown source densities are related to the boundary conditions on the hull and the free surface
in a nonlinear manner, compare section 3.1. However, the present optimization approach is based on
explicit, linear perturbation relations for the gradients evaluation in order to
access the inverse design problem in a straightforward way, i.e., seeking a disturbance of the hull
boundary (and free surface boundary) which matches a desired, predefined hull source density
distribution,
bypass costly direct numerical solutions of the boundary value problem by means of CFD simulations for a large set of variables.
The perturbation method is considered an approximation procedure permitting a straightforward treatment of the nonlinear boundary value problem. The fundamental idea of the perturbation approach is
as follows: A small parameter is introduced into the problem such that an already known 0th -order
solution corresponds to = 0 and that all higher order perturbations of this basic solution are traceable
by their specific power of . Perturbation approaches are asymptotic power series expansions with the
property that all higher order series terms approach zero faster than all other preceding series terms
as approaches zero. The benefit of this approach is that any nonlinear problem can be decomposed
4

hull geometry in terms of a hull surface discretization by flat quadrilateral panels

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

45

4 Sensitivity analysis
into a sequence of linear problems with ascending order by sorting in powers of . The linear subproblems can then be solved separately one after the other. Yet, a restriction to the 1st -order problem
already yields a meaningful approximation of the fully nonlinear problem, presuming the perturbation
magnitude is kept small. The approximation is known to get closer the smaller gets. A detailed
representation of the perturbation theory can be found in Cole (1968) and van Dyke (1975).
Practical applications of perturbation methods frequently deal with the inverse design problem, i.e.,
searching for a design which meets one or more desired, predefined performance criteria. A systematic
perturbation study is carried out for a parent design which then gets adapted in a subsequent step in
order to satisfy the desired, predefined criteria. Other implementations of perturbation methods are
aiming at an acceleration of the optimization process. Expensive direct numerical evaluations of the
cost function are by-passed by simple, explicit perturbation relations. The present work benefits from
both these practical aspects of the perturbation method as will be pointed out shortly.
Performance criteria in inverse design applications are flow related state variables like a desired distribution of the dynamic hull pressure, the flow velocities along the hull or in the wake. Geometrical
quantities are used in terms of offset coordinates or local inclinations with respect to the incident flow
direction. The geometry is varied either by simply shifting offset points or spline vertices, by distortion
approaches or the variation of high-level form parameters. In case of discretized geometries, cells,
patches or panels are either shifted or inclined or both. In inverse design applications the decisive flow
and geometrical quantities are both expanded in terms of perturbation expressions which are applied to
the governing equations of the flow problem. By utilization of series expansions and sorting in powers
of the perturbation parameter the nonlinear boundary value problem is then transferred to mostly linear relations linking flow and geometry. Practical applications of the perturbation method to 2-D and
axis-symmetric inviscid flow problems are reported, e.g., in Schmidt (1976), Bristow (1974, 1980) and
Favre et al. (1987). Kraus (1989) extended the application range to 3-D problems. Huan and Huang
(1998) formulated both a perturbation and an adjoint equations method which they applied to the inviscid nonlinear free surface flow problem. They successfully validated both techniques by a comparison
with direct CFD solutions and experimental data for the flow about the Salvesen hydrofoil below a
nonlinear free surface. An application to an inverse hydrofoil design problem is also considered in their
work. Schulze (1996) reported a comprehensive investigation of the practicability of locally confined
perturbations for both inviscid and viscous flows about 2-D profiles.
The perturbation relations derived in this chapter are based on discretized boundary equations as the
discrete view of the underlying boundary element flow solver implicates. This does not imply a loss
of generality since the discretized relations do approach their continuous counterparts with increasing
mesh refinement. The governing equation for the panel source densities is given in discretized form by
(3.11)
NF S

j Ai j = Bi

j=1

i = 1, , NFS .

NFS accounts for the total number of panels which includes the hull panels NH and the free surface
panels N f s , i.e., NFS = NH + N f s . Flat quadrilateral panels are used with a constant source density
distribution across the panel.
The perturbation relations are derived by expanding both flow and geometrical quantities in terms of
perturbation expressions which are then applied to the boundary value problem. The panel source
densities represent the unknowns in the flow solution and at the same time are selected as optimization

46

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

4.2 Perturbation approach


variables. Expanding the normalized (by U ) panel source densities per unit area in a perturbation
expression and truncating after the linear term yields
(0)

(1)

n = n + n + O(2 ) ,
(1)

n = n + O(2 )

n = 1, , N NH ,

with (0) indicating the 0th -order solution, i.e., the basic solution. Superscript
solution, i.e., the perturbation expression.

(4.1)
(1)

denotes the 1st -order

The corresponding perturbation of the hull geometry is realized by perturbating the ship hull surface in
its normal direction. The hull surface is approximated by flat quadrilateral panels. Hence, the panels
are relocated in their normal directions in- or outwards of the hull panel mesh. The panel relocation
technique is described in section 4.3. Expressing the normalized (by LPP ) hull panel normal relocations
in terms of a perturbation notation and truncating after the linear term yields
~Pk = ~P(0) + ~n(1) + O(2 ) ,
k
k
(1)
~Pk ~nk = ~nk + O(2 )

k = 1, , NK NH .

(4.2)

~Pk may be any point on the flat quadrilateral panel k. ~Pk will be later associated with the panel control
point. NK is the number of active hull panels applied in the panel normal relocation scheme. The
(1)
(1)
magnitude of the panel shift in panel normal direction ~nk is composed of the shift amplitude nk and
the unit vector normal nk
(1)

(1)

~nk = nk nk

~nk = nk nk .

The hull part (i = 1, , NH ) of the influence coefficients Ai j and the inhomogeneous term Bi in (3.11)
are pure geometrical quantities solely depending on the geometry, the relative position and the orientation of the panels. The free surface part (i = NH + 1, , NFS ) in addition includes the flow velocities,
generally, the free surface velocities and their derivatives from a previous step in a nonlinear free surface iteration. However, these velocities and derivatives are treated as constant, known terms in the current evaluation of the influence coefficients. Hence, the influence coefficients and the inhomogeneous
term are actually affected only by the geometry perturbations in terms of the panel normal relocations
(1)
nk . Since this dependency is implicit, both the influence coefficients and the inhomogeneous term
need to be expanded in Taylor series which may be truncated after the linear terms

(0)

Ai j = Ai j +

NK

(0)

(1)

nk

k=1
(0)

Bi = Bi +

NK

(1)

nk

k=1

Ai j

nk

+ O(2 ) ,

(0)

Bi
+ O(2 )
nk

i, j = 1, , NFS ,

(4.3)

in order to obtain explicit and linear relations in the perturbation parameter . Kraus (1989) verified the
validity of the Taylor expansions by means of systematic perturbation studies with a 2-D profile and a

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

47

4 Sensitivity analysis
rotationally symmetric Rankine ovoid, both deeply submerged. He computed the change of the influence coefficients Ai j due to a y-perturbation with varying magnitude of selected panel corner points.
For small y-perturbations in the range 10% of the initial y-offset Kraus found good agreement of the
Taylor expansions (4.3) with results from exact computations applying the Hess & Smith relations for
the Ai j . A similar study conducted by the author for the 3-D case confirmed the results of Kraus. As a
test case the flow around the standard Wigley hull at Fn = 0.3 including the free surface was studied.
Selected hull panels were systematically shifted in panel normal direction in- and outwards of the hull
surface. The resulting change in the influence coefficients Ai j , Xi j , Yi j , Zi j and the inhomogeneous term
Bi were computed by means of three different methods: SHIPFLOW/XPAN, a test program prepared
by the author and the linearized relations (4.3). A comparison of the the results approve the applicability of the linearizations (4.3) for moderate hull panel normal relocations in the range of 10% of a
typical hull panel size. 5 Larger perturbation values are feasible but should be verified beforehand for
the particular application case.
The linearization of the boundary relations can be conducted now by introducing the perturbation
approaches (4.1) and (4.2) together with the Taylor expansions (4.3) into the boundary value problem
(3.11)
NFS

j=1

(0) (0)
(1) (0)
(0)
j Ai j + j Ai j + j

(0)

NK

k=1

(1)
nk

Ai j

nk

(1)
+ 2 j

NK

k=1

(1)
nk

(0) )

Ai j

nk

(0)

= Bi +

NK

(1)

nk

k=1

i = 1, , NFS .

(0)

Bi
nk

(4.4)

Neglecting higher order contributions of O(2 ) and sorting in orders of yields for the 0th -order
NF S

(0) (0)
Ai j

j=1

(0)

i = 1, , NFS ,

= Bi

(4.5)

which constitutes the basic equation system. For the 1st -order terms one gets
NF S

j=1

NK

(0) )

(1) (0)
(0)
(1) Ai j
j Ai j + j nk
nk
k=1

NK

(1)

nk

k=1

(0)

Bi
nk

i = 1, , NFS ,

(4.6)

which represents the contribution due to the perturbation. After rearrangement, (4.6) can be written in
compact matrix notation
A(0) ~(1) = T(0) ~n(1)
,
[NFS NFS ] [NFS ] = [NFS NK ] [NK ]

(4.7)

with the transfer matrix 6


T
5
6

(0)

(0)
Tik

(0)

(0)

NFS
B
(0) Ai j
= i + j
nk
nk
j=1

i = 1, , NFS
.
k = 1, , NK

(4.8)

e.g., typical hull panel size in terms of the average square root of the hull panel areas, as will be discussed in section 4.2.2
(0)
A hull panel normal relocation does not affect the inhomogeneous terms, i.e., in this case Bi /nk = 0 for all k.

48

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

4.2 Perturbation approach


Equation (4.7) represents the fundamental linear perturbation relation relating the source density variations ~(1) and the panel normal relocations ~n(1) . The coefficient matrix A(0) in (4.7) is already known
from a base flow solution. The evaluation of the transfer matrix T(0) is the expensive part since the
partial derivatives of the coefficients with respect to the hull panel normal relocations have to be calculated. This can be done either analytically or numerically as will be addressed in section 4.8. As
expression (4.8) reveals the transfer matrix comprises the variation effect on both the hull and the free
surface portion. However, during the evaluation of the transfer matrix the free surface panelization
itself is temporarily frozen.
The hull panel source densities are selected as the free variables in the optimization process. Nevertheless, the optimization aims at an improvement of the wave-making characteristics which finally is
the result of a favourable adaptation of the hull form. Relation (4.7) provides the link between the
source density variations and the corresponding hull panel adaptations. However, the inverse relation
is required as will be derived below.
The inverse solution

The inverse solution yields a variation of the hull form which matches the predefined variation of
the hull source densities. The inverse relation is obtained by inversion of the transfer matrix T(0) in
(4.7) and solving for the hull panel normal relocations ~n(1) . For the inverse solution all rows and
columns of the vector and matrix elements in (4.7) which are associated with the free surface portion
(NK,H < i, j NFS ) can be omitted, temporarily, since only the hull source densities ~(1) are considered
as variables in the optimization, thus
~n(1) = T(0)1 A(0) ~(1)
.
[NK ] = [NK NK ] [NK NK ] [NK ]

(4.9)

In all inverse cases utilized in the optimization the transfer matrix is non-singular, diagonally dominant
and, as addressed above, temporarily square. Hence, the number of hull panel relocations equals the
number of equations which in turn permits a unique solution. 7 In the optimization context the equation
system (4.9) is to be solved N times for N optimization variables, i.e., N hull source density variations
~(1) . Since each time only the hull source densities on the right hand side vary equation (4.9) is solved
numerically in a straightforward manner by LU-decomposition and back substitution, for details of the
technique be referred to Press et al. (1992).
For the sake of completeness the over-determined case may be discussed briefly. In the over-determined
cases the number of unknowns, i.e., the hull panels included in the panel normal relocation scheme, is
less than the number of equations. For instance, one allows only a subset of all hull panels to adjust
by moving normally in- and outwards of the mesh (NK < NH ), whereas, at the same time a predefined,
desired variation of the hull source densities is to be met at all hull panels. This may be the case, e.g.,
when all panels which are part of the flat of side (FOS) and/or flat of bottom (FOB) are frozen (not
allowed to move) in order to maintain those specific hull characteristics in the optimization. However,
at the same time the same panels are forced to actively adjust a specific -variation which, for instance,
could be a zero source density variation.
7

In particular, as shown by Cramers rule, there is a unique solution if T(0) has a matrix inverse.

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

49

4 Sensitivity analysis
A solution of the over-determined system is possible only in a mean sense, i.e., by a minimization of
the residuals. For this purpose the perturbation relation (4.9) is reformulated in terms of a functional
expression which is to be minimized in a least-square sense
!2
NH

F=

i=1

NFS

(1) (0)

NK

(1)

(0)

j Ai j + nk Tik

j=1

= min .

(4.10)

k=1

The first order necessary conditions for a local extremum are


F

(1)
nl

=0

l = 1, , NK < NH ,

which yield
NH

(0)

Til

i=1

NFS

(1) (0)
Ai j +

j
j=1

NK

(1)

nk

k=1

(0)

Tik

=0

l = 1, , NK < NH .

(4.11)

Finally, the inverse relation for the over-determined case may be written in compact matrix notation
e (0)1 T(0)T A(0) ~(1)
~n(1) = T
,
[NK ] = [NK NK ] [NK NH ] [NH NH ] [NH ]

(4.12)

with
NH

e (0) Te(0) = T (0) T (0) = T(0)T T(0)


T
kk
ik
ik
i=1

k = 1, , NK < NH .

Equation system (4.12) represents the normal equations of the linear least-square problem (4.10). For
NK = NH the linear system for the over-determined case (4.12) reduces to the special case of (4.9).
The direct solution

Contrary to the inverse problem in the direct approach one seeks to determine the source density variations for a preceding variation of the hull form. Again the perturbation relation (4.7) is applied but this
time it is solved for the source density variations which result from a preceding variation of the hull
panelization, i.e., a normal relocation of the hull panels. Thus, in the direct case the source densities no
longer serve as optimization variables as in the inverse approach. Rather, this time the variation of the
source densities, particularly, those at the free surface result from a preceding hull perturbation. Since
it is searched for the variation of all panel source densities (hull and free surface) this time the free
surface portion in (4.7) must be taken into account, thus giving
~ (1) = A(0)1 T(0) ~n(1)
.
[NFS ] = [NFS NFS ] [NFS NK ] [NK ]

(4.13)

Equation system (4.13) is to be solved repeatedly, generally N times according to the number of
optimization variables, with varying right hand sides. This calls for a numerical treatment by LUdecomposition and back substitution.

50

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

4.2 Perturbation approach


In practical applications the free surface domain needs to extend far enough up- and downstream and
sidewise in order to allow for a proper wave cut analysis. At the same time the panelization has to
be dense enough to capture much of the wave information. This in turn results in large systems of
equations with a huge coefficient matrix 8 which is costly both in computation time and in memory
consumption. This gives rise to reducing the portion of the free surface domain considered in the
solution of the perturbation relation (4.13). 9 A validation study conducted by the author has shown
that a truncation of the free surface panelization sidewise is practicable to a certain extent as long as
the perturbations are kept small.
Alternatively, the source density variations may be calculated only in a mean sense. At the same time
the influence of the whole free surface domain is kept. Assuming that the number of source panels just
including a subset of the free surface domain, e.g., truncated sidewise, is NFS
f with (NFS
f < NFS ), then,
analogous to (4.10), the direct problem (4.13) can be expressed in terms of a functional expression
which is minimized in a least-square sense. This yields

with

e (0)1 A(0)T T(0) ~n(1)


~ (1) = A

 
 

,
NFS
f = NFS
f NFS
f NFS
f NFS [NFS NK ] [NK ]

(4.14)

NFS

e (0) A
e(0) = A(0) A(0) = A(0) T A(0)
A
jj
ij
ij
i=1

j = 1, , NFS
f < NFS .

Equation system (4.14) represents the normal equations of the linear least-square problem. For NFS
f =
NFS the linear system (4.14) becomes identical to (4.13).
4.2.2 Perturbation magnitudes

As outlined in the preceding section a perturbation scheme is applied to linearize the boundary value
problem. The basic assumption is that the magnitude of the perturbation is small. Since the hull
panel source densities and the hull panel normal relocations are selected as perturbation quantities
their magnitudes both need to be bounded in a perturbation sense. In order to keep the scheme robust
and straightforward the same perturbation magnitude is adopted for all hull panels.
The perturbation magnitude for the normalized hull panel source densities, in (4.1), is chosen as a
fixed small value in order to comply with the perturbation assumptions. The perturbation magnitude
may be simply related to a characteristic velocity U (e.g. the incident flow velocity)
0 < 1 ,

n = = U

(4.15)

controlled by a perturbation parameter . Alternatively, the perturbation magnitude may be related


to a typical source density value, for instance the average absolute value of all hull panel source
densities
NH

(0)

|i |

n = = i=1
NH
8
9

0 < 1 ,

(4.16)

A coefficient matrix A typically possesses 25106 up to 100106 elements.


Note: The underlying CFD simulation by SHIPFLOW still includes the whole free surface domain in the solution.

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

51

4 Sensitivity analysis
or to the average of the weighted absolute value of all hull panel source densities. If the hull panel
source densities are weighted by their panel areas Si the perturbation magnitude may be defined as
NH

(0)

Si |i |

n = = i=1 N
H
Si

0 < 1 .

(4.17)

i=1

Since the numerator in (4.17) can be viewed as the perturbation of the volume flow through hull panel
i, with
(0)

Vi = Si |i |

0 < 1 ,

definition (4.17) can be equivalently interpreted as an increment of the average volume flow through
all hull panels. The advantage of the definition (4.17) for the perturbation magnitude is that it accounts
for the specific hull panelization and the flow case in question which allows a more intuitive and
appropriate choice of .
Similarly, the perturbation magnitude for the hull panel normal relocations, nk in (4.2), can be selected. For instance, the perturbation magnitude may be simply related to a characteristic length L (like
the length between perpendiculars or the waterline length)
0 < ~n 1 ,

nk = n = ~n L

(4.18)

controlled by a perturbation parameter ~n . Alternatively, the perturbation magnitude may be related to


a typical hull panel size, e.g., the average hull panel diagonal or the average square roots of the hull
panel areas Si
NH

nk = n = ~n

i=1

NH

Si
0 < ~n 1 .

(4.19)

Again the advantage of the more elaborate definition is that it accounts for the specific hull panelization,
particularly, in terms of the panel mesh resolution 10 which allows a more intuitive and appropriate
choice of n.
Many more choices of the perturbation magnitudes are conceivable. Which of the alternatives are
eventually utilized to define the perturbation magnitudes is a practical decision by the user. For the
perturbation scheme it is of minor importance as along as the perturbation magnitudes are kept small
in a perturbation sense.
In the present optimization scheme the more general definitions (4.17) and (4.19) for the perturbation
magnitudes are implemented. In either definition the perturbation magnitudes are already related to the
specific hull panelization and the flow case in question. This allows a more general determination of
the small perturbation parameters and ~n so as to cover a broad range of hull shapes, flow conditions
and hull panelizations at the same time. From a practical application point of view the perturbation
magnitudes and n should correspond. They must still be selected large enough in order to produce
detectable variations 11 of the system state quantities and at the same time they need to be small enough
10
11

The perturbation magnitude n decreases with increasing panel mesh refinement.


in terms of variations in a significant digit which is required to be at least some order of magnitude higher than the adopted
mantissa length

52

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

4.3 Panel relocation scheme

Input Points Flat H&S Panels

z
0.5
2

0
0

y
1

1
x

2
3

Fig. 4.1: Artificial bicubic surface patch and panelization by 3 3 flat quadrilateral panels (right).
in order to comply with the perturbation approach. A systematic study conducted by the author with
different ship applications and with varying panel mesh resolutions determined a value of = ~n = 0.1
for the perturbation parameters as most suitable when defining the perturbation magnitudes according
to (4.17) and (4.19).

4.3 Panel relocation scheme


Whether the wave-making characteristics of a ship hull are considered good or bad is basically determined by the hull shape. Hence, the final outcome of an optimization is a new hull shape or a
favourable hull shape variation.
The present optimization scheme is based on a discrete view of the hull surface, i.e., a hull panelization
by flat quadrilateral Hess & Smith panels with a constant source density across each panel. The panel
patch-work rendering the hull surface is G0 -discontinuous in this case. That means, small gaps will
arise between adjacent panel edges. The method of formation of flat quadrilateral Hess & Smith panels
from a regularly ordered input or offset point mesh is briefly discussed in section 3.1.2. A detailed
description is given in the original publication by Hess and Smith (1962).
The hull modification originates from a variation of the hull source densities. Different choices for
the modification of the hull panel mesh are feasible. Two alternatives should be addressed. A simple
but effective approach is to relocate the underlying input point mesh which constitutes the hull panels.
From a topological point of view input points are equivalent to offset points which often are arranged
along hull sections and profile curves and are located directly on the hull surface. Different schemes
exist. Still popular and simple is to shift hull points in lateral (mostly in y-direction) direction in order
to attain a hull form variation, e.g. see Kraus (1989). In this case the panel mesh adjusts indirectly. An
alternative technique is to shift single hull panels directly in- and outwards of the mesh. Unless special
precautions are taken considerable gaps may develop between adjacent hull panels which requires a
certain control mechanism.

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

53

4 Sensitivity analysis
Input Points Flat H&S Panels Transformations

Fig. 4.2: Relocation of the corner points of the center panel in panel normal direction (left), effect of
the equivalent relocation of the four central input points (right).
The advantages and disadvantages of either approach may be summarized briefly, starting with the
input/offset point relocation scheme:
Easy application points are simply shifted by their offset coordinates,
no special treatment of the panels is necessary, they are self-adjusting, accordingly,
the initial topology of the input point mesh is kept at all time,
points are free to move since no dependency on the neighbourhood needs to be maintained, 12
constraints for the input points like, e.g., a symmetry condition at the midships plane are relatively easy to enforce, but
on the other hand the forced relocation of the hull panels is a dependent transformation only
indirectly controllable by the transformation of the underlying input points,
panel shape and area considerably change and hence are not invariant against the transformation,
at the same time panel rotations are enforced,
special effort is required to maintain geometric boundary conditions and,
since in practice the number of optimization variables does not match the number of the underlying input points an over-determined system is to be solved which leads to an averaging in a
least-square sense with somehow ambiguous results.
Advantages and disadvantages of the hull panel normal relocation scheme are:
It can be considered the discrete version of a continuous hull surface variation scheme,
the variation of the hull panels can be controlled directly,
12

unless the initial topology of the input point mesh is violated, e.g., by point intersections

54

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

4.3 Panel relocation scheme


Input Points Flat H&S Panels Transformations

Fig. 4.3: Homogeneous panel relocation (left), effect of the joint relocation (right).
no panel rotations are enforced, 13
panel shape and area are invariant against the panel transformation, 13
since the (variable) hull source densities and the hull panel relocations can be mapped one-toone, typically a unique solution of a determined linear system is expected, but
on the other hand a more expensive implementation is required, because panel-connectivities as
well as relocations of both the panel corner and the input points have to be traced,
special effort is necessary to maintain geometric boundary conditions,
the topology of the input point mesh is violated since the inverse formation of a coherent input
point mesh from independently displaced panels is impossible (contradicting demands),
relocation of adjacent panels in different directions or with different magnitudes cause considerable gaps in the panel mesh
which requires a special treatment of neighbouring panels and a reconstruction scheme in order
to restore a unique hull surface from the emerging incoherent panel cluster.
With the above considerations in mind it is meaningful to combine both approaches so as to combine the
advantages of either approach and at the same time to control or even to neutralize the disadvantages.
Consequently, the following hull panel relocation scheme has been developed which, in some sense,
performs both approaches successively.
13

as long as no hull surface reconstruction scheme has been applied

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

55

4 Sensitivity analysis
Input Points Flat H&S Panels Transformations

Fig. 4.4: Stepped panel relocation (left), effect of the joint relocation (right).
4.3.1 Joint relocation of hull panels

The idea behind the joint relocation scheme is simple: first, let the hull panels shift separately in their
normal directions according to the optimization requirements and, in a subsequent step, redistribute the
panel normal relocations back to the associated input point mesh by a vector averaging. In the present
work this scheme is supported by an inequality constraint in the optimization which keeps track of
adjacent panel relocations in order to ensure smooth transitions in the distorted panel mesh, see section
5.2.2.
The relocation scheme is illustrated by means of an artificial bicubic surface patch (figures 4.1 to
4.4), a generic ferry-type stern hull surface patch possessing y-symmetry (figures 4.5 to 4.8) and by a
SHIPFLOW panelization example featuring the transition area from the bulb to the bow (figures 4.9 to
4.13). The surface patches in the examples are panelized by flat quadrilateral panels according to the
Hess & Smith method. Initial small gaps, i.e., G0 -discontinuities, can be observed between adjacent
panel edges as is immanent in the method.
Figure 4.2 illustrates the effect of moving a single panel outwards of the mesh in its normal direction,
see picture to the left. The panel normal relocation is achieved by simply shifting the four panel corner
points homogeneously in panel normal direction. Apparently, large gaps arise between the panel edges
of the shifted center panel and the yet untouched adjacent panels. 14 Alternatively to the relocation of
the panel corner points the associated four input points may be shifted in the panel normal direction by
the same amount. Even though, in both cases, the final shifted position of the center panel is identical,
in the latter case the adjacent panels are forced to follow the movement of the input points and, thus,
panel gaps have been reduced considerably, see picture to the right. Unfortunately, this property is
purchased by the penalty of an unwanted panel rotation of the adjacent panels.
14

Note: For illustration purposes the panel relocations in the examples are exaggerated, compared to the panel perturbations
allowed in the hull optimization.

56

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

4.3 Panel relocation scheme

Input Points Flat H&S Panels

z 1

0.5

0
0

1
-1

2
x

3
4

-2

Fig. 4.5: Generic ferry-type stern hull surface patch (upside down), y-symmetry, and panelization by
4 4 flat quadrilateral panels (right).
So far only a single panel has been touched at a time. In view of the perturbation scheme this is a
somehow artificial case never arising in practical applications. In the variation scheme all hull panels
or at least a panel cluster will be adapted at the same time. Figure 4.3 shows a homogeneous panel
relocation of the whole patch and in figure 4.4 a stepped panel relocation is presented. Apparently,
as can be observed in the pictures to the left, shifting each panel separately ignoring the neighbourhood causes considerable gaps since the panels start detaching or intersecting which is amplified with
increasing panel displacement. Obviously, a technique is required to resize the gaps to normal size
as implied by the Hess & Smith method and to maintain the initial topology of the input point mesh.
However, at the same time the characteristic of the panel relocation must be kept as best as possible.
The idea is to redistribute the panel normal relocations back to the associated input point mesh by a
vector averaging in a subsequent step.

Let IPm denote the vector representation of the m-th input point and Nm the number of adjacent panels 15
with 1 Nm 8. In order to reconstruct a coherent input point mesh with a well defined topology the
(1)
panel normal relocations ~nk have to be redistributed back to the associated input points. This can be
achieved by vector averaging

(0)

IPm = IPm + IPm


1

(0)
= IPm +
Nm

Nm

(1)

~nk(i

im =1

m ,m)

m = 1, , NM
.
1 Nm 8

(4.20)

The final adaptation of the panels then is deduced from the thus reconstructed input point mesh, again
by means of the Hess & Smith method. Examples for the joint relocation scheme are presented in the
pictures on the right hand side in figures 4.3 and 4.4 and in the following figures.
15

In case of quadrilateral panels a maximum of four panels are linked to a single input point. Since in case of symmetrical
ship hulls the mirror panels across the symmetry plane are also considered in the relocation scheme, topologically twice
the number of panels may be linked to a single input point at maximum.

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

57

4 Sensitivity analysis
Input Points Flat H&S Panels Transformations

Fig. 4.6: Homogeneous panel normal relocation (left), effect of the joint relocation (right).
The joint relocation scheme is applied to construct a smooth input point mesh with a well defined
topology. The point mesh can be further processed by surface interpolation or approximation and fairing techniques, see e.g. Nowacki et al. (1998), Nowacki and Kaklis (1998) and Westgaard (2000). In
the present work the computer aided ship hull design (CASHD) tool MultiSurf from AeroHydro Inc.,
Letcher et al. (1999), is utilized as a geometrical postprocessor to the optimization scheme. MultiSurf
is applied in a script technique in order to generate higher-level B-Spline hull surface representations
from the input point mesh data in an automated procedure, compare section 5.4. Moreover, this simplifies the export of the hull geometries to other systems and/or applications via portable and freely
convertible formats (like IGES).
Properties of the joint relocation scheme

Taking into account the exaggeration of the panel relocations in the examples and considering that
sign and magnitude of the panel relocations in ship application will be somehow smoothly distributed
across the hull the following properties of the joint relocation scheme can be concluded:
The technique is considered a discrete version of a continuous hull surface variation scheme,
the initial panel orientations are kept by the scheme,
panel shape and area are changed only slightly,
a coherent input point mesh with a well defined topology, the initial mesh topology, is kept,
gaps between adjacent panel edges remain small, as is typical of the Hess & Smith method,
the variable hull source densities and the panel normal relocations can be mapped one-to-one
which leads to unique solution of the perturbation relations, but,
undesired yet unavoidable panel rotations are introduced due to the averaging procedure and

58

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

4.3 Panel relocation scheme


Input Points Flat H&S Panels Transformations

Fig. 4.7: Upstream stepped panel normal relocation (left), effect of the joint relocation (right).
compared to the simple relocation of input points the joint relocation scheme is more expensive
in implementation.

4.3.2 Relocation range

In the perturbation scheme the variation of the hull form only is a derived quantity since actual control
of the hull variation is attained by the variation of the hull source densities. Nevertheless, like in all
engineering tasks topological and geometrical constraints have to be imposed in order to achieve valid
and meaningful hull form variations.
In the joint panel relocation scheme the hull panel mesh is gradually moved in surface normal direction
which can be considered a topology-preserving form variation. 16 Topological bounds are imposed to
avoid prohibited panel interactions, like panels approaching each other to close or even crossing each
other which may cause singular value problems. Moreover, topological bounds are required to avoid
panel relocations violating the initial panel alignment or violating symmetry conditions, for instance
panels crossing the midship plane. Another focus is the preservation of topological hull features like
the flat of side and bottom or the preservation of other self contained hull regions.
In this context geometrical bounds are imposed which introduce a permissible range for each panel
relocation. Typically, in ship application the hull surface is forced to lie within the bounding box
defined by the maximum length, the maximum (half-)breadth and the maximum draft.
The relocation bounds are applied to restrict the hull panel normal relocations at the first stage of the
joint relocation scheme. Whereas, the subsequent reconstruction of a coherent input point mesh by
vector averaging does not demand any further constraints handling. The reason is, that the averaging
procedure forces the hull panels to stay within the envelope of the preceding panel normal relocation.
16

presuming the magnitudes of the panel relocations are, in a perturbation sense, small and that the curvature distribution
of convex-concave hull areas, like the transition from bulbous-bow to the bow, is moderate

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

59

4 Sensitivity analysis
Input Points Flat H&S Panels Transformations

Fig. 4.8: Downstream stepped panel normal relocation (left), effect of the joint relocation (right).
Via the perturbation approach, see equation (4.9), both the hull panel source density variations and the
derived hull panel normal relocations are linearly linked to each other. Hence, the bounds imposed
on the hull panel relocations can be directly mapped into the range of the hull panel source density
variations.
All subsequent definitions and derivations in this section are related to the SHIPFLOW coordinate
system and to the ships initial floating position at rest. 17 Recalling the SHIPFLOW coordinate system,
a right-handed Cartesian system is defined with the origin located at the fore perpendicular in the
undisturbed free surface level, x pointing downstream and z pointing up. Since, in the present work it
is focused on laterally symmetrical ship hulls it always is referred to the positive, i.e., to the starboard

side. Furthermore, all definitions and derivations are related to the panel control points CT . The reason
is to keep the scheme unique but comprehensive and to comply with the boundary conditions of the
panel method which, likewise, are satisfied at the panel control points.
Topological bounds

Topological bounds are imposed to avoid prohibited panel interactions, to avoid panel relocations violating the initial panel alignment or violating the symmetry condition at midships plane and to preserve
topological hull features like the flat of side and flat of bottom.
Midship plane

The present work is focused on laterally symmetrical flow cases. Since only one half of the hull
is considered 18 in the flow simulation special precautions are required to avoid panels crossing or
17

Since the perturbation is applied to the dynamic ship condition considering the wetted hull panelization under the wavy
free surface and the effective floating position of the advancing ship including trim and sinkage, the hull virtually is
transformed back to its initial floating position at rest.
18 the starboard side

60

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

4.3 Panel relocation scheme

Fig. 4.9: SHIPFLOW panelization of a ferry hull, perspective view.


Input Points Flat H&S Panels Transformations

2
1

-1
0

1
-1

0
y
-1
-2

Fig. 4.10: Cut-out of the panelization showing the intersection region at the fore perpendicular FP and
the design waterline DW L (upper edge of the transition area from the bulb to the bow).
detaching the symmetry plane and panels touching the plane at obtuse angles. 19 Otherwise, in case of
laterally asymmetric flows where the hull has to be taken on the whole the panel relocation scheme can
be applied, in general, without special precautions of this kind.
19

Panels which are in close vicinity and even parallel to the symmetry plane are prohibited in potential theory methods since
the mutually induced velocities at the symmetrical counter-panel usually causes singular value problems.

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

61

4 Sensitivity analysis

Fig. 4.11: Panel overlapping at design waterline as a consequence of a homogenous panel normal relocation (left), effect of the joint relocation (right) considering symmetry at the midships
plane.
Figure 4.5 shows a generic ferry-type stern hull patch (upside down) with y-symmetry at midships
plane. To the left the corresponding panelization by 4 4 flat quadrilateral panels is given. If the
panels next to the symmetry plane start to move in their normal directions they inevitably detach or
cross the symmetry plane as illustrated by the pictures on the left hand side in figures 4.6 to 4.8. A
simple solution to prevent the panels next to the symmetry plane either from detaching or crossing the
plane during panel relocations is to simply freeze the component of their panel normal which points
perpendicular to the symmetry plane. 20 As a result the corresponding panels are forced to stay in a
plane parallel to the midship plane, thus, ensuring a closed input point mesh. Even though this approach
is easy to implement, yet, it restricts a free panel relocation.
An alternative approach, as followed in the present work, takes into account not only the panels next
to the symmetry plane but also their mirror images. In terms of the relocation scheme that means
to simply include the mirror panels across the symmetry plane in the vector averaging of the joint
relocation scheme. In this way the input point mesh is forced to remain at the symmetry plane while
the panels next to the symmetry plane are, at the same time, free to move. This particular behaviour of
20

normally the y-component of the panel normal vector

62

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

4.3 Panel relocation scheme

Fig. 4.12: Stem panels crossing the midships plane as a consequence of an inward normal relocation
of the upper bow panels (left), effect of the joint relocation (right) considering symmetry at
the midships plane.
the joint relocation scheme is illustrated by the pictures to the right in figures 4.6 to 4.8.
The above outlined procedure is a straightforward way to maintain a coherent input point mesh. However, problems arise if arbitrary hull panels intend to cross the midship plane. This case is illustrated by
means of a typical hull panelization of a ferry-hull, as presented in figure 4.9 and 4.10. Let us focus on
the intersection region of the fore perpendicular FP and the design waterline DW L (upper edge of the
transition area bulbous-bow bow). Figure 4.12 shows the special case of a hull panel relocation with
the stem panels crossing the symmetry plane (see the enlarged views at the bottom of the figure). The
application of the joint relocation scheme, alone, is not sufficient in this case in order to reconstruct a
coherent input point mesh. Apparently, another relocation constraint has to be imposed a minimum
distance between the hull panels and the midships plane needs to be preserved, too.
Specific hull features

In hull optimization distinct hull features are often considered as fixed due to requirements such as
water depth restrictions, harbour or canal restrictions, cargo handling and storage, accommodation of

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

63

4 Sensitivity analysis

Fig. 4.13: Lengthwise stepped panel normal relocation (left), effect of the joint relocation (right) considering symmetry at the midships plane.
the propulsion system, specific stern shapes with respect to propulsive efficiency etc..
Specific hull features or self contained hull regions may be frozen during the hull optimization process.
If this gets translated into the notation of the panel relocation scheme this is equivalent to suppressing
the panel normal relocations of specific hull panels. Mathematically this is achieved by employing
equality constraints
(1)

~nk p = 0

k p = 1, , Nk p
,
Nk p < NK NH

(4.21)

with the index k p indicating frozen or passive hull panels. In terms of the present implementation
those panels do not appear in the equation systems right from the beginning. Instead of equalities,
alternatively, inequality constraints
(1)

nk p max k~nk p k 0

nk p max nka max ,

(4.22)

can be employed to allow for a little leeway, i.e., small changes of the passive panels around their
initial positions. However, the absolute maximal leeway of the passive panels, defined as nk p max , is
small compared to the maximum allowed relocation of the active hull panels, nka max .

64

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

4.3 Panel relocation scheme

e
Cas

e
Cas

e
Cas

CT j( 0 )

y
(0)

(0)
k

CT
Initial panel alignment

CT k j

CT j(1)
(1)

CT k j
Panel relocation Case 1

CTk(1)

Fig. 4.14: Generic 2-D panel relocation example examining two panels arranged in a blunt angle (e.g.
the transition from FOB to the bilge radius) permitted and prohibited (crossed out) relocation cases are compared.
The benefit of allowing all hull panels to adjust freely, even though the passive panels, however, only
on a small scale, is that the overall relocation of the panel mesh promises to be more smooth since no
abrupt breaks in the relocation pattern are expected.
Panel alignment

In the present work both the CFD simulation and the vector averaging of the joint relocation scheme
depend on a regularly aligned panel mesh. Therefore, it is mandatory for the panel relocation scheme
to preserve the initial panel alignment or topology. From an application point of view this means
avoiding all events where panels approach each other too close or even crossing each other during
panel relocation.
In SHIPFLOW the hull panels are basically aligned length- and sectionwise forming a regular mesh as
required by the CFD system. A typical panelization is shown in figure 4.9. Problems with panel overlapping as a consequence of panel relocations are prevailing in case of hull regions with pronounced
convex-concave shape characteristics, for instance at the intersection region of fore perpendicular FP
and design waterline DW L, see figure 4.10. Figure 4.11 demonstrate the possible panel overlapping at
the design waterline following an outwardly directed panel normal relocation. Although, the degree

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

65

4 Sensitivity analysis
of panel overlapping in figure 4.11 is still within permissible limits a more pronounced normal relocation would cause serious problems within the mesh reconstruction scheme and, hence, needs to be
prevented beforehand.
Undesired panel collisions can be avoided by defining appropriate bounds for the panel normal relocations. Figure 4.14 outlines permitted and prohibited relocation cases by means of a simple 2-D panel
relocation example. For the sake of clarity the problem is reduced to two dimensions, but, all conclusions drawn are valid also in three dimensions. In the initial panel alignment (filled grey) two connected
panels are arranged in a blunt angle. Both panels get relocated independently in their normal directions.
Relocation case 1 and case 2 are permitted since the initial panel alignment is maintained, although,
both panels already start to detach or to cross each other, respectively. The relocations in case 3 and
case 4 (dashed) are prohibited and have to be prevented. Although the initial panel alignment is still
preserved in relocation case 3 the panel control points already approached each other below a certain
limit. In case 4 the panel alignment is reversed and, hence, the initial mesh topology is violated.
The degree of panel alignment during panel relocations can be measured by the change of the distance
vector,
(1) (1) (1)
CT k j = CT j CT k ,
(0)
in relation to the initial 21 distance vector, CT k j . It represents the vector pointing from the control
point of panel k to the control point of panel j, see figure 4.14.
On closer examination of figure 4.14 it becomes apparent that not the modified 22 (due to perturbation)
distance vector itself but rather that part of the vector in direction of the corresponding initial distance
vector may serve as a suitable measure of the degree of panel alignment. In other words, the projection
of the modified vector onto the unit initial distance vector,
c (0)
CT
kj =

(0)
CT k j
,
(0)
kCT k j k

yields, by virtue of the inner product,


(1) c (0)
CT k j CT
kj ,

the desired measure which can be formulated in terms of an inequality constraint


(1) c (0)
(0)
CT k j CT
k j PA kCT k j k 0 ,

(4.23)

with PA being a slack parameter controlling the panel alignment. PA is a user predefined parameter
(0)
to control the panel approaching in terms of the initial control point distance kCT k j k. Equation (4.23)
21
22

indicated by (0)
indicated by (1)

66

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

4.3 Panel relocation scheme


(1)

may be recasted in terms of the panel normal relocations ~nk, j in the form

0 < PA < 1

k = 1, , NK

f
j = 1, , NH , , NFS
(0)
(0)
(1)
(1)
c
(~n j ~nk ) CT k j + (1 PA) kCT k j k 0
.
NK NH

nk n j

k 6= j

kj 6= { j k}T

(4.24)

The inequality constraint (4.24) is utilized to bound the hull panel relocations in order to preserve a
coherent panel alignment, avoiding serious panel collisions and panel crossings. The parameter PA can
be estimated by analysing the most hazardous relocation case of two adjacent panels directed against
each other or are arranged in some sharp angle. Tests with ship applications conducted by the author
identified a value of PA 0.5 or slightly higher to be appropriate.
Monitoring of the panel alignment may concentrate particularly on the direct neighbourhood of a hull
panel. But, to be on the safe side and to keep the scheme comprehensive the inequality constraint (4.24)
is evaluated for all hull panels against all other hull panels, no matter how far the panels are located
from each other. Furthermore, the worst case is assumed in monitoring of two active hull panels. That
is, both panels approaching each other with the same magnitude during panel normal relocation. In
addition the panel alignment is monitored for the active hull panels against a subset (NFS
f ) of the free
23
surface panels in the close vicinity to the hull.
Geometrical bounds

Geometrical bounds are utilized to preserve a permissible range for the hull panel relocations. Typically, in ship application the hull surface is forced to lie within the bounding box defined by the
maximum length, the maximum beam and the maximum draft. In addition, in order to comply with
the perturbation approach the panel relocations are required to be small with respect to a reference
length.
Maximum panel relocation

As described above the present optimization approach utilizes gradients in the optimization. The gradients are evaluated by means of explicit linear relations. A perturbation scheme is applied to linearize
the boundary value problem. The basic assumption is that the magnitude of the perturbation is small.
Since the hull panel normal relocations are selected 24 as perturbation quantities their relocation magnitude need to be reasonably bounded.
Similar to the treatment of the passive hull panels the maximum allowed normal relocation of the
active 25 hull panels can be formulated in terms of an inequality constraint
(1)

nk max k~nk k 0

k = 1, , NK .

(4.25)

23

The free surface panels are passive in this context.


in addition to the hull panel source densities
25 active hull panels are indicated by index k
24

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

67

4 Sensitivity analysis
Different choices of the maximum panel normal relocation nk max are conceivable. Following the definitions of the perturbation magnitudes in section 4.2.2 the maximum panel relocation may be simply
related to a characteristic length L (e.g. the length between perpendiculars or the waterline length)
0 < ~nmax 1 ,

nk max = nmax = ~nmax L

(4.26)

controlled by a parameter ~nmax . Alternatively, the maximum panel relocation may be related to a
typical hull panel size, e.g., the average square roots of the hull panel areas Si
NH

nk max = nmax = ~nmax

i=1

Si
0 < ~nmax 1 .

NH

(4.27)

Which of the two expressions is eventually utilized to define the maximum normal panel relocation
nmax is a practical decision by the user. For the perturbation scheme it is of minor importance as long
as the maximum panel relocation is kept small.
Experiences gained from the optimization applications suggest to determine a perturbation magnitude
n according to (4.19) which at the same time acts as the maximum panel normal relocation, i.e.,
nmax = n.
Longitudinal bounds

In ship optimization the ship length most often is given as a fixed or at least only slightly modifiable
quantity. Accordingly, the longitudinal panel relocation, here in x-direction, may be bounded at both
bow and stern.
For the ships longitudinal direction this requires for the panel relocations
(0)

(1)

xBow CTk x + nk x xStern

k = 1, , NK ,

(4.28)

with CTk x indicating the x-component of the k-th panel control point. xBow and xStern are the fore- and
aftmost hull point, respectively. Equation (4.28) can be recast as follows:
(0)

(1)

CTk x + nk x (xBow Bow LPP ) 0

(0)
(1)
(CTk x + nk x ) + xStern + Stern LPP

0 Bow 1 ,

0 Stern 1 .

(4.29)
(4.30)

Introducing slack parameters Bow,Stern the longitudinal panel relocations at the bow and stern can be
variably controlled, respectively. For instance, small values of the slack parameters allow for a little
leeway, i.e., the active hull panels are permitted to move a small distance beyond the extents of the hull.
However, if the panels are to stay within the extents then the slack parameters would have to be set to
zero.
Lateral bounds

Similar to the longitudinal extension of the ship the hull often has to comply with a maximum breadth.
Accordingly, the lateral panel relocation, here in y-direction, may be bounded by the symmetry plane
and the maximum breadth.

68

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

4.4 Process flow of sensitivity analysis


For the ships lateral direction this requires for the panel relocations
(0)

(1)

yMS < CTk y + nk y yB max

k = 1, , NK ,

(4.31)

with CTk y indicating the y-component of the k-th panel control point. yB max represents the maximum
lateral hull extension and yMS the y-coordinate of the midship plane in the SHIPFLOW coordinate
system, usually yMS = 0. Equation (4.31) can be interpreted in terms of inequality constraints for the
midship plane and the maximum breadth, respectively,
B
(0)
(1)
CTk y + nk y (yMS + MS ) 0
2
B
(0)
(1)
(CTk y + nk y ) + yB max + SS 0
2

0 < MS 1 ,

(4.32)

0 SS 1 ,

(4.33)

with B denoting the ships breadth. By means of the slack parameters MS,SS the lateral panel relocations at the midship plane and the shipside can be variably controlled. For instance, a small value of SS
enables the active hull panels to move a small distance beyond the half-breadth of the hull. Otherwise,
to keep the panels exactly within the half-breadth then SS would be set to zero. However, the lower
bound associated with MS must be chosen large enough to keep the panels away from the midship
plane.
Vertical bounds

Finally, the hull design must not violate the maximum permitted draft. Like the ships breadth the draft
is usually limited. In order to avoid the upmost hull panels (at the level of the wavy free surface) to
move above a certain waterline level an upper vertical bound might be defined. The appropriate zW FS
bound corresponds to the z-coordinate of the control point of the upmost wetted hull panel typically
located in the wavy free surface level (WFS) along the hull side.
For the ships vertical direction, the z-direction in the SHIPFLOW coordinate system, this requires for
the panel relocations
(0)

(1)

zT max CTk z + nk z zW FS

k = 1, , NK ,

(4.34)

with CTk z indicating the z-component of the k-th panel control point. zW FS represents a certain waterline
above the wavy free surface level and zT max the z-coordinate of the maximum vertical extension at
keel. Equation (4.34) can be interpreted in terms of inequality constraints for the maximum vertical
extensions
(0)

(1)

CTk z + nk z (zT max Keel T ) 0

0 Keel 1 ,

(4.35)

(CTk z + nk z ) + zW FS + W FS T 0

0 W FS 1 ,

(4.36)

(0)

(1)

with T denoting the ships draft. Again using slack parameters, here Keel,W FS , the vertical panel
relocations can be variably controlled. For instance, a small value of Keel permit the active hull panels
to move a small distance below the draft of the hull. Otherwise, to keep the panels exactly within the
draft Keel would be set to zero.

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

69

(Re-)Initialization of
optimization task
(variables, constr., bounds)

Transfer matrices

Systematic hull
perturbation analysis

Initialization

4 Sensitivity analysis

Impact on hull form


(inverse perturb. solution)

Impact on flow field


(wave elevat., flow quant.)

Impact on
wave energy distribution
(free wave spectrum, RWP)

Gradients matrices

Sensitivity analysis

Systematic variation of
design variables
(flow quantity related)

Evaluation of gradients
(MoM, constraints, bounds)

Fig. 4.15: Process flow of hull perturbation and sensitivity analysis (detail magnification of the process
flow, figure 2.1).

4.4 Process flow of sensitivity analysis


The process flow of the sensitivity analysis is depicted in figure 4.15. Figure 4.15 gives a detail of the
general optimization process flow, see figure 2.1.
The sensitivity analysis is the central unit of the optimization scheme. The sensitivity analysis comprises the cause and effect chain from the variation of the optimization variables (the cause) to the
variation of the wave pattern, the free wave spectra and the wave pattern resistance (the effects). The
sensitivity analysis is based upon a preceding flow solution and wave cut analysis. In the present
scheme a converged nonlinear free surface CFD simulation by SHIPFLOW is utilized as a base flow
solution.
In a preparation step the perturbation scheme is initialized by evaluation of the transfer matrices which
serve as the basic instrument for the sensitivity analysis. The partial derivatives of the influence coefficients with respect to the hull panel normal relocations in the transfer matrices are numerically
evaluated. This is done by subsequently shifting each active hull panel by the perturbation magnitude
n, as defined by (4.19), in its panel normal direction. Then, in principle, the influence coefficients
are evaluated for the initial, undisturbed and the relocated panel position and the derivatives in the
transfer matrices are numerically determined by forward differencing. The transfer matrices comprise
the variation effect on both the hull and the free surface portion. However, at the stage when the

70

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

4.5 Impact on hull form and main form parameters


transfer matrices are evaluated the free surface panelization itself is temporarily frozen. This process
continues for all active hull panels until the transfer matrices are completely assembled. However,
since the normal relocation of a single hull panel only affects a single row and a single column in the
influence coefficient matrices a fast and memory conserving scheme was developed for the successive
construction of the transfer matrices.
Once the transfer matrices are assembled the systematic variation of the optimization variables may
start. A single optimization variable, i.e., a single hull source density, is varied at a time. The source
density variation essentially affects the neighbouring hull panels. In order to restore the hull boundary
condition 26 the adjacent panels are forced to adjust their positions, accordingly. The panel adjustment
by panel normal relocations is controlled by virtue of the transfer matrices and the inverse perturbation
relation (4.9). Hence, the variation of a single source density results, at first instance, in a normal
relocation of the neighbouring hull panels. This local deformation of the panel mesh again causes a
variation of the hull form parameters.
The locally confined panel normal relocation in turn has an impact on the flow quantities in terms
of the flow potential and the wave elevation, as will be seen in section 4.6. The impact on the free
surface elevation determined from (4.79) considering the hull panel normal relocations from (4.9), the
perturbation potentials from the direct perturbation relation (4.13) and the transfer matrices (4.80).
Details are given in section 4.6. Prior to further processing the perturbation wave elevations at the free
surface collocations points are interpolated in order to yield the perturbation wave cut at a predefined
offtrack position.
Finally, the variation of the free wave spectra and the impact on the wave pattern resistance is determined by a longitudinal wave cut analysis of the perturbation wave cut. The reader is referred to
section 4.7 for details. The cause and effect chain is thus completed up to the objective function of
optimization, i.e., the wave pattern resistance.
Principally, the above procedure is to be repeated for all hull panel source densities which are considered as optimization variables. However, the formulation of the perturbation relations allows for a joint
treatment of all optimization hull panels in one single step. Hence, the above steps need to be processed
only once in order to perform the whole sensitivity analysis. By far the most resource intensive part is
the evaluation of the transfer matrices in the preceding preparation step.
For further processing the sensitivity relations are utilized in terms of gradients. The gradients are
numerically evaluated by means of forward differences. Details of the gradient evaluation are given in
section 4.8.

4.5 Impact on hull form and main form parameters


The present optimization scheme is based upon locally confined hull variations. Hence, the variation
of a single hull source density is supposed to result in a locally restricted hull panel normal relocation.
Whereas, the global hull variation results from a superposition of all local variation effects with the
objective of minimizing the wave pattern resistance. In this section an approach is outlined how to
confine the hull variations locally.
26

in a perturbation sense

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

71

4 Sensitivity analysis

............
......... 4
......... 3
......... 2
......... 1
Level 0

Fig. 4.16: Schematic sketch of the hull panel influence segment of a single source density variation at
level 0, edge cases (left) and central location (right).
In ship hull optimization the main hull form parameters, generally, are considered as fixed given quantities or are kept within lower and upper bounds. The reason is to maintain, for instance, a certain
predefined deadweight, to maintain a specific floating position or to comply with stability requirements. Normally, many criteria are in the catalog of requirements. In the present optimization scheme
the main hull form parameters, like the displacement, the waterline area, the center of buoyancy and
the longitudinal center of flotation are considered.
The present approach is based on discretized hull boundaries. A hull panel normal relocation naturally
causes a variation of the hull form parameters. This implicates the evaluation of the form parameter
variations on a discrete basis, too. This practice coincides with the linear perturbation approach. 27
The hull optimization is carried out for the effective floating position of the advancing ship including
trim and sinkage. Consistently, the form parameter variations are evaluated for the advancing ship
considering the wetted hull panelization under the wavy free surface and the effect of trim and sinkage. 28
Again, all definitions and derivations in this section are related to the SHIPFLOW coordinate system
(compare section 4.3.2).
Hull form

Since the present optimization scheme is based upon locally confined hull variations the variation of
a single hull source density is supposed to result in a locally restricted hull panel normal relocation.
Mathematically this is expressed by the inverse perturbation relation (4.9). A single source density
variation in (4.9) is realized by a single nonzero entry in the right hand side vector ~(1) with otherwise
zero components. In other words, it is searched for a hull panel normal relocation which complies with
a single nonzero source density variation, but zero variations at the remaining hull panels. However,
27

Comparisons with results from continuous integration of the form parameters confirm the validity of evaluating the form
parameter variations on a discrete basis.
28 This practice somewhat differs from the common approach to assign the form parameters and hydrostatic coefficients to
the ships floating position at rest. However, both approaches are realizable within the present scheme, alternatively.

72

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

4.5 Impact on hull form and main form parameters


this constraint is too strict for the hull panel relocation as tests with different ship applications and
different variation scenarios revealed. Mathematically the inverse relation (4.9) provides a unique
solution for this case. However, from an application point of view the resulting hull panel relocation
can not be considered local since undesired, awkward panel movement can be observed at considerable
distance from the origin of the source density variation. The undesired panel movement mainly occurs
at the (flat of) bottom and (flat of) side of the ship where the panels are almost aligned with the main
flow direction. Due to their vanishing angle of incidence the disturbance velocities are small or even
vanishing at those panels. This, normally, is accompanied by a comparatively small source density
value. As a consequence, these hull panels are highly sensitive to any, yet remote, perturbation of the
flow boundaries. In order to overcome this problem a slight modification of the above procedure should
be introduced:
A local panel relocation can be easily attained by restricting the solution of (4.9) solely to the hull
panels in the neighbourhood of the source density variation. This is a somewhat practical approach and
it proves to be very effective in application. Moreover, this approach is substantiated by the special
property of a single spacial source singularity whose main influence effect is locally confined, since
the velocity induced by the source singularity vanishes rapidly with 1/r3 , r being the distance to the
singularity.
In figure 4.16 the influence segment of the source density variation is sketched. For example an influence segment up to level 1 means that solely the hull panels up to level 1 are shifted according to the
inverse relation (4.9) in order to adjust a given source density variation at level 0 and, simultaneously,
preserving the source densities at the adjacent hull panels up to level 1. At the same time all panels
exterior to the segment are untouched.
From an application point of view the influence segment must be locally confined, thus, allowing a
locally traceable variation effect. A too large influence segment may also cause considerable trouble
in the optimization when preserving the geometrical constraints. On the other hand the influence
segment must be large enough so that the variation effects are of sufficient magnitude to be utilized
in terms of gradients. From a systematic optimization study an influence segment up to level 1
(or 2) was identified as a suitable compromise. In the special case of an influence segment which
is completely reduced to level 0, i.e., to the variation hull panel itself, there is effectively no more
difference between varying the hull source densities or relocating the hull panels directly. Optimization
studies with a degenerated influence segment (to level 0) showed a tendency to produce somewhat
bumpy shape variations.
Finally, the global hull variation results from a superposition of all local variation effects with the
objective of minimizing the wave pattern resistance.

Displacement volume

The effective displacement volume may be represented in terms of the perturbation notation as the sum
of the basic displacement V (0) and its variation V 29
V = V (0) + V .
29

(4.37)

alternatively, the displacement volume often is indicated by

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

73

4 Sensitivity analysis
The displacement variation resulting from the normal relocation of the single hull panel k can be calculated on a discrete basis by the panel area Sk multiplied by the magnitude of the panel normal relo(1)
cation nk
(1)

(1)

Vk = Sk ~nk nk = Sk nk 1

k = 1, , NK .

(4.38)

Evidently, the displacement variation due to the simultaneous normal relocation of all active hull panels
is given by the sum of the Vk
NK

V =

Vk ,

(4.39)

k=1

Generally, the displacement variation is bounded or even set to zero. In the present optimization scheme
the variation of the derived hull form parameters is controlled by variable lower and upper bounds. The
bounds for the displacement variation (4.39) can be expressed in terms of two inequality constraints
Vlow Vinit + V 0
Vup Vinit V 0

0 < Vlow 1 ,
0 < Vup 1 ,

(4.40)
(4.41)

with Vlow/up being parameters controlling the displacement variation and Vinit is the initial displacement.
In the present scheme Vinit is selected as the displacement of the effective submerged hull portion under
the wavy free surface of the advancing ship. Vinit is deduced from a CFD simulation for the initial hull
considering the wavy free surface and the effect of trim and sinkage.
Center of buoyancy

The effective longitudinal (x-direction) and vertical (z-direction) center of buoyancy may be repre(0)
(0)
sented in terms of the perturbation notation by the basic location LCB , VCB varied by LCB , VCB
(0)

LCB = LCB + LCB ,


(0)

VCB = VCB + VCB .

(4.42)

The variation of the center of buoyancy due to the normal relocation of the single hull panel k can be
calculated on a discrete basis by
1 (1)
(0)
LCB k = CTk x + nk x ,
2
1 (1)
(0)
VCB k = CTk z + nk z
2

k = 1, , NK .

(4.43)

The variation of the center of buoyancy LCB , VCB due to the simultaneous normal relocation of all
active hull panels can be expressed in the form
(0)

LCB = (LCB LCB init ) (LCB LCB init ) ,


(0)

VCB = (VCB VCB init ) (VCB VCB init ) ,

74

(4.44)

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

4.5 Impact on hull form and main form parameters


with LCB init indicating the initial center of buoyancy. The center of buoyancy in (4.44) is evaluated
by
NK

(0)

LCB V (0) + LCB k Vk


k=1

LCB =

V (0) + V
NK

(0)

VCB V (0) + VCB k Vk


k=1

VCB =

V (0) + V

(4.45)

Apparently, in (4.45) both the numerator and the denominator are linear functions of the perturbation
quantities. Thus, the rational functions LCB and VCB depend in a nonlinear manner on the perturbation
quantities. Unfortunately, this violates the basic assumptions of the linear perturbation approach. In
order to avoid rational expressions in the optimization scheme the 1st -order statical moments may be
used instead whose variations are given by virtue of the relations (4.44) by
(0)

MSL = (LCB LCB init )V (LCB LCB init )V (0) ,


(0)

MSV = (VCB VCB init )V (VCB VCB init )V (0) .

(4.46)

On closer examination of above expressions, considering the relations


NK

(0)

LCB V = LCB V (0) + LCB k Vk ,


k=1

(0)

NK

VCB V = VCB V (0) + VCB k Vk ,


k=1

equation (4.46) can be rewritten in the form


NK

MSL =

LCB k Vk LCB init V ,

k=1
NK

MSV =

VCB k Vk VCB init V .

(4.47)

k=1

The sum terms in (4.47) are still of higher order in the perturbation quantities. In order to comply with
the linear perturbation assumption equation (4.47) requires a consistent linearization. The perturbation
terms may be temporarily written by introducing the small perturbation parameter . Then applying
(4.38), (4.39) and (4.43) to (4.47) yields
NK

MSL =

1 (1)
(0)
(CTk x + 2 nk x ) Vk LCB init
k=1
NK

MSV =

(0)

(CTk z

k=1

1 (1)
+ nk z ) Vk VCB init
2

NK

Vk ,

k=1
NK

Vk .

(4.48)

k=1

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

75

4 Sensitivity analysis
Finally, the terms in parenthesis in (4.48) are expanded and higher order contributions associated with
O(2 ) are neglected. 30 This results in linear expressions for the variation of the 1st -order statical
moments of buoyancy
NK

(0)

Vk LCB init V + O(2 ) ,

(0)

Vk VCB init V + O(2 ) .

CTk x

MSL =

k=1
NK

CTk z

MSV =

k=1

(4.49)

Presuming the hull perturbations to be small, the variation of the 1st -order statical moments of buoyancy
(4.49) can be considered an equivalent to the variation of the center of buoyancy (4.44). 31
According to the displacement variation variable lower and upper bounds for the variation of the 1st order statical moments of buoyancy (4.49) can be defined in terms of inequality constraints
MSL low MSL init + MSL 0

0 < MSL low 1 ,

(4.50)

MSV low MSV init + MSV 0

0 < MSV low 1 ,

(4.52)

MSL up MSL init MSL 0

MSV up MSV init MSV 0

0 < MSL up 1 ,

0 < MSV up 1 ,

(4.51)
(4.53)

with MSL low/up and MSV low/up being parameters controlling the variation of the statical moments. MSL init
and MSV init are the initial 1st -order statical moments of buoyancy which are deduced from the effective
submerged hull portion under the wavy free surface of the advancing ship, with
MSL init = LCB init Vinit 6= 0 ,

MSV init = VCB init Vinit 6= 0 .

Waterline area

The effective waterline area may be represented in terms of the perturbation notation as the sum of the
(0)
basic waterline area AW L and its variation AW L
(0)

AW L = AW L + AW L .

(4.54)

In the present context the waterline area is associated with the effective wavy free surface contour along
the hull side of the advancing ship. That is, the waterline area is put up by horizontal cuts through the
hull along the effective wavy free surface contour at the hull side. The variation of the waterline area
due to the normal relocation of a single active waterline panel k is deduced on a discrete basis from the
length of the upper panel edge multiplied by the panel normal relocation. Both the upper panel edge
and the panel normal relocation are, beforehand, projected onto the horizontal xy-plane. The upper
panel edge in the SHIPFLOW frame of reference is for each hull panel defined by the upper panel
30
31

in a linear perturbation sense


A study, conducted by the author, with realistic perturbation cases revealed a very close correlation of the variation of the
center of buoyancy and the variation of the 1st -order statical moments.

76

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

4.5 Impact on hull form and main form parameters

corner points (CP) number #2 and #3. 32 Thus, the variation of the waterline area due to the normal
relocation of the single active waterline panel k can be calculated on a discrete basis by



(0)
(0) (1)


AW L k =
k = 1, , NK NK ,
(4.55)
CP 3 k xy CP 2 k xy nk knk xy k
with

(0)

(0)


CP3 k x CP2 k x



(0)

(0)
(0)
(0)
CP

CP
=


CP
CP
2 k xy
3 k xy
2 k y
3 k y



0

and



nk x




knk xy k = nk y .


0

The variation of the waterline area AW L due to the simultaneous normal relocation of all active waterline panels is given by the sum of the AW L k
AW L =

NK

AW L k

k =1

(4.56)

Variable bounds for the variation of the waterline area (4.56) can be expressed in terms of inequality
constraints
AW L low AW L init + AWL 0

0 < AW L low 1 ,

AW L up AW L init AWL 0

0 < AW L up 1 ,

(4.57)
(4.58)

with AW L low/up being the parameters controlling the variation of the waterline area. AW L init is the initial
waterline area which is deduced from the effective wavy free surface contour along the hull side of the
advancing ship.
Longitudinal center of flotation

The effective longitudinal center of flotation may be represented in terms of the perturbation notation
(0)
by the basic location LCF varied by LCF
(0)

LCF = LCF + LCF .

(4.59)

The variation of the longitudinal center of flotation can be derived according to the center of buoyancy.
The variation of the longitudinal center of flotation due to the normal relocation of the single active
waterline panel k can be calculated on a discrete basis by
1 (1)
(0)
LCF k = Pk + nk x
2

k = 1, , NK NK .

(4.60)

with
(0)

Pk =
32

1
(0)
(0)
(CP2 k x +CP3 k x ) .
2

(4.61)

For an observer looking from outside in positive y-direction into the hull the panel corner points are for each hull panel
oriented clockwise with the first corner point located in the lower left panel corner.

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

77

4 Sensitivity analysis
The variation of the center of flotation LCF due to the simultaneous normal relocation of all active
waterline panels can be expressed in the form
(0)

LCF = (LCF LCF init ) (LCF LCF init ) ,

(4.62)

with CF init indicating the initial center of flotation. The center of flotation in (4.62) is determined by
(0)

NK

(0)

LCF AW L + LCF k AW L k
k =1
(0)
AW L + AWL

LCF =

(4.63)

Similar to the derivation of the center of buoyancy, in order to avoid rational expressions, not the center
of flotation (4.63) is used in the optimization scheme but the 1st -order statical moment whose variation
is given by virtue of relation (4.62) as
(0)

(0)

MSLCF = (LCF LCF init ) AW L (LCF LCF init ) AW L .

(4.64)

On closer examination of the above expression and considering the relation


LCF AW L =

(0) (0)
LCF AW L +

NK

LCF k

k =1

AW L k ,

equation (4.64) can be rewritten in the form


MSLCF =

NK

LCF k

k =1

AW L k LCF init AW L .

(4.65)

The sum in (4.65) is still of higher order in the perturbation quantities. In order to comply with the linear
perturbation assumption equation (4.65) requires a consistent linearization. The perturbation terms may
be temporarily expressed by introduction of the small perturbation parameter . Then applying (4.55),
(4.56) and (4.60) to (4.65) yields
MSLCF =

NK

(0)

(Pk

k =1

1 (1)
+ nk x ) AW L k LCF init
2

NK

AW L k

k =1

(4.66)

Finally, the term in parenthesis in (4.66) is expanded and higher order contributions associate with
O(2 ) are neglected. 33 This results in a linear expression for the variation of the 1st -order statical
moment of flotation
MSLCF =

NK

(0)

Pk

k =1

AW L k LCF init AW L + O(2 ) .

(4.67)

Variable bounds for the variation of the 1st -order statical moment of flotation (4.67) can be defined in
terms of inequality constraints
MSLCF low MSLCF init + MSLCF 0
MSLCF up MSLCF init MSLCF 0

33

0 < MSLCF low 1 ,


0 < MSLCF up 1 ,

(4.68)
(4.69)

in a linear perturbation sense

78

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

4.6 Impact on wave elevation and flow quantities

Fig. 4.17: Outward normal shift of a panel segment at the bow (to the right) of the Wigley hull (top
side view). Predicted change of the hull source density distribution by the direct perturbation
approach (middle) compared to the result of a SHIPFLOW recomputation (bottom).
with MSLCF low/up being the parameters controlling the variation of the statical moment. MSLCF init indicates the initial 1st -order statical moment of flotation which is deduced from the effective wavy free
surface area of the advancing ship, with
MSLCF init = LCF init AW L init 6= 0 .

4.6 Impact on wave elevation and flow quantities


The variation of a single optimization variable, i.e., a single hull source density, results in a locally
confined hull form variation, as discussed in section 4.5. The local hull form variation in turn has an
impact on the flow quantities in terms of the flow potential and the wave elevation.
Source density distribution

In the direct perturbation approach 34 (4.13) it is searched for a modification of the hull and, particularly,
of the free surface source density distribution subject to a preceding hull form variation. The preceding
hull form variation again is the result of a variation of the optimization variables at the first instance of
the perturbation scheme.
Figure 4.17 exemplarily illustrates the effect of an imaginary outward normal relocation of a 3 3
panel segment 35 at the bow of the standard Wigley hull featuring a double symmetric mathematical
hull form with a rectangular center line contour (see section 6.2 for details of the Wigley hull). The
panel perturbation is based upon a preceding free surface flow solution at a Froude number of 0.3 (the
34
35

opposite to the inverse formulation


stepped relocation of a 3 3 panel segment with the maximum outward shift at the middle panel

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

79

4 Sensitivity analysis
free surface is not pictured). A comparison, on a qualitative basis, of the predicted change of the hull
source density distribution by the direct perturbation relation (4.13) with the result of a SHIPFLOW
recomputation shows an excellent agreement. In figure 4.17 it can be observed that at the shifted
panel segment itself the magnitudes of the panel source densities decreased, whereas in the immediate
vicinity they increased.
Wave elevation and wave cut

The normalized 36 free surface wave elevation is determined by the linearized dynamic free surface
boundary condition (3.8)
"

#
 2  2  2


1 2

+
+
.
(x, y) = Fn 1 +
+
+
2
2
x
y
z
x x y y z z
Here the SHIPFLOW notation is used in which is the flow potential of the previous iteration step 37
whereas corresponds to the current flow solution, e.g. the converged solution.
The effective wave elevation may be represented in terms of the perturbation notation as the sum of the
basic wave elevation (0) and its variation
(x, y) = (0) (x, y) + (x, y) .

(4.70)

The basic wave elevation (0) in (4.70) corresponds to the latest flow solution which is determined by
a nonlinear free surface CFD simulation. This flow solution is considered fully converged with respect
to the wave formation, the source density distribution and the freely adjusted ships floating position.
The free surface domain is discretized by Rankine panels. The boundary condition is satisfied in
a discrete sense at the free surface panel collocation points. Accordingly, the wave elevations are
evaluated in a discrete sense with the wave elevation at collocation point i given by (3.12)
i =


1 2
Fn 1 + 2x i + 2y i + 2z i 2 (x i x i + y i y i + z i z i )
2

i = NH + 1, , NFS ,

with x = /x etc. and NH < i NFS for the free surface panels. The velocities according to the
previous SHIPFLOW iteration step (1) 38 in (3.12) are determined from
NFS

x i = Ux i =

j=1
NFS

y i = Uy i =

(1) (1)
j +U y

Yi j

j=1
NF S

z i = Uz i =

(1) (1)
j +U x

Xi j

(1) (1)
j +U z

Zi j

j=1

i = NH + 1, , NFS .

(4.71)

36

by LPP
In the 1st step of a nonlinear free surface iteration is selected as the double-body flow or simply as the undisturbed flow
potential.
38 the last but one SHIPFLOW iteration step related to the converged solution

37

80

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

4.6 Impact on wave elevation and flow quantities


The hull perturbation is carried out around the base flow solution which corresponds to the last converged SHIPFLOW solution (0) . Hence, the effective velocities at the free surface collocation points
are composed of the basic velocities and the perturbation effects
NF S

x i = ux i =

Xi j j +U x ,

j=1
NF S

y i = uy i =

Yi j j +U y ,

j=1
NFS

z i = uz i =

Zi j j +U z

i = NH + 1, , NFS .

j=1

(4.72)

The influence coefficients Xi j Zi j in (4.72) are pure geometrical quantities depending on the geometry, the relative position and the orientation of the panels. The influence coefficients are affected by the
(1)
perturbation, in particular by the panel normal relocations nk . Since this dependency is implicit the
influence coefficients are expanded in Taylor series which are truncated after the linear terms
(0)

NK

(0)

Xi j = Xi j +

(1)

nk

Xi j

nk

k=1

Yi j =

(0)
Yi j +

NK

k=1

Zi j =

(0)
Zi j +

NK

k=1

+ O(2 ) ,

(0)

(1)
nk

Yi j

nk

+ O(2 ) ,

(0)

(1)
nk

Zi j

nk

+ O(2 )

i, j = NH + 1, , NFS ,

(4.73)

where the perturbation terms are expressed by the small perturbation parameter . The effective source
density distribution is composed of the base solution (4.5) and the source density variation which is
determined by the solution of the direct perturbation relation (4.13)
(0)

(1)

j = j + j

+ O(2 )

j = 1, , NFS .

(4.74)

Now, if the relations (4.73) and (4.74) are introduced to the free surface velocities (4.72) one gets

NF S

x i =

(0)
Xi j +

j=1
NF S

j=1

k=1

(0) (0)
Xi j j +

y i = ,
z i =

NK

(1)
nk

NFS

j=1

(0) ! 

Xi j

nk

(0)

(1)

j + j

(0) (1)
Xi j j +

NF S

j=1

(0)
j

NK

k=1

+U x

(1)
nk

(0)

Xi j

nk

NF S

j=1

(1)
j

NK

k=1

i = NH + 1, , NFS .

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

(0)

(1)
nk

Xi j

nk

+U x ,

(4.75)

81

4 Sensitivity analysis
In the sense of a consistent linearization, higher order contributions of O(2 ) in the perturbation parameter are neglected in (4.75) which yields

NFS

x i =

NFS

(0) (0)

Xi j j +

j=1
NFS

y i =

j=1
NFS

z i =

j=1

(0) (0)
Yi j j +

(0) (0)
Zi j j +

(0) (1)

Xi j j

NFS

(0)

NK

(0)

j=1

j=1

k=1

NF S

NFS

NK

j=1
NFS

j=1

(0) (1)
Yi j j +

(0) (1)
Zi j j +

j=1
NFS

j=1

(0)
j

(0)
j

Xi j

(1)

nk

k=1

k=1

+U x + O(2 ) ,

(0)

(1)
nk

NK

nk
Yi j

nk

+U y + O(2 ) ,

(4.76)

(0)

(1)
nk

Zi j

nk

+U z + O(2 )

i = NH + 1, , NFS .

Finally, the expressions for the free surface velocities from the previous, i.e., the next to last
SHIPFLOW iteration (4.71) and the velocities according to the current, converged solution comprising
the perturbation effects (4.76) are introduced to the linearized dynamic free surface boundary condition
(3.12). After sorting in orders of this yields for the 0th -order the basic wave elevation according to
(3.12) where the flow potential = (0) equals the base flow potential according to the current, converged solution before perturbation. If (3.12) is formulated in terms of the free surface velocities the
basic wave elevation follows as
(0)

i =


i
1 2h
(0)
(0)
(0)
Fn 1 +Ux2i +Uy2i +Uz2i 2 Ux i ux i +Uy i uy i +Uz i uz i
2

i = NH + 1, , NFS . (4.77)

The 1st -order variation of the wave elevation at the i-th free surface collocation point results in

i =

Fn2

"

NFS

Ux i

j=1

(0) (1)
Xi j j +

NF S

Uy i

j=1

(0)
j

NFS

NK

k=1

(1)
nk

(0) !

Xi j

nk

(0) !

NK

(0) (1)
(0)
(1) Yi j
Yi j j + j nk nk
j=1
j=1
k=1

NF S

Uz i

NF S

j=1

(0) (1)
Zi j j +

NFS

j=1

(0)
j

NK

k=1

(1)
nk

(0) !#

Zi j

nk

i = NH + 1, , NFS .

(4.78)

Equation (4.78) can be written in compact matrix notation


h 

~ x i X(0) ~ (1) + TX (0) ~n(1) +
~ = Fn2 U


~ y i Y(0) ~ (1) + TY (0) ~n(1) +
U

i
~ z i Z(0) ~ (1) + TZ(0) ~n(1)
U

[N f s ] =

82

[N f s ] ([N f s NFS ] [NFS ] + [N f s NK ] [NK ]) ,

(4.79)

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

4.6 Impact on wave elevation and flow quantities


with N f s being the number of free surface panels excluding the hull panels. TX TZ in (4.79) are
transfer matrices for the variation of the free surface wave elevation with

TX
TY
TZ

(0)

(0)

(0)

(0)
TX ik

(0)
TY ik

(0)
TZ ik

NF S

j=1
NFS

j=1
NFS

j=1

(0)

(0)
j

Xi j

nk

(0)

(0)
j

Yi j

nk
(0)

(0)
j

Zi j

nk

,
i = NH + 1, , NFS
.
k = 1, , NK

(4.80)

All the transfer matrices, TX TZ and T in equation (4.8), are evaluated simultaneously in the preparation step where the perturbation scheme is initialized, see section 4.4.
As already mentioned, the basic quantities indicated by (0) in the above relations correspond to the last
flow solution which is determined by a nonlinear free surface iteration utilizing SHIPFLOW. This flow
solution can be considered fully converged, particularly, with respect to the wave formation so that the
maximum wave change at convergence is at least one order less than the expected variations due to the
hull perturbation. To be on the safe side, for the wave elevation a convergence criterium 39 of 107
to 1010 is selected in the CFD simulation. An advantageous side effect of the small convergence criterium is that the actual nonlinear free surface boundary conditions, (3.6) and (3.7), can be considered
fully satisfied on a discrete, numerical basis with
= (0) . Therefore, the variation of the free surface
wave elevation as an effect of the hull perturbation takes place, practically, around the actual nonlinear
free surface.
In practical applications the free surface domain needs to extend far enough up- and downstream and
sidewise in order to allow for a proper wave cut analysis. At the same time the panelization has to be
dense enough to capture much of the wave information. This in turn causes large equations systems
which are costly both in computation time and in memory consumption. This gives rise to reduce
the portion of the free surface domain considered in the evaluation of the variation impact on the
free surface source densities and wave elevations. As already addressed in section 4.2.1, a validation
study conducted by the author revealed that a truncation 40 of the free surface panelization sidewise
at approximated 2/3 (or even 1/2) of the regular 41 lateral extension is acceptable due to a negligible
influence on the results as long as the perturbations are kept small.
Prior to further processing, by means of a longitudinal wave cut analysis, the discrete wave elevations including the variation effects from (4.79) are interpolated laterally and longitudinally at the free
surface collocation points in order to yield a longitudinal wave cut at a predefined lateral position,
compare section 3.2.2. At first the wave elevations are interpolated laterally in y-direction for constant
x-positions by means of a piecewise cubic spline interpolation. The curve mesh, thus constructed,
then is interpolated longitudinally in x-direction for a constant lateral wave cut position. In the present
scheme a linear interpolation is utilized longitudinally. A comparison study applying experimental
results indicated this interpolation to be stable, also compare Janson (1997).
39

depending on the application case


The underlying CFD simulation by SHIPFLOW still includes the whole free surface domain in the solution.
41 at minimum covering the Kelvin wedge

40

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

83

4 Sensitivity analysis

# 538 (90% LPP)

VS

# 252 (40% LPP)


# 76 (10% LPP)
# 68 (10% LPP)
Fig. 4.18: Panelization of the Wigley hull. Source density perturbations at selected panels (for the sake
of clarity only the panels at the starboard side are marked).
The effective longitudinal wave cut including the effect of variation may be represented in terms of the
perturbation notation as the sum of the basic wave cut (0) and its variation
(x, y) = (0) (x, y) + (x, y) .

(4.81)

Formally the perturbation relation (4.81) is equivalent to relation (4.70), except that (4.81) applies to a
constant transverse cutting plane or offtrack position y = const..

4.7 Impact on free wave spectra and wave pattern resistance


The variation of the free wave spectra and the impact on the wave pattern resistance is determined by
a wave cut analysis of the longitudinal wave cut, including the variation effects. The cause and effect
chain is thus completed up to the objective function of optimization, the wave pattern resistance.
Free wave spectra

The sine (odd) and cosine (even) components of the free wave spectrum, in non-dimensional form, are
determined as functions of the component wave direction from the weighted Fourier transforms of a
longitudinal wave cut according to (3.35)
s
[SW FT (s) cos(u y) +CW FT (s) sin(u y)] ,

s
AC () = [SW FT (s) sin(u y) CW FT (s) cos(u y)] .

AS () =

84

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

4.7 Impact on free wave spectra and wave pattern resistance


prediction ~ single restart step | | | fully converged solution | |
Wave pattern resist. change CWP 1000 [-]

Wave elevation change / LPP 1000 [-]

prediction ~ single restart step | | | fully converged solution | |


0.1

0.05

-0.05

-0.1

-0.15
0

-0.02

-0.04

-0.06

x / LPP [-]

prediction ~ single restart step | | | fully converged solution | |


Free wave spectra change AC 1000 [-]

Free wave spectra change AS 1000 [-]

prediction ~ single restart step | | | fully converged solution | |


2

-1

-2

20
40
60
Component wave direction []

-1

-2

20
40
60
Component wave direction []

20
40
60
Component wave direction []

Fig. 4.19: Improvement of the wave making characteristics of the Wigley hull at Fn = 0.3. The result
of the first sub-optimization loop of an optimization process, considering all but the center
line hull source densities as free variables, is shown.
The weighted Fourier transforms of a longitudinal wave cut are given by virtue of (3.34)

SW FT (s) =

ZxE p

s2 1

s2 1

ZxE

CW FT (s) =

(x, y) sin(s x) dx +

Z p

xE

(x, y) cos(s x) dx +

s2 1 e
(x, y) sin(s x) dx ,

Z p

xE

s2 1 e
(x, y) cos(s x) dx .

In practical applications, due to limited computer resources (CFD) or restrictions due to tank wall
reflections (EFD), longitudinal wave cuts (x, y) need to be truncated at a finite distance downstream
of the hull. In order to retain the wave energy contained within the wave pattern downstream of the
truncation point a special truncation correction is applied. An analytical function is introduced e
(x, y)
and fitted to the tail end of the wave cut signal up to the truncation point. For details see section 3.2.3.
So as to keep the following derivation clear and straightforward it may be temporarily assumed that
the wave cut signal continues far downstream avoiding the need for a truncation correction. Hence,

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

85

4 Sensitivity analysis
prediction ~ single restart step | | | fully converged solution | |
Wave pattern resist. change CWP 1000 [-]

Wave elevation change / LPP 1000 [-]

prediction ~ single restart step | | | fully converged solution | |

0.0004
0.0002
0
-0.0002

-0.0004

-0.00005

-0.0001

-0.00015

-0.0002
0

20
40
60
Component wave direction []

x / LPP [-]

prediction ~ single restart step | | | fully converged solution | |


Free wave spectra change AC 1000 [-]

Free wave spectra change AS 1000 [-]

prediction ~ single restart step | | | fully converged solution | |


0.006
0.004
0.002
0
-0.002
-0.004
-0.006
0

20
40
60
Component wave direction []

0.006
0.004
0.002
0
-0.002
-0.004
-0.006
0

20
40
60
Component wave direction []

Fig. 4.20: Change of the wave pattern characteristic due to variation of one single hull panel source
density in its optimum gradient direction until the perturbation bound is reached. Here panel
# 68, located above keel level 10% LPP downstream of the bow, is selected.
without loss of generality, the weighted Fourier transforms simplify to

SW FT (s) =

Z p

s2 1 (x, y) sin(s x) dx ,

s2 1 (x, y) cos(s x) dx .

CW FT (s) =

(4.82)

If the longitudinal wave cut, including the effect of variation (4.81), is introduced to (4.82) then the
(0)
(0)
weighted Fourier transforms follow as the sum of the basic transforms SW FT and CW FT and their variations SW FT and CW FT

(0)

SW FT (s) = SW FT (s) + SW FT (s) ,


(0)

CW FT (s) = CW FT (s) + CW FT (s) ,

86

(4.83)

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

4.7 Impact on free wave spectra and wave pattern resistance


prediction ~ single restart step | | | fully converged solution | |
Wave pattern resist. change CWP 1000 [-]

Wave elevation change / LPP 1000 [-]

prediction ~ single restart step | | | fully converged solution | |


0.02

0.01

-0.01

-0.002

-0.004

-0.006

-0.008

-0.01
0

x / LPP [-]

prediction ~ single restart step | | | fully converged solution | |


Free wave spectra change AC 1000 [-]

Free wave spectra change AS 1000 [-]

prediction ~ single restart step | | | fully converged solution | |

0.2

-0.2

-0.4
0

20
40
60
Component wave direction []

20
40
60
Component wave direction []

0.2

-0.2

-0.4
0

20
40
60
Component wave direction []

Fig. 4.21: Change of the wave pattern characteristic due to variation of the hull panel source density
# 76, located directly beneath the wavy free surface level 10% LPP downstream of the bow.
with SW FT and CW FT determined from
SW FT (s) =

Z p

s2 1 (x, y) sin(s x) dx ,

s2 1 (x, y) cos(s x) dx .

CW FT (s) =

(4.84)

(0)

(0)

Accordingly, this yields for the free wave spectra the sum of the basic wave spectra AS and AC and
their variations AS and AC
(0)

AS () = AS () + AS () ,
(0)

AC () = AC () + AC () .

(4.85)

with AS and AC determined from


s
[ SW FT (s) cos(u y) + CW FT (s) sin(u y)] ,

s
AC () = [ SW FT (s) sin(u y) CW FT (s) cos(u y)] .

AS () =

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

(4.86)

87

4 Sensitivity analysis
prediction ~ single restart step | | | fully converged solution | |
Wave pattern resist. change CWP 1000 [-]

Wave elevation change / LPP 1000 [-]

prediction ~ single restart step | | | fully converged solution | |


0.001
0.0005
0
-0.0005
-0.001
-0.0015
0

0.0004

0.0002

-0.0002

-0.0004
0

x / LPP [-]

prediction ~ single restart step | | | fully converged solution | |


Free wave spectra change AC 1000 [-]

Free wave spectra change AS 1000 [-]

prediction ~ single restart step | | | fully converged solution | |


0.03
0.02
0.01
0
-0.01
-0.02
-0.03
0

20
40
60
Component wave direction []

20
40
60
Component wave direction []

0.03
0.02
0.01
0
-0.01
-0.02
-0.03
0

20
40
60
Component wave direction []

Fig. 4.22: Change of the wave pattern characteristic due to variation of the hull panel source density
# 252, located directly beneath the wavy free surface level 40% LPP downstream of the bow.

Wave pattern resistance

The wave pattern resistance serves as the objective function of the hydrodynamic hull optimization.
The non-dimensional form of the wave pattern resistance, in a notation going back to Havelock, is
given by (3.30)

RW P =

Z/2
0


A2S () + AC2 () cos3 () d .

Introducing the free wave spectra from (4.85) to the above expression yields

Z/2
2 
2 
(0)
(0)
RW P =
AS () + AS () + AC () + AC ()
cos3 () d .

(4.87)

This expression for RW P contains square terms of the variations of the free wave spectra. In a consistent
linearization the square terms, A2S () and AC2 (), are dropped from (4.87) since they are of higher
order and small in terms of the perturbation parameter . However, as will be discussed in section 5.2,

88

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

4.8 Evaluation of gradients


prediction ~ single restart step | | | fully converged solution | |
Wave pattern resist. change CWP 1000 [-]

Wave elevation change / LPP 1000 [-]

prediction ~ single restart step | | | fully converged solution | |


0.02
0.01
0
-0.01

-0.02

-0.03
0

0.002
0
-0.002
-0.004
-0.006
-0.008
-0.01
0

x / LPP [-]

prediction ~ single restart step | | | fully converged solution | |


Free wave spectra change AC 1000 [-]

Free wave spectra change AS 1000 [-]

prediction ~ single restart step | | | fully converged solution | |


0.6
0.4
0.2
0
-0.2
-0.4
-0.6
0

20
40
60
Component wave direction []

0.6
0.4
0.2
0
-0.2
-0.4
-0.6

20
40
60
Component wave direction []

20
40
60
Component wave direction []

Fig. 4.23: Change of the wave pattern characteristic due to variation of the hull panel source density
# 538, located directly beneath the wavy free surface level 10% LPP upstream of the stern.
it is preferred to set-up and to solve the objective functional in a least-square approach. Hence, it is
advantageous to keep expression (4.87) in its original form which inherently is of least-square type.
On closer examination of (4.87) it becomes apparent that a reduction of RW P can be achieved only by
a minimization of the free wave spectra, weighted by cos3/2 (), i.e.,
h

i2 
(0)
3/2
min
AS () + AS () cos ()
h

i2 
(0)
3/2
min
AC () + AC () cos ()

and
for all .

(4.88)

The objective function of optimization will be defined and derived in detail in section 5.2.

4.8 Evaluation of gradients


The present optimization scheme is based on gradients. The impact of the variation of the optimization
variables, i.e., the hull panel source densities, on the hull form, on the wave pattern, the free wave
spectra and the wave pattern resistance is measured in terms of gradients. Two principal types of

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

89

Wave pattern resist. change DCWP / CWP (0) [%]

4 Sensitivity analysis

0.03
0.02
0.01
0
-0.01
-0.02
-0.03

-3

-2

-1
0
1
Source density ratio Ds68 / Ds68 max [-]

Fig. 4.24: Change of the wave pattern resistance due to variation of the single hull panel source density
# 68 (an abscissa value of +1 means that the source density variation is pushed to its upper
bound allowed in optimization).
gradients characterize the sensitivity analysis. One is related to the perturbation approach, in particular,
to the set-up of the transfer matrices
(0)
~X1
,
~n

(4.89)

and the second specifies the sensitivities of the objective function and the constraints with respect to
the optimization variables, i.e., the hull source densities
(0)
~X2
,
~

(4.90)

(0)

with ~X1,2 being place holders for any of the system state quantities.
In principle, the partial derivatives of the state variables can be evaluated either analytically or numerically. In the analytical approach the partial derivatives of the system state quantities need to be
evaluated in closed analytical form. This rapidly becomes a highly complex task since the boundary
value problem depends in a complicated nonlinear manner on the optimization variables and on the
hull panel locations. Yet the influence coefficients Ai j , Xi j , Zi j are nonlinear functions of the panel
geometry. A derivation of the analytical relations may be based on the Hess & Smith relations. 42 However, the present scheme is based on a numerical flow solution. This calls for a numerical evaluation
of the partial derivatives of the influence coefficients with respect to the hull panel normal relocations
(4.89). Therefore, the gradient (4.89) is approximated here by forward differences in a straightforward
way, see further below.
42

Kraus (1989) formulated closed analytical expressions for the gradients of the influence coefficients for 2-D flows about
NACA profiles omitting the free surface. In his work the hull variation was conducted by shifting the panel input points
in lateral direction.

90

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

Wave pattern resist. change DCWP / CWP (0) [%]

4.8 Evaluation of gradients

0.5

-0.5

-1

-1.5
-3

-2

-1
0
1
Source density ratio Ds76 / Ds76 max [-]

Fig. 4.25: Change of the wave pattern resistance due to variation of hull panel source density # 76.
In their initial form the objective function and the constraints are related in complicated nonlinear
manner to the optimization variables. However, in the present scheme the cause and effect chain is
linearly interlinked since the nonlinear boundary value problem is transferred to explicit equations
depending linearly on the optimization variables. Hence, the gradients of the objective function and
the constraints with respect to the optimization variables (4.90) simply turn into the slopes of the
corresponding straight-line equations. The slopes can be determined simply by forward differences
(4.92) too.
In the limiting case when the perturbation quantities approach zero the forward differences become the
partial derivatives
(0)

(0)
(0)
~X1 ({CT}
~X
+ ~n) ~X1 ({CT}(0) ) ~X1
=
= 1
lim
~n0
~n
~n
~n
(0)
(0)
(0)
(0)
~X2 (~ + ~) ~X2(~ ) ~X2
~X
lim
=
= 2 .
~
~
~
~0

(4.91)
(4.92)

The gradients of the influence coefficients with respect to the hull panel normal relocations (4.91) are
utilized in the evaluation of the transfer matrices, (4.8) and (4.80). The gradients are approximated by
forward differences according to (4.91) by successively shifting each active hull panel by the perturbation magnitude n, as defined by (4.19), in its panel normal direction. Since the normal relocation of
a single hull panel only affects a single row and a single column in each influence coefficient matrix
a fast and memory conserving scheme was developed for the successive construction of the transfer
matrices. Applying a marching technique, the evaluation of the gradients due to each single hull panel
normal relocation only requires the computation of two rows and two columns for each influence coefficient matrix. Nevertheless, the transfer matrices and the coefficient matrix A(0) (utilized in the direct
perturbation relation (4.13)) need to be processed and temporarily stored in memory. In practical applications relatively large free surface domains are handled which require an appropriate number of

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

91

Wave pattern resist. change DCWP / CWP (0) [%]

4 Sensitivity analysis

0.04

0.02

-0.02

-0.04
-3

-2

-1
0
1
2
Source density ratio Ds252 / Ds252 max [-]

Fig. 4.26: Change of the wave pattern resistance due to variation of hull panel source density # 252.
panels. This in turn causes large equations systems with a large coefficient matrix. To overcome the
problem the portion of the free surface domain which is considered in the evaluation of the transfer
matrices and in the solution of the direct perturbation relation (4.13) might be reduced, as outlined in
section 4.6.
The second type of gradients (4.92) characterizes the sensitivities of the objective function and the
constraints with respect to the optimization variables. These gradients constitute the essential result of
the sensitivity analysis and are directly utilized by the optimization scheme, see chapter 5. They are
evaluated at each stage of the cause and effect chain.
Figures 4.19 to 4.27 show a representative perturbation study. The gradients of the objective function,
i.e., the free wave spectra and the wave pattern resistance, are investigated by means of a numerical
study for the standard Wigley hull at Fn = 0.3. Figure 4.19 shows the optimization result 43 of the first
sub-optimization loop of an optimization process. All but the center line hull source densities were
selected as free variables in the optimization. Figures 4.20 to 4.23 show the change of the wave pattern
characteristic due to the variation of selected panel source densities along the Wigley hull, see sketch
in figure 4.18. The selected source density values were pushed in their respective optimum gradient
directions up to their individual variation bounds allowed in optimization. The variation bounds were
defined so that the hull panel normal relocations did not exceed the perturbation magnitude n as
computed from (4.19), with ~n = 0.1. Finally, in figures 4.24 to 4.27 the change of the wave pattern
resistance due to the variation of the selected source density values is represented. Each source density
variation was normalized by its variation bound (for instance, an abscissa value of +1 means that the
source density variation was pushed to its upper bound allowed in the optimization).
The perturbation relations derived in this chapter were applied to predict the gradients of the objective
function. The predicted gradients (prediction) were compared to results from direct SHIPFLOW
43

The longitudinal wave cut at the offtrack position y/LPP = 0.25 was considered for the wave cut analysis.

92

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

Wave pattern resist. change DCWP / CWP (0) [%]

4.8 Evaluation of gradients

0.5

-0.5

-1

-3

-2

-1
0
1
2
Source density ratio Ds538 / Ds538 max [-]

Fig. 4.27: Change of the wave pattern resistance due to variation of hull panel source density # 538.
recomputations. 44 In the SHIPFLOW recomputations two cases were considered: a single restart step
from the converged base solution and a fully nonlinear free surface iteration until convergence. As
long as the perturbations stay in line with the perturbation bounds a good correspondence was found
for both the continuous distributions (figures 4.19 to 4.23) and the scalar wave pattern resistance values
(figures 4.24 to 4.27). The results are excellent and prove the validity of the method.

44

In this case, the SHIPFLOW runs were conducted directly with the individually relocated, gaping hull panels as provided
by the perturbation scheme. No joint panel relocation was applied.

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

93

4 Sensitivity analysis

94

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

5 Improvement of the wave-making


characteristics
The developed hydrodynamic optimization scheme aims at an improvement of the wavemaking characteristics and a reduction of the wave (pattern) resistance by a favourable
variation of the ship hull form. The optimization scheme utilizes gradient information of
both the objective function and the constraints with respect to the optimization variables.
The definition and evaluation of the gradients was given in chapter 4. The present chapter
focuses on the set up and the solution of the constrained minimization problem. An outline
of the method for the adaptation of the new, improved hull form completes the chapter.
The developed optimization scheme aims at an improvement of the wave-making characteristics and
a reduction of the wave (pattern) resistance of ship hulls. As already outlined it differs from related
schemes in four substantial aspects:
Hull variations are controlled by a flow related quantity, i.e., the hull panel source densities.
Neither, are essential prerequisites to the hull representation and its characteristic features nor
to the hull variation method imposed. Moreover, the optimization variables constitute a link
between the hull geometry and their hydrodynamic properties.
The objective function of optimization is directly related to the steady ship wave systems in
terms of free wave spectra and the wave pattern resistance as determined by a longitudinal wave
cut analysis. This improves the system identification since waves can be influenced and reduced
more systematically and with higher control.
The hull optimization is carried out for the most realistic case of the advancing ship in its effective
floating position including trim and sinkage. Variations are directly applied to the wetted hull
portion under (and slightly above) the actual wavy free surface level.
The optimization process is established in terms of an iterative marching scheme of successive
sub-optimization loops which built on top of each other. At each sub-optimization stage the
boundary value problem is simplified by a linear perturbation approach around the current base
point yielding a convex quadratic image of the solution space in a trust-region, i.e., the suboptimization space around the current base point. This simplified image is known to possess a
single minimum determined by the active constraints. Thus, fast and straightforward evaluation
of the gradients of the objective function and the constraints is enabled, allowing a simultaneous
treatment of numerous, locally acting optimization variables (of order 103 ) which introduces
great freedom to the hull variation.
The present chapter focuses on the hull improvement stage, as a part of the optimization process,
compare figure 5.1. The constrained minimization problem is discussed, the optimization functional is

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

95

Minimization of objective
function implying constraints
(iterative scheme)
Generation of
improved hull form &
prediction of gains / changes

Hull improvement

5 Improvement of the wave-making characteristics

Fig. 5.1: The hull improvement stage (detail magnification of the process flow, figure 2.1).
defined, the normal equations are derived and a special, iterative solution scheme is introduced. Finally,
the adaptation of the new, improved hull form is explained.

5.1 Constrained minimization problem


In this section a brief overview of constrained minimization problems is given with a focus on the
present minimization task. The reader is referred to the comprehensive literature to this topic for a
detailed survey. Following literature resources may serve as a starting point: Nowacki (1976, 1994),
Nowacki et al. (1995), Birk and Harries (2003), The NEOS Guide (2004), Mittelmann and Spellucci
(2004).
Engineering design problems can normally be formulated in the format of optimization problems. For
a mathematical treatment by an optimization strategy the real world is first described by a physical
model which then is transferred to an appropriate mathematical model. The latter usually is solved
by a suitable numerical algorithm. Most engineering optimizations have the format of constrained
minimization problems which can be generally stated as:
Find the vector of free optimization variables ~x = (x1 , x2 , , xN )T in the N-dimensional space,
bounded by
ximin xi ximax

i = 1, , N ,

which minimizes the objective function


!

f (~x ,~p) = min { f (~x,~p)} ,


subject to the inequality constraints
g j (~x ,~p) 0

j = 1, , NG

and the equality constraints


hk (~x ,~p) = 0

k = 1, , NH .

~x indicates the vector of free variables of the minimum design. ~p represents the vector of optimization parameters. Optimization parameters are necessary to describe the design problem but they are
kept constant during optimization. More precisely, optimization parameters are understood as passive

96

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

5.1 Constrained minimization problem


optimization variables, whereas active optimization variables are allowed to freely adjust during the
optimization. If g j (~x ,~p) 0 and hk (~x ,~p) = 0 holds true for all j and k, then the vector of free variables ~x is feasible and represents a valid design candidate. Whereas, if any of the inequality constraints
has a negative value or if just one equality constraint has a non-zero value, the design candidate is
infeasible. The set of all feasible design candidates constitutes the design space.
Engineering design problems usually are either of linear (LP) or nonlinear (NLP) programming type. A
linear programming problem exists if both the objective function and the constraints are linear functions
of the free variables. In case of an LP-problem the minimum is always governed by one or more
constraints. In nonlinear programming problems either the objective function, a constraint or both
together are nonlinear functions of the free variables. The majority of engineering design problems are
of nonlinear type. In section 5.2 the present optimization problem is discussed in this respect.

Lagrange functional formulation

In the above formulation of the constrained minimization problem the constraints are specified explicitly. Special algorithms are required to deal with explicit constraints. Search methods for instance often
implement special algorithms which adopt a certain strategy as soon as the pattern search intends to
violate a constraint. However, constrained minimization problems can be re-written as unconstrained
problems with a modified objective function. This allows a treatment of the constrained problem by
unconstrained optimization algorithms. The modification of the objective function is done by adding
penalty functions. Penalty function methods are subdivided into interior, or barrier penalty function
methods, and exterior penalty function methods. For a detailed description of penalty function methods the reader is referred to the pertinent literature, e.g., see the cross references at The NEOS Guide
(2004) and Mittelmann and Spellucci (2004). The constrained minimization problem can be recasted
in an equivalent unconstrained form in terms of a free variational problem based on the Euler-Lagrange
approach
NH

NG

k=1

j=1

F(~x,~,~) = f (~x) + k hk (~x) j g j (~x) .

(5.1)

The objective functional (5.1) still depends on the constant optimization parameters ~p, but for the
sake of clarity they are omitted in the above representation. ~ and ~ are the vectors of a priori unknown Lagrange multipliers. Hence, the constrained minimization problem has been transferred to
an unconstrained one at the cost of introducing new additional unknowns. Frequently, the inequality
constraints are transformed beforehand to equality constraints by introducing so called slack variables
~s, i.e., g j (~x) 0 is transformed to g j (~x) s2j = 0. However, the slack variables introduce even further
unknowns to the problem.
In engineering design problems the number of constraints often becomes rather large. This is especially
true if certain positional constraints have to be maintained. Then the number of additional unknowns
in terms of necessary, but undesired Lagrange multipliers and slack variables may rapidly exceed the
actual number of free optimization variables. In this case the Lagrangian approach may be ineffective
(CPU-intensive and memory consuming). For this reason the constraints handling is performed in a
different way in the present scheme, see section 5.2.

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

97

5 Improvement of the wave-making characteristics


First order necessary conditions

The first order necessary conditions for a local extremum of the objective functional (5.1) require that
the gradient equals the null vector, i.e., the partial first order derivatives with respect to all variables
vanish
! ~
~x F(~x ,~ ,~ ) =
0
!

~ F(~x ,~ ,~ ) = ~0
! ~
~ F(~x ,~ ,~ ) =
0.

(5.2)

As a further condition the free variables vector ~x at the extremum point is supposed to lie on the active
constraints. Equality constraints are always active. All inequality constraints which are inactive at the
extremum point vanish by j = 0 for j
/ active set and, thus, do not affect the result. The vector ~x
which satisfies the first order necessary conditions is called a stationary point.
Sufficient conditions are introduced in order to classify the extremum, i.e., whether it is a minimum,
a maximum or an inflection point. Generally, the sufficient conditions are associated with the second
order derivatives of the objective function 1 with respect to all variables. The sufficient condition for ~x
to be a minimum is fulfilled, provided the necessary conditions are satisfied, if the Hessian matrix of
the second order derivatives, including the mixed derivatives, is positive definite, i.e., all its eigenvalues
are positive. However, in practical application the proof of the sufficient conditions is an expensive task
and, hence, the sufficient conditions are often not checked explicitly.
Solution space

In the previous section the mathematical model of the constrained minimization problem was outlined.
However, for a proper choice of a suitable solution strategy the specific shape of the solution space is
of particular importance.
In cases where the objective function and the constraints can be expressed explicitly as functions of
one single free variable (univariate problem) or a small set of variables, a picture of the shape of the
solution space is often known a priori or can be simply deduced by sampling of the respective functions.
However, hydrodynamic optimization problems are of complicated nonlinear nature with several free
variables (multi-variate problem) and constraints and frequently with multiple competing objective
functions (multi-objective problem). Moreover, this class of problems usually represent multi-stage
calculation processes involving different analysis tool for the evaluation of the objective function(s)
and the constraints. Here it is difficult or even impossible to draw a picture of the solution space. Even
the representation of the objective function over a single free variable is too expensive in most cases
since the evaluation of the objective function even for a single design candidate may be very costly.
For the moment it might be assumed that the solution space is known or its shape can be approximated by a response surface. Of particular importance in the classification of optimization problems is
whether the shape of the solution space is convex or non-convex. If the solution space is convex both
the objective function and the constraints need to be convex functions of the free variables. Convex optimization problems are characterized by a solution space with a single global optimum and are called
1

i.e., the objective function and the constraints have to be twice differentiable in the whole solution space

98

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

5.2 Optimization functional


uni-modal. Otherwise, if the objective function or a constraint is non-convex the optimization problem
is called multi-modal. Here the feasible space is composed of one or more local optima, one of which
being the global optimum. Hydrodynamic optimization problems fall into the latter category. In recent
time several algorithms have been invented which are particularly designed to deal with multi-modal
cases. Their aim is to find the global optimum at a justifiable expense. Multi-modal optimization tasks
are start-point sensitive since the selection of the starting point significantly affects the course the algorithm takes and the quality of the optimum it finds. Normally, in hydrodynamic hull optimization
an existing, approved parent hull form is selected as starting point. Any significant improvement of
the parent hull by means of an optimization strategy indicates a successful application, independently
of the character of the optimum reached (typically a local optimum or just an improvement is encountered).
The present optimization process, as presented in section 2.3, is established in terms of an iterative
marching scheme of single successive sub-optimization loops which built on top of each other. At
each sub-optimization stage the boundary value problem is accessed by a linear perturbation approach
around the actual state of the system, i.e., the current base point 2 , see section 4.2. Both the objective
function and the constraints are formulated as quadratic expressions and are combined in a functional
formulation in terms of a convex sum. That way, the simplified convex quadratic image of the solution
space in the sub-optimization space around the current base point is attained which is known to possess
only one single minimum 3 determined by the active constraints. The perturbation approach enables a
fast and straightforward evaluation of the gradients of the objective function and the constraints with
respect to the optimization variables in the sub-optimization space. The acceleration of the evaluation
process is of particular importance in the present scheme since the number of optimization variables in
terms of hull panel source densities, naturally, tends to be comparatively large (of order 103 ).
Each sub-optimization task only renders a sub-optimization space, i.e., a confined region of the full
solution space. From the point of view of the sub-optimization task the actual solution space appears
as a simplified cut-out which is temporarily bounded, hiding away the global picture. The upcoming
sections address the set-up and the solution of the sub-optimization problem.

5.2 Optimization functional


In this section a modification of the general optimization functional (5.1) is introduced which is used in
the present optimization approach. First, a motivation for the modification of the functional is given.
The present optimization scheme is based on local hull variations which are controlled by a variation of
the hull panel source densities (the optimization variables), compare chapter 4. An average hull panelization for a potential flow CFD simulation comprises approximately 1000 panels. For a calculational
example it may be assumed that all hull panels are allowed to actively adjust their positions by panel
normal relocations. This implicates the following constraints:
The hull source density variations are bounded by a lower and an upper bound which are determined by the perturbation approach approximately 1000 constraints.
2
3

the actual state of the system is determined by a converged nonlinear free surface flow solution for the current hull variant
This is only valid for the approximation of the objective function in the vicinity of the current base point.

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

99

5 Improvement of the wave-making characteristics


Each single hull panel normal relocation is forced to stay within a lower and an upper bound
which are determined by the perturbation approach and/or topological as well as geometrical
bounds approximately 1000 constraints.
Relocation of adjacent panels in different directions or with different magnitudes cause considerable large gaps in the panel mesh, unless the gaps between adjacent panels are constrained to
a permissible range approximately 3000 constraints.
The variations of the main form parameters are bounded, e.g. displacement, center of buoyancy,
water plane area, etc. 5 to 10 constraints.
All constraints are of inequality type. If they are considered in terms of a Lagrange approach, slack
variables have to be introduced to transform the inequality to equality constraints. Hence, the objective
functional F(~x,~,~s) represents a function of the Lagrange multipliers, the slack variables and the free
optimization variables which are all considered as unknowns. Applying the first order necessary conditions for a local minimum (5.2) to the above outlined example roughly results in an equation system
with 104 linearly independent equations.
Equation systems of this size are about an order of magnitude smaller then what commonly are called
large-scale linear systems. 4 However, in the present hull optimization applications the coefficient
matrix was found to be of dense structure, frequently ill-conditioned or very close to singular. Considerable problems were initially encountered when attempting to solve this equation system on a state
of the art PC for typical hull optimization applications. Both CPU and memory consumption were far
beyond justifiable limits. 5
Apparently, a different approach is required. Even though in hydrodynamic hull optimization the
objective function and the constraints are nonlinear functions of the optimization variables, in the
present approach these relations are linearized via the perturbation approach in a sub-optimization
space around the current state of the system. Hence, the actual solution space may be realized as
a linearized image which is temporarily bounded. This gives rise to applying a linear programming
algorithm, though only in the sub-optimization space around the current system state. Nevertheless,
the application of LP was discarded for the following reasons:
The integrand in the expression for the wave pattern resistance (3.30), i.e., the sine and cosine
free wave spectra to the square, inherently is of least-square type and, hence, is predestinated for
a formulation of the optimization functional in a least-square sense.
In the present scheme a perturbation approach is introduced in order to access the inverse design
problem and to accelerate the computation of the gradients of the objective function and the
constraints with respect to the optimization variables. The perturbation approach is truncated
after the linear term. However, one may later want to include higher order contributions in the
perturbation approach. Since this would prohibit the application of an LP algorithm, a different
solution scheme is therefore better suited for extensions.
Moreover, it is straightforward to combine the objective function and the constraints in an optimization functional as weighted convex sum of quadratic least-square terms. An appropriate
4

Large-scale linear systems are encountered for instance in the simulation of ocean currents or in weather simulations and
may posses 105 to 107 unknowns.
5 For the numerical solution different decomposition algorithms of the MATHEMATICA package, Wolfram (2003), like the
Cholesky decomposition or the singular value decomposition (SVD), were applied.

100

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

5.2 Optimization functional

(- )

C (j 0 )

C j av

C j org

center

C j shift

DC j org

DC j max
active constraint

r
C j (x )

(+ )

DC j max
active constraint

inactive constraint

Fig. 5.2: Generalized inequality constraint bounded by a lower (-) and an upper (+) bound.
weight distribution is determined by an iteration scheme. In this way, the number of equations is
drastically reduced and a greater flexibility is introduced to the solution scheme.
Nevertheless, an important feature of the LP formulation also applies to the devised optimization approach: the solution space of the sub-optimization task is always convex, hence, the problem is unimodal with one single minimum which is determined by the active constraints.
In the present approach a modified version of the Lagrangian functional formulation (5.1) is used. All
terms related to equality constraints are dropped from the functional in favour of inequalities. Inequalities introduce freedom to the problem since boundary conditions do not need to be met exactly but
rather they have been released to a certain band, i.e., bounded by a lower and an upper bound. Nevertheless, if required, equality conditions maybe recovered by gradually shrinking the freedom (i.e.,
the band-width). On the other hand, by tangible releasing the bounds inequality constraints are effectively enabled. As another feature of the modified functional formulation the Lagrangian multipliers ~
have been exchanged by positive weights. The weights are treated as parameters, i.e., the optimization
functional is not an explicit function of the weights. The functional expression combines the objective
function and the (inequality) constraints in terms of a positively weighted convex sum of least-square
terms
NG

F(~x) = w0 f (~x) w j g j (~x)

w0 > 0 , w j 0 .

j=1

(5.3)

w0 may be any non-zero positive scaling factor and ~w are either positive scalar weights indicating
active constraints, or are equal to zero in case of inactive constraints. A suitable weight distribution is
determined by an iteration scheme which is described in detail in section 5.3.
The objective function f (~x) in (5.3) is directly derived from the non-dimensional expression for the
wave pattern resistance (3.30), i.e.,
f (~x) := fRW P (~x) .

(5.4)

A detailed definition is given in the upcoming section 5.2.1.


The inequality constraints g j (~x) in (5.3) are defined in general form as
(0)

g j (~x) := C j max |C j (~x) C j C j shi f t | 0

C j max > 0 for j = 1, , NG .

(5.5)

In figure 5.2 the generalized inequality constrained with lower and upper bound is sketched. C j max
(0)
being the maximum allowed tolerance or the lower and upper bound of the j-th constraint C j (~x), C j
(0)

indicates the base value and C j shi f t the center-shift of the j-th constraint. The base value C j

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

of the

101

5 Improvement of the wave-making characteristics

Generalized constraint w1 g1* [-]

-1

-2
-1

-0.5
0
0.5
1
Generalized optimization variable x / D xmax [-]

Fig. 5.3: Characteristic of the generalized constraint function subject to different weights w1 .
current sub-optimization problem is required to be inside the bounds. An initial value or an origin C j org
of the constraint may be defined. Depending on the kind of constraint C j org can either take an initial
constant value, or it may be the average point C j av of the lower and upper bound or it may be equal to
the base value of the sub-optimization problem. For a consistent treatment of all constraints within the
solution scheme the generalized inequality constraint (5.5) is related (indicated by an over-bar) to its
bound C j max . Moreover, since the optimization functional is set up and minimized in a least-square
sense, expression (5.5) is reformulated in a suitable form by an equivalent transformation
!2
(0)
C j (~x) C j C j shi f t
0
C j max > 0 for j = 1, , NG .
(5.6)
g j (~x) := 1
C j max
The constraints terms may be alternatively used in terms of penalty functions, since this terminology
better suits their actual meaning, as the following discussion of the weight distribution suggests. In
the functional expression (5.3) the objective function and the terms accounting for the (inequality)
constraints are weighted against each other. Inactive constraints are dropped from the functional by
vanishing weights. High weights for the active constraints generate a stiff problem with less freedom
for the variation, where the constraints dominate the problem at the cost of a poor optimization result.
On the other hand, small weights generate a loose problem with much leeway for the minimization but
with a tendency of constraints violation. Active constraints directly lie on a bound or are outside the
bounds


C j (~x) C(0) C j shi f t


j
C j max > 0 for j active set .
(5.7)

1


C j max

Whereas, inactive constraints stay within the permissible range




C j (~x) C(0) C j shi f t


j
C j max > 0 for j
/ active set ,

<1


C j max

102

(5.8)

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

5.2 Optimization functional


compare figure 5.2.
The influence of different weights on the characteristic of the constraint function is shown in figure 5.3.
In the simplified example it is adopted that the generalized optimization variable x itself is constraint
by a lower (1) and an upper (+1) bound with the base point and the origin located at the average
point. Hence, the generalized constraint can be expressed here simply as
#
"
2
x
w1 g1 (x) = w1 g1 (x) = w1
1
w1 0 , xmax > 0 .
xmax
Figure 5.4 illustrates the principle of application of the constraints weights within the optimization
functional. For this purpose the above example was extended by an objective term so that it features a
single objective and a single constraint term which are joined together in the generalized optimization
functional F(x) = w0 f (x) + w1 g1 (x). Since the same generalized constraint, as introduced above, is
utilized the permissible variation range is limited to 1 x/ xmax +1. For the objective f (x) term
a simple function is taken for illustration which depicts, like the wave pattern resistance, a quadratic
function in the optimization variable
"
#
2
x
w0 f (x) = w0
2 +1
w0 > 0 , xmax > 0 .
xmax
From an observation of the objective f (x) function in figure 5.4 it is apparent that the global minimum
is expected at the active upper constraint bound. But how is the bound detected by means of a de facto
unconstrained functional approach? This task is accomplished by the constraint term which works,
in this respect, in the same way as a penalty function. The amount of penalty is controlled by the
associated weight w1 . Figure 5.4 reveals how different weights w1 of the constraint term are utilized to
change the characteristic of the functional expression F(x). A too small weight (w1 = 0.5) causes an
infeasible minimum of the optimization functional at x/ xmax = +4/3, since it violates the constraint
1 x/ xmax +1, compare the lower picture in figure 5.4. On the other hand a too large weight
(w1 = 2) shifts the minimum of F(x) inside the feasible domain at x/ xmax = +2/3, as the upper
picture in figure 5.4 illustrates. Although this solution is feasible, yet it is not the seeked minimum
of the objective function f (x). Apparently, the true minimum is located at the active upper constraint
bound x/ xmax = +1 which is correctly detected by a proper selection of the weight (w1 = 1), see
middle picture in figure 5.4. This simple example reveals that the appropriate choice of the weights
is of particular importance in the present optimization scheme. Therefore, in section 5.3 an iteration
scheme is introduced for the iterative adaptation of the weight distribution. The scheme keeps track of
the active set of constraints.
Based on the preceding discussion the generalized optimization functional, finally, is obtained by introducing (5.4) and (5.6) to the functional formulation (5.3) which yields
NG

F(~x) = w0 fRW P (~x)+ w j


j=1

(0)

C j (~x) C j C j shi f t
C j max

!2

w0 > 0 , w j 0 , C j max > 0 . (5.9)

The formulation (5.9) combines the objective function and the inequality constraints in terms of a
positively weighted convex sum of quadratic least-square terms. That way a convex solution space
(of the sub-optimization task) is constituted which is characterized by a single minimum (uni-modal

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

103

5 Improvement of the wave-making characteristics

Generalized functional terms [-]

12
10
8
6
4

Minimum

2
0
-2
-1

-0.5
0
0.5
1
1.5
Generalized optimization variable x / D xmax [-]

Generalized functional terms [-]

12
10
8
6
4
2

Minimum

0
-2
-1

-0.5
0
0.5
1
1.5
Generalized optimization variable x / D xmax [-]

Generalized functional terms [-]

12
10
8
6
4
2
Minimum

0
-2
-1

-0.5
0
0.5
1
1.5
Generalized optimization variable x / D xmax [-]

Fig. 5.4: Influence of the weight w1 on the characteristic of an optimization functional by means of an
example featuring a single objective w0 f (x) and a single constraint term w1 g1 (x).

104

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

5.2 Optimization functional


problem). Moreover, the minimum is determined by the active constraints. This suggests a scheme
which keeps track of the active set of constraints (and free variables).
The first order necessary conditions for a local minimum of the optimization functional (5.9) require
that the gradient equals the null vector, i.e., the partial first order derivatives with respect to the optimization variables vanish
! ~
~x F(~x ) =
0.

(5.10)

The optimization functional (5.9) is of type of a linear least-squares problem. Hence, the application of
the first order necessary conditions (5.10) results in the so called normal equations of the linear leastsquares problem, see e.g. Press et al. (1992) and Bjrck (1996). In the upcoming sections the normal
equations are derived for the objective function and the constraints expressions.

5.2.1 Objective function

The wave pattern resistance RW P , in the non-dimensional form (3.30), serves as the objective function
of the hydrodynamic hull optimization. For instance, a normalization according to the wave pattern
resistance coefficient (3.31)
CW P =

RW P
,
0.5 Sre f k02

is unsuited as objective function since it is normalized by the wetted hull surface area Sre f which may
considerably vary in the optimization process. 6 Likewise, the ratio of wave pattern resistance to the
displacement (buoyant) force according to
RW P RW P
=
,

g
is not used as objective function in the present implementation. In the opinion of the author a practical
optimization task is the optimization of the wave-making characteristics for a specified transportation
capacity. Naturally, wave-making and wave (pattern) resistance are crucial in case of medium speed
and fast volume critical 7 cargo ships like container carriers or ferries. However, for this type of ships
transportation capacity is difficult to measure correctly in terms of displacement. Hence, the resistancedisplacement ratio in this case does not provide an appropriate measure of merit. Moreover, from a
practical implementation point of view the resistance-displacement ratio is not directly usable in the
above fractional form since both the numerator and the denominator are functions of the optimization
variables.
Hence, a different strategy is followed. The wave pattern resistance is applied as the objective function
of optimization in the non-dimensional form of (3.30)

Z/2

RW P =
A2S () + AC2 () cos3 () d ,
0

6
7

unless CW P is related to a fixed wetted hull surface area, e.g., those of the initial hull
opposite to weight critical cargo ships, e.g. bulker

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

105

5 Improvement of the wave-making characteristics


while the variations of selected main form parameters, like the displacement, are kept within bounds
by additionally defined inequality constraints, see section 5.2.2.
The free wave spectra AS and AC in (3.30) are derived in the previous chapter according to equation
(4.85) as linearized functions of the perturbation quantities. The free wave spectra can be expressed
(1)
explicitly in terms of the optimization variables by expanding AS and AC in all free variables n . This
(0)
(0)
yields the tangent plane in hyperspace attached to the basic point AS and AC
N

(0)

AS () = AS () +

n=1
N

(0)

AC () = AC () +

(0)
(1) AS ()

(0)
(1) AC ()

n=1

+ O(2 ) ,
+ O(2 ) .

(5.11)

The partial derivatives of the free wave spectra in (5.11) are evaluated numerically by forward differences as discussed in section 4.8. Introducing the expansions (5.11) to the wave pattern resistance
integral (3.30) results in 8

!2
!2
Z/2
(0)
(0)
N
N
A
()
A
()
(0)
(1)
(0)
(1)
S
C
cos3 () d . (5.12)
RW P = AS () + n
+ AC () + n

n
n
n=1
n=1
0

In view of equation (5.4), the objective function of optimization can be expressed directly in terms of
the wave pattern resistance (5.12) with

(1)

fRW P (~ ) =

Z/2"

(0)
AS () +

(0)
(1) AS ()
n
n
n=1

Z/2"

(0)
AC () +

(0)
(1) AC ()
n
n
n=1

cos

3/2

cos

3/2

#2

d +

#2

d ,

()

()

(5.13)

which is then to be minimized.


Before applying the first order necessary conditions to the objective function, the continuous integral
over the component wave directions in (5.13) is restated for numerical treatment. Since the free wave
spectra are evaluated numerically by a longitudinal wave cut analysis, see section 3.2, they are already
sampled at discrete -values in the range 0 < /2. According to section 3.2 the continuous integral
over is evaluated numerically by means of the trapezoidal rule. This yields for the objective function
(5.13)
fRWP (~(1) ) =

NM

wem

m=0
8

(0)

(0)
(1) AS m
A S m + n
n
n=1

omitting the perturbation parameter

106

!2

NM

wem

m=0

(0)

(0)
(1) AC m
AC m + n
n
n=1

!2

, (5.14)

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

5.2 Optimization functional


with the numerical integration factors w
em defined as 9
(
0.5 : for m = 0 and NM
3
w
em = cos (m )
.
1 : otherwise
In the majority of optimization applications the entire wave (pattern) resistance is minimized. Occasionally only finite components of the steady ship wave system are the objective of optimization. For
instance, one aims at a reduction of the transverse waves which may generate an adverse wake flow
for propeller and rudder operation. Furthermore, fast and high-speed vessels may be prevented from
emitting steep divergent waves which cause adverse effects in the far field. In such particular cases it
may be advantageous to focus on a specific wave range, for instance the transverse or divergent waves
alone. In the present optimization approach this is achieved in a straightforward way by simply limiting
the evaluation of the objective function (5.14) to the desired wave range, i.e., the integration in (5.13)
or (5.14) is carried out solely for the particular interesting component wave directions: min max
rather then 0 /2. 10
The first order necessary conditions for a minimum require that the partial derivatives of the optimization functional with respect to the optimization variables vanish. According to (5.10) the partial
derivatives of the objective function (5.14) with respect to the optimization variables yields
fRW P
(1)

NM

= 2

m=0
NM

w
em

wem

m=0

(0)
AS m +

(0)
(1) AS m
n
n
n=1

(0)
AC m +

(0)
(1) AC m
n
n
n=1
N

(0)

AS m
+
l
(0)

AC m
l

l = 1, , N .

(5.15)

Let the gradients matrices in (5.15) be denoted as


AS

(0)

(0)
AS nm

(0)

A
= Sm ,
n

AC

(0)

(0)
AC nm

(0)

A
= Cm
n

n = 1, , N
,
m = 0, , NM

and transfer the numerical integration factors w


em into a diagonal matrix (applying the identity matrix)

w
e0 0 0

e1 0
0 w
f

emm =
W W
m = 0, , NM .
.. . .
..
..

. .
.
.
0

w
em

Then, after interchanging the summation order, equation (5.15) can be written in compact matrix notation
(0)
~ (1) fR = AS (0) f
W AS (0) T ~(1) + AS (0) f
W ~AS +
WP

9
10

(0)
AC (0) f
W AC (0) T ~(1) + AC (0) f
W ~AC ,

(5.16)

equidistant sampling provided


Note: The wave pattern resistance, the free wave spectra and the gradients are still determined for the whole wave directions
range.

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

107

5 Improvement of the wave-making characteristics


as a system of N equations. The constant factor 2 in (5.15) might be dropped since it is considered
implicitly by the leading weighting factor w0 of the objective function, compare (5.3).
Expression (5.16) is the normal equation form of the objective function. In the upcoming section a
similar derivation is carried out for the penalty functions.

5.2.2 Penalty functions

The optimization functional (5.9) is composed of the objective function and the terms accounting for
the inequality constraints. The constraints are treated via penalty functions which introduce a certain
penalty to the objective function as soon as a constraint becomes active. The constraints or penalty
terms are formulated in a least-squares sense related to their respective maximum allowed bounds.
According to (5.6) the general form of a constraint term is expressed by
(0)

gj (~(1) ) = g j (~(1) ) =

C j (~(1) ) C j C j shi f t
C j max

!2

C j max > 0 for j = 1, , NG . (5.17)

In the previous chapter the constraints terms were linearized by means of a perturbation approach. The
constraints can be expressed explicitly in terms of the optimization variables by expanding the C j in all
(1)
(0)
free variables n . This yields the tangent plane in hyperspace attached to the basic point C j
(0)
Cj = Cj +

(0)
(1) C j
n
n
n=1
N

+ O(2 )

j = 1, , NG .

(5.18)

The partial derivatives of the constraints functions in (5.18) are evaluated numerically by forward
differences as discussed in section 4.8. If the expansion (5.18) is introduced to (5.17) the general form
of the constraint term is obtained 11
gj (~(1) ) =

(0)

(1) C j

C j max 2

n=1

C j shi f t

!2

j = 1, , NG .

(5.19)

(0)

Since the current basic value C j of the constraint function is not necessarily identical to the average
point C j av of the lower and upper constraint bound, a center point shift C j shi f t is introduced, see figure 5.2. Depending on the type of constraint function a center point shift may be required or not as will
be discussed below.
According to (5.10) the partial derivatives of the constraints terms (5.19) with respect to the optimization variables yield
gj
(1)
l
11

C j max

(0)
(1) C j
n
n
n=1
N

2
2

C j shi f t

(0)

C j

l = 1, , N
.
j = 1, , NG

(5.20)

omitting the perturbation parameter

108

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

5.2 Optimization functional


Moreover, according to the optimization functional (5.9) the sum of all constraints terms multiplied
by their respective weights is to be composed. The summation of (5.20) for all constraints terms NG
results in
! (0)
(0)
NG
NG
N
C j
gj
wj
(1) C j
l = 1, , N ,
(5.21)
w j (1) = C 2 n n C j shi f t l
j max

n=1
j=1
j=1
l

whereas the multiplication factor 2 in (5.20) is considered implicitly by the weights w j . Equation
(5.21) can be written in compact matrix notation by introducing the abbreviations 12
(0)

(0) C (0) =
C
nj

C j
1
,
C j max n

C j shi f t
~

C
shi f t C j shi f t =
C j max

n = 1, , N
,
j = 1, , NG

and transferring the weights vector ~w into a diagonal matrix form

w1 0 0

0 w2 0
W Wj j =
.. . .
.
..

. ..
.
.
0 0 wj

j = 1, , NG .

Then, after interchanging the order of summations, equation (5.21) follows in compact matrix notation
~
~ (1) ~g ~w = C
(0) W C
(0)T ~(1) C
(0) W C
shi f t ,

(5.22)

as a systems of N equations. The product ~(1)~g on the left hand side of (5.22) indicates the dyadic
product of the two vectors. 13
Expression (5.22) is the normal equation form of the constraints.
In the present optimization scheme different types of inequality constraints are utilized which are all
expressed consistently by a lower and an upper bound. Constraints are defined for selected main hull
form parameters, for the hull panel normal relocations, for the relative normal relocations (induced
gaps) of adjacent hull panels and also for the optimization variables.
Main hull form parameters

In section 4.5 the displacement, the longitudinal and vertical center of buoyancy, the waterline area
and the longitudinal center of flotation are selected as the main hull form parameters to be monitored.
The centers of buoyancy and flotation, respectively, are treated in terms of their 1st -order statical moments.
12

The over-bar indicates that the constraints functions are related to their respective bounds C j max .

(1)
(1)
(1)
g1 /1
g2 /1
gNG /1

g /(1) g /(1) gN /(1)


N NG
2
2
2
2
(1)
1
G
13 ~

~
(1) g := g j /l l j =
l j
..
..
..
..

l j
.
.
.
.

(1)
(1)
(1)
g1 /N
g2 /N
gNG /N

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

109

5 Improvement of the wave-making characteristics


All form parameter constraints are of the same type and therefore are referred to in the general form C j .
The initial form parameter values before optimization are labelled C j org C j init . Generally, neither the
(0)
initial form parameter values nor their basic points C j of the sub-optimization problems are identical
to the average points C j av of the lower and upper constraint bound. This situation is illustrated in
figure 5.2 by means of the generalized inequality constrained featuring a lower and an upper bound.
Normally, the form parameter bounds are user defined prior to the actual optimization process. An
initial origin shift C j org is indicated if, for example, one of the bounds is to be relaxed. For instance,
if the displacement should not fall below 95% of the initial displacement but may increase to 110% or
even higher.
The center-shift C j shi f t of the j-th form parameter constraint is evaluated by
(0)

(0)

C j shi f t = C j av C j = C j org + C j org C j

j form param. constr. ,

(5.23)

with the origin shift


0 < C j org 1 for j form param. constr. ,

C j org = C j org C j max

(5.24)

and the maximum allowed variation of the j-th form parameter defined by
C j max = C j max C j org

0 < C j max 1 for j form param. constr. .

(5.25)

The form parameter variations are controlled by means of the user defined parameters C j max and
C j org .
The main hull form parameters introduce five penalty terms to the optimization functional.
Hull panel normal relocations

In section 4.3.2 topological and geometrical bounds for the hull panel normal relocations are for(0)
mulated. The basic points C j of the hull panel normal relocations are defined by the zero normal
relocations, i.e., the hull panelization prior to variation. Hence, the center shift is determined by the
average point C j av of the lower and upper variation bound alone
(0)

C j shi f t = C j av C j = C j av

(0)

with C j = 0 for j hull panel norm. reloc. constr. ,

(5.26)

The lower and the upper bound for each hull panel normal relocation, respectively, is determined by
the maximum of all negative and the minimum of all positive topological and geometrical bounds
(determined according to section 4.3.2) for the corresponding hull panel normal relocation.
The hull panel normal relocations introduce NK additional penalty terms to the optimization functional.
Adjacent relative hull panel relocations

Panel normal relocations of adjacent hull panels in different directions or with different magnitudes
cause considerable gaps in the panel mesh. In section 4.3.1 the idea of a joint panel relocation scheme
by vector averaging is introduced to smooth the panel mesh and to redistribute the panel normal relocations back to the associated input point mesh. In the present work this scheme is supported by

110

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

5.2 Optimization functional

Fig. 5.5: Schematic sketch of possible occurrences of induced gaps between adjacent panels (exemplary selection)
employing additional inequality constraints which keep track of adjacent relative panel relocations in
order to ensure smooth transitions in the distorted panel mesh.
In figure 5.5 possible occurrences of induced gaps between adjacent panels are sketched. Each doublesided arrow in figure 5.5 stands for an induced gap between adjacent hull panels and, hence, demands
for an own constrained term. The constraint C j is directly derived as the relative difference of adjacent
hull panel normal relocations weighted by the inverse of the cosine of the included angle between both
panel normal directions. The weighting is introduced to account for an amplification of the induced
gaps which occurs at highly curved hull surface areas where adjacent panel normal directions are
far from being parallel. No center shift C j shi f t is applied with this type of constraint since negative
and positive relative panel relocations are bounded likewise. The maximum allowed relative panel
relocations may be related, according to
C j max = C j max n

0 < C j max 1 for j relative panel norm. reloc. constr. ,

(5.27)

to the hull panel perturbation magnitude n as defined by (4.19). The induced gaps are controlled
via the user defined parameter C j max which, typically, is selected constant for all relative hull panel
relocation constraints. A value of 10% of the maximum possible hull panel perturbation magnitude
was found to work well.
However, this type of constraint needs to be applied with caution. On the one hand its application
generates smooth transitions in the distorted panel mesh but on the other hand the variety of potential
hull form variations is reduced since adjacent panels are prohibited to relocate fully independently of
each other as might be desired by the optimization. Hence, a proper choice of the parameter C j max is
important for a good balance between both contradicting requirements. It is elaborated further on this
task in the application chapter.
Depending on the specific hull application and the selected optimization range this type of constraint
introduces roughly three to four times NK additional penalty terms to the optimization functional.
Optimization variables

The optimization variables are utilized within the perturbation approach in terms of small perturbation
parameters. In order to comply with the perturbation assumption the variation range of the optimization

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

111

5 Improvement of the wave-making characteristics


variables must be bounded. Hence, the maximum bound for the variation of the optimization variables,
i.e. the hull panel source densities, may be equated as follows
C j max = C j max

0 C j max 1 for j optim. variables constr. ,

(5.28)

being the source density perturbation magnitude as defined by (4.17). The variation bound may be
adjusted via a user defined parameter C j max which, typically, is selected constant for all variables.
Moreover, recalling the conclusions of section 4.5, the immediate effect of the variation of an optimization variable is a locally confined hull form variation in terms of hull panel normal relocations.
Above, constraints were already defined to restrict each hull panel normal relocation to a certain bound.
However, this type of constraints apply to the accumulated normal relocation of each single hull panel
due to a simultaneous variation of all optimization variables. Consequently, is makes sense to keep the
hull panel normal relocations as the effect of the variation of a single optimization variable in the same
bounds. But, different from the formulation of the constraints for the hull panel normal relocations
this type of constraint needs to be formulated for the optimization variables themselves. Accordingly,
the lower and upper bound for the optimization variables are directly determined from the linear relation between the hull source density variations and the hull panel normal relocations, relation (4.9),
and from the constraints already defined for the hull panel normal relocations. Since the variation of
a single optimization variable results in a normal relocation of a set of hull panels in the neighbourhood, a set of bounds readily arises for each optimization variable which equals the number of the
involved panel relocations. In addition, as mentioned above, the bounds to comply with the perturbation assumption (5.28) may be considered for each optimization variable. 14 Thus, the actual lower and
upper bound for each optimization variable, respectively, finally is determined by the maximum of all
negative and the minimum of all positive bounds.
(0)

The basic constraints values C j of the optimization variables are zero (no variation). Due to the
bounds imposed by the panel relocation conditions, the basic points are generally not identical to the
center points of the lower and upper constraint bounds. Hence, a center shift is introduced which equals
the average points
(0)

C j shi f t = C j av C j = C j av

(0)

with C j = 0 for j optim. variables constr. .

(5.29)

The restrictions of the optimization variables introduce N (equal to the number of variables) additional
penalty terms to the optimization functional.

5.3 Solution scheme


In the preceding sections 5.2.1 and 5.2.2 the specific objective function and the penalty terms were
derived. Applying the first order necessary conditions for a local minimum, according to (5.9), to the
specific optimization functional
! ~
~ (1) F(~(1) ) =
0.

14

(5.30)

However, in practical application it was found redundant to explicitly add the bound (5.28) for the source density perturbation magnitude.

112

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

5.3 Solution scheme


results in an equation system in the unknown optimization variables ~(1) . Introducing the contribution
of the objective function (5.16) and the contribution of the penalty terms (5.22) to condition (5.30)
yields a system of normal equations for the optimization variables ~(1) , in compact matrix notation
M ~(1) = ~B ,
with the system matrix


(0) W C
(0)T ,
M = w0 AS (0) f
W AS (0) T + AC (0) f
W AC (0) T + C

(5.31)

(5.32)

and the right hand side vector



(0)
(0) f ~ (0)
~
(0) W C
~B = C
W AS + AC (0) f
W ~AC .
shi f t w0 AS

(5.33)

The system matrix M is a square matrix of dimensions N N, symmetric, non-sparse but diagonally
dominated, full rank and positive definite. 15 Linear systems of this type can usually be solved numerically, a suitable algorithm being the Cholesky decomposition, see Press et al. (1992). However, in practical application the coefficient matrix may turn out to be badly conditioned and close to singular. This
then calls for a numerical treatment by means of singular value decomposition (SVD), see Press et al.
(1992). The present implementation of the optimization scheme is based on the MATHEMATICA
package, Wolfram (2003). For the solution of the equation system (5.31) the coefficient matrix is first
analysed and then solved by the appropriate numerical algorithm, essentially switching between the
Cholesky decomposition and the application of the so called pseudo inverse (SVD) in case of a badly
conditioned matrix.
The solution of (5.31) primarily depends on the appropriate choice of the weight function w j for the
constraints. The purpose of the weights is to distinguish active and inactive constraints. The minimum
of the sub-optimization problem is determined solely by the active constraints, inactive constraints
vanish at the minimum. The weighted constraints behave in the same way as penalty functions. The
principle of application of the constraints weights was illustrated in figure 5.4. Different weighting of
the constraints can be effectively utilized to adapt the characteristic of the optimization functional. A
too small weight generates too little penalty and, hence, causes an infeasible minimum of the optimization functional since the constraint is violated. On the other hand a too large weight generates to much
penalty with a putative minimum of the optimization functional, although inside the feasible domain,
which is not the actual minimum of the objective function. However, with an appropriate choice of the
weight the global minimum of the objective function at the active upper constraint bound is detected
correctly. The example in figure 5.4 reveals the particular importance of an appropriate choice of the
weights for the optimization outcome.
However, the appropriate weights are not known a priori. In the present approach the weights are
adapted by an iteration scheme. The scheme keeps track of the active set of constraints. In each iteration
step the normal equations are set up and solved for a new set of weights. The iteration terminates
if the process has converged to the actual minimum of the sub-optimization problem. Numerically,
the process is considered converged if the relative change of the objective function in two successive
iteration steps falls below a certain limit


f (~(1) i+1 ) f (~(1) i )
RW P

RW P
(5.34)
105 ,

(1)
i+1


fRW P (~
)

15

i.e., all eigenvalues are positive

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

113

5 Improvement of the wave-making characteristics


and if all constraints are satisfied


C j (~(1) i+1 ) C(0) C j shi f t

j

1



C j max

C j max > 0 for j = 1, , NG .

(5.35)

At convergence the active constraints lie on their bounds (= 1) and inactive constraints stay within the
permissible range (< 1).
The scheme for the iterative adaptation of the weights is derived from the criteria to the active and inactive constraints at the minimum of the sub-optimization problem. For this purposes the optimization
functional (5.9) may be recalled

!2
(1) ) C (0) C
NG
~
C
(

j
j
shi
f
t
j
F(~(1) ) = w0 fRW P (~(1) )+ w j
1 w0 > 0 , w j 0 , C j max > 0 .
C
j max
j=1
The minimum of the optimization functional is solely determined by the active constraints. According
to (5.35), the active constraints terms are identical to zero at the very minimum

!2
(0)
(1)

~
C j ( ) C j C j shi f t
!
1 = 0
w j > 0 for j active set .
(5.36)
wj
C j max
Since the inactive constraints lose weight during the iteration progress, this implies for the optimization
functional that it takes the value of the objective function at the minimum
F(~(1) ) = w0 fRW P (~(1) )

w0 > 0 ,

compare the middle picture in figure 5.4. For the inactive constraints to vanish at the minimum, their
leading weights must vanish
!

wj = 0

for j
/ active set ,

(5.37)

since, according to (5.35),


(0)

C j (~(1) ) C j C j shi f t
C j max

!2

1 6= 0

for j
/ active set .

From the conditions (5.36) and (5.37) for the active and inactive constraints, respectively, the scheme
for the iterative update of the weights may be derived as follows:
(0)

w i+1
= w ij
j

C j (~(1) i ) C j C j shi f t
C j max

!2

j = 1, , NG .

(5.38)

The iterator (5.38) is strictly satisfied both for the active and inactive constraints at convergence to
the actual minimum of the sub-optimization problem. For the active constraints the square of the
constraints terms, see parenthesis in (5.38), are referred to as the quadratic residuals. They converge
to one, according to (5.35). At the same time the weights for the inactive constraints drop to zero.
Depending on the applications it may be advantageous to alter the convergence history by exchanging

114

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

5.3 Solution scheme


the quadratic residual terms in (5.38) by their absolute values. However, this was not the case in the
present applications.
In the present optimization scheme the iterator (5.38) is applied in a somewhat modified form. It
turned out that the convergence of the process close to the minimum can be effectively accelerated by
introducing a small convergence factor cf
w i+1
= w ij
j



!2
C j (~(1) i ) C(0) C j shi f t

j


+ cf


C j max

0 cf 1 .

(5.39)

Furthermore, with a nonzero convergence factor a possible temporary zero-crossing of the residual term
during the iteration process can be efficiently intercepted, avoiding an extra inquiry. The convergence
factor may be individually adjusted for each type of constraint, yet, a universal convergence factor in the
range 0.05 cf 0.1 was found to work well. The application of the convergence factor does not affect
the inactive constraints. Otherwise, the active constraints are somewhat shifted. In other words, the
minimum is somewhat shifted inside the feasible domain. However, in practical application this effect
was found to be acceptable since it is small and does not affect the character of the optimization.
Similar to other iteration schemes a certain under-relaxation may be introduced in addition. Underrelaxation is used to stabilize the convergence process and to avoid unwanted oscillations. Introducing
under-relaxation to the iterator (5.39) yields

w i+1
= w ij + rf w ij
j



!2
C j (~(1) i ) C(0) C j shi f t


j

+ cf w ij


C j max

0 cf 1 , 0 rf 1 . (5.40)

The relaxation factor rf again may be individually adjusted for each type of constraint which was
frequently utilized in the present optimization applications.
The iteration process for the weights of the constraints is as follows:
Initial values are assigned to the weights. 16
Set-up and solution of the equation system (5.31) with the initial weights.
Update of the weights according to (5.40).
Set-up and solution of the equation system (5.31) with the updated weights.
Iteration of the latest steps until convergence is reached according to (5.34) and (5.35).
Inactive constraints which are effectively approaching zero during the iteration process are removed
from the equation system in the next step. However, if those constraints happen to become active in a
subsequent iteration step they are compelled to reenter the iteration process. In the present application
cases convergence was achieved in less then 200 (frequently about 50) iterations. The task which
consumes most of the computation time is the set-up of the normal equations for the updated constraints
weights in each iteration step.
16

normally ~w1 = 1, or the final weight vector of a preceding iteration with an appropriate substitution of the zero weights

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

115

5 Improvement of the wave-making characteristics

5.4 Hull form adaptation


The result of the iterative solution of the governing equation system (5.31) is a vector of optimization
variables hull panel source densities which characterizes a minimum of the wave pattern resistance
in the current sub-optimization.
Since a ship hull optimization results in a favourable hull shape the optimization variables must be
linked to the hull geometry. In the present scheme the hull shape modification is expressed by hull
panel normal relocations which are linked to the optimization variables by the inverse perturbation
relation (4.9). The superposition of all local effects brings about a global hull variation in terms of a
relocated hull panel mesh, see sections 4.3 and 4.5.
At this stage the joint relocation scheme is applied to construct a smooth point mesh with a well defined
topology from the relocated hull panel mesh, compare section 4.3.1. The point mesh may either be used
directly in terms of intermediate offsets for a subsequent sub-optimization task or it may be further
processed by surface interpolation or approximation and fairing techniques, see e.g. Nowacki et al.
(1998), Nowacki and Kaklis (1998) and Westgaard (2000). In the present work the computer aided
ship hull design (CASHD) tool MultiSurf from AeroHydro Inc., Letcher et al. (1999), is utilized as a
geometrical postprocessor to the optimization scheme. MultiSurf is applied in a script technique in
order to generate higher-level B-Spline hull surface representations from the point mesh data in a fully
automated procedure. MultiSurf builds on a hierarchical structure from point data to (master) curves,
from (master) curves of primarily sectional character to surfaces and features a so called relational
geometry model. MultiSurf comes with a hydrostatics module which was used in a postprocessing
stage for a detailed shape comparison of the initial and the optimized hull shapes. Moreover, the
application of a tailored CASHD tool like MultiSurf simplifies the export of the hull geometries to
other systems and/or applications via portable and freely convertible formats (like IGES).
Examples for the hull form adaptation by the joint relocation scheme as well as for the application of
CASHD techniques are given in the next chapter.

116

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

6 Applications
In this chapter the application of the optimization method to exemplary test cases is presented. A double-symmetric mathematically defined hull form the standard Wigley hull
and a state of the art twin screw ferry the FANTASTIC FantaRoRo typical of fast short-sea
shipping featuring a bulbous bow and tunnelled transom stern were selected as test cases.
Both ships are supposed to operate around a Froude number of 0.3 where the wave pattern
resistance approaches a hump and, hence, tangible improvements were expected. The optimization results are discussed in view of the ability of the present optimization scheme to
efficiently improve the wave-making characteristics.
The suggested optimization method aims at an improvement of the wave-making characteristics in
conjunction with the minimization of the wave pattern resistance. Naturally, this optimization task
concerns medium speed and fast volume critical ships like container carriers or ferries.
The wave-making characteristics of two exemplary application test cases,
the standard Wigley at Fn = 0.3 and
the FANTASTIC FantaRoRo ferry at Fn = 0.311
were optimized for the intended design speeds. Both ships operate around a Froude number of 0.3
where the wave pattern resistance approaches a hump. Hence, tangible improvements were expected.
The well known and widely used Wigley hull was selected as an academic test case. It represents a
double symmetric mathematically defined simple hull form. Hence, the Wigley hull is well suited for
a calibration of the optimization parameters by means of a systematic optimization study. As a second
and realistic application case a state of the art twin screw ferry the FANTASTIC FantaRoRo typical
of fast short-sea shipping was chosen. The ferry features a bulbous bow and a tunnelled transom stern.
The ferry hull was chosen since it was devised as an elaborate test case within the European R&D
project FANTASTIC. It was already investigated within the scope of a comprehensive optimization
study both numerically and experimentally.

6.1 Introductory comments


For a general description of the optimization process flow the reader is referred to section 2.3.
The optimization environment MinSWASH

The optimization method was implemented by the author as a MATHEMATICA based code and was
called MinSWASH (Minimum Ship Wash). The optimization environment MinSWASH basically controls the optimization process flow, it comprises the sensitivity analysis, including the evaluation of the

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

117

6 Applications
perturbation quantities up to the wave elevation and the wave cut, it features the optimization kernel
and the longitudinal wave cut method SWASH and it establishes the connection to the CFD system
SHIPFLOW. In this way the entire optimization process is fully automated. Optimization runs are initiated and controlled by user edited initialization files. The intermediate and final optimization results
as well as the process progress are documented by various data and log file output. All hull geometry is
stored both in terms of pure hull panel offsets and as interpreted NURBS-representation (Non Uniform
Rational B-Splines) in the MultiSurf model file format. A MATHEMATICA based postprocessor was
developed for the visualization of the optimization results. Several diagrams and plots presented in this
chapter are generated by means of the postprocessor. The optimization may be run, optionally, on a
Windows or Linux operating system.

Nonlinear free surface flow simulation by SHIPFLOW

The hydrodynamic analysis of the steady ship wave system was carried out with the aid of the nonlinear free surface Rankine panel module XPAN of the CFD system SHIPFLOW. SHIPFLOW solves the
underlying boundary value problem by discretization of the boundaries of the flow domain, essentially
the hull and the wavy free surface. In the first order panel method the boundaries are approximated
by flat quadrilateral panels. The nonlinear wave pattern are considered the converged solution of an
iteration process which gradually adapts the free surface elevation and accounts for the dynamic floating position. SHIPFLOW adopts the so called raised panel method, i.e., the free surface panels are
raised above the free surface collocation (control) points in conjunction with an upstream collocation
point shift. Important quantities, in the present context, that SHIPFLOW provides from nonlinear free
surface flow computations are longitudinal (and transverse) wave cuts, the flow velocities, the hull
pressure distribution, dynamic trim and sinkage and the wave resistance from pressure integration over
the wetted hull surface. Moreover, SHIPFLOW Rel. 2.8.10 provides additional, specially customized
interface capabilities like the I/O of the panel geometry, of selected influence coefficients and certain
mechanism for the SHIPFLOW execution control. For a closer description of the nonlinear free surface
Rankine panel method be referred to section 3.1.

Viscous effects and propulsion

The influence of viscous effects on the formation of the wave system is not considered since the present
optimization approach builds on potential theory based boundary element method (BEM). However,
CFD validation studies comparing computed and measured wave pattern approved the excellent ability
of nonlinear free surface Rankine panel methods to model the steady wave systems of the ship types
and speed ranges in question, compare, e.g., Raven (1996), Janson (1997), Raven and Prins (1998a,b),
Nowacki et al. (1999), Heimann (2000), Valdenazzi et al. (2003), Heimann and Harries (2003). Moreover, no propulsion devices were considered in the present optimization study. This is primarily due
to a lack of information regarding the propulsion parameters like the propeller geometry, the thrust
coefficient and the advance ratio. However, propulsion devices may be included in the optimization
scheme. SHIPFLOW/XPAN utilizes a propeller in terms of an actuator disk which is considered a
straightforward but simple model of the propeller wake flow.

118

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

6.1 Introductory comments


Grid/Mesh-dependency

A well known and extensively studied feature of CFD methods, whether viscous or inviscid, is their
dependence upon the grid or mesh on which the discretized flow equations are solved numerically.
This property also applies to Rankine panel methods. However, in case of steady flows and at medium
range Froude numbers around 0.2 to 0.4 the grid-dependence is controllable and convergence to a
grid-independent solution, normally, can be pushed to a satisfactory limit. Within the scope of the
SHIPFLOW solver grid-dependence and CFD validation studies were reported, e.g., by Janson (1997),
Harries and Schulze (1997), Schulze and Harries (1997) and Lee (2003). Preliminary to the optimization runs the author conducted grid-dependence studies with the selected application test cases with
the objective to find a good trade-off between mesh-convergence and computational costs in terms of
CPU time and memory consumption. Besides mesh-convergence further constraints to the panel mesh
resolution are imposed by the optimization scheme itself. The computational expense for the perturbation analysis and the number of optimization variables in terms of hull panel source densities increases
with increasing panel mesh resolution. Moreover, with a too dense hull panel mesh the variation effects
of a single source density variation may fall below a warranted numerical precision. In addition, the
free surface panel row next to the hull need to preserve a certain lateral size in order to avoid control
point collisions with the hull panels during panel normal relocations. On the other hand they ought to
be small enough to capture the displacement effect of the hull which in turn causes the formation of
the free surface waves. Further requirements to the free surface panelization are in particular imposed
by the longitudinal wave cut analysis. The free surface domain must extent far enough downstream to
allow for a reliable truncation correction and wide enough to the side so as to cover the Kelvin wedge.
To gather most of the shorter wavelength effects a sufficient number of free surface panels per fundamental wavelength must be placed longitudinally and only a slight stretching may be used laterally.
Longitudinal stretching of the free surface panels ought to be avoided everywhere to keep numerical
damping low. The highest wave frequency that can be resolved by the free surface panelization is defined by the Nyquist cut-off frequency which is proportional to the width of the outermost lateral panel
strip, if lateral panel stretching is used. Keeping the above precautions in mind, the identical panel
meshing was used per application test case for all flow evaluations during the optimization runs. In this
way, a correct ranking of the variants is to be expected.
A different aspect of the mesh resolution is its influence on the optimization process since both the
optimization variables and the hull shape variation inherently depend upon the hull panel mesh. This
is because frequency and length of hull changes are determined by the mesh resolution. Hence, the
optimization outcome is to some degree controlled by the hull panel mesh resolution. This must be
accepted as an immanent feature of the present optimization approach, but it applies to all numerical
hull form optimization methods. An optimization study, conducted by the author, with systematically
varied hull panelizations revealed that already a relatively coarse hull panel mesh suffice to converge
to a certain optimization trend. Moreover, excessive mesh refinement only had a marginal effect on the
optimization trend and, as already outlined above, introduced performance and precision problems to
the optimization process.
Wave cut analysis and free surface domain

Tracing the results from grid-dependency studies revealed that computed wave pattern and, consequently, the resistance calculations based on them, prove to be less sensitive to flow field discretization
than hull pressure integration. Studies to this topic reveal that pressure integration often is poorly

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

119

6 Applications
conditioned and small changes of the hull panelization may lead to serious alterations in the wave resistance, e.g. Raven (1996), Janson (1997) and Raven and Prins (1998a,b). The wave resistance from
hull pressure integration RW appeared to be especially sensitive to the panelization of the bow and stern
regions (e.g. to the panel resolution close to the forward stagnation point at a bulbous bow).
Within the present optimization approach the aim is to minimize the energy losses associated with the
wave formation along the hull. The present approach is based on longitudinal wave cut analysis which
requires, in case of flows symmetric to the ship center plane, a single wave cut parallel to the ship to
perform the whole analysis. Practical application of the longitudinal wave cut analysis mainly depends
on the lateral position of the cutting plane, the utilizable length of the wave signal until the truncation
point (the downstream edge of the computational free surface domain) and the ships speed. In order to
retain the wave energy contained within the wave pattern downstream of the truncation point wave cuts
of finite length are treated by a truncation correction. To avoid serious analysis errors and to ensure
validity of the method the cut is to stay outside a region close to the hull where local flow effects
are predominant. On the other hand, in view of the numerical damping inherently in the free surface
CFD simulation, the lateral distance needs to be small enough to capture the wave signal as best as
possible. According to the authors experience satisfactory results are obtained for medium speed and
fast monohulls at lateral cut positions in the range 0.2 y/LPP 0.3 (provided the cut extends far
enough downstream to allow for a truncation correction). This is in agreement with related studies,
e.g. Lalli et al. (2000). For the present application cases a lateral wave cut position of y/LPP = 0.25
was used throughout the optimizations. For a closer description of the wave cut analysis method be
referred to section 3.2.
Selection of optimization variables

In the present optimization approach the hull panel source densities are utilized as optimization variables, compare section 4.1. The practical selection of the hull panels to be included in the optimization
is done by switching on (and off) selected quadrilateral mesh arrays per panelization hull group. Moreover, single hull panels can be selected or dropped if appropriate. The hull panels utilized for the hull
panel normal relocation scheme must comprise at minimum all hull panels selected for optimization.
Optimization bounds

For the application test cases practical geometrical bounds were maintained in order to achieve sustainable optimization results. For instance, as a geometrical bound the bounding box enclosing the hull
surface is defined by the maximum length, the maximum half-breadth and the maximum draft of the
ship at rest. Furthermore, lower and upper bounds for selected main hull form parameters were applied. Deviating from the common practice, the bounds were evaluated for the wetted hull panelization
under the wavy free surface and the effective floating position of the advancing ship including trim and
sinkage.
Convergence criteria

The optimization process is a sequence of sub-optimization loops. Each sub-optimization task is composed of a prediction and a correction step which are alternating. As a result the prediction step provides a set of hull panel normal relocations which promises to minimize the wave pattern resistance. In
the correction step the joint relocation scheme is applied to the predicted optimal hull panel relocations

120

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

6.2 Wigley hull


in order to reconstruct a coherent input point mesh. In the second phase of the correction step the
input or offset point mesh is sent to SHIPFLOW for a validation run (improvement assessment) with
nonlinear free surface condition and free trim and sinkage. This is followed by a longitudinal wave cut
analysis which yields the free wave spectra and the wave pattern resistance. The result then serves as
the starting point for the subsequent sub-optimization tasks. For details be referred to section 2.3.
Three different criteria are utilized to terminate the optimization process. The formal convergence
criterium is given by


R s+1 R s
WP
WP
(6.1)
105 ,

s+1


RW P
correct
with s being the counter for the optimization loops. Furthermore, the process is required to converge
strictly monotonic, i.e.,
 s+1

RW P < RWs P correct .

(6.2)

A further practical criterium is to hold a maximum number of optimization loops Ns max


s Ns max .

(6.3)

Normally, the formal convergence criterium (6.1) do not become active since differences in the results
from the prediction and the correction step do not allow for a very strict convergence. However, the
convergence accuracy in (6.1) may be relaxed.
Reference coordinate system and normalization

The optimization environment MinSWASH uses the SHIPFLOW coordinate system as reference system. All numerical and graphical results presented in this chapter do refer to this system. It represents
a right-handed Cartesian coordinate system with the origin located at the fore perpendicular in the
undisturbed free surface level, x pointing downstream and z pointing upwards. Since in the present
work it is focused on laterally symmetrical ship hulls it is always referred to the positive side, i.e., to
starboard.
The optimization scheme builds on normalized representations. All geometry related quantities are
normalized by the length between perpendiculars and all velocity related quantities are normalized by
the ship speed. Further normalization is applied within the wave cut analysis, compare section 3.2.
Run-time and memory consumption

All present optimization were performed on a Linux PC, AMD Dual Athlon MP 2400+ with a CPU
speed of 2 GHz and 4 GB RAM. The computation time for a full optimization study strongly depends
on the specific application case and on the hardware resources. The present optimization runs took
from 12 hours to a couple of days. The memory requirements primarily are determined by the number
of panels used which in turn determines the size of the coefficient matrices. The present applications
required approx. 1 to 1.5 GB of RAM.

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

121

FP

0
10

optimized

10

initial

AP

6 Applications

Fig. 6.1: Ship lines of the initial and the optimized Wigley hull.

6.2 Wigley hull


The well known and widely used standard Wigley hull was selected as an optimization test case. The
hull is well suited for a calibration of the optimization parameters by means of a preliminary opti-

122

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

6.2 Wigley hull


Initial

Optimized

Fn

0.3

VS

18.26 kn

1.025 t/m3

Gains &
Changes

Optimization bounds

LPP

100 m

100 m

LOS

100 m

100 m

[+1% LPP forward FP, +1% LPP aft AP]

10 m

10.153 m

+1.53%

+2%

6.25 m

6.25 m

kept within tolerance

-4.06%

[-5%, +5%]2

m3

2776.9

LCB 1

50 m

49.775 m

-0.45%

unbounded2

VCB 1

-2.344 m

-2.347 m

-0.13%

unbounded2

AW

666.51 m2

633.42 m2

-4.96%

[-5%, +5%]2

LCF 1

50 m

50.106 m

+0.21%

unbounded2

1480 m2

1478.58 m2

-0.10%

CB

0.446

0.421

-5.61%

CP

0.668

0.590

-11.68%

CM

0.667

0.713

+6.9%

CW P

0.669

0.626

-6.43%

S 1

0.025

-0.003

-0.028

zS

-0.207 m

-0.242 m

-16.91%

KM

5.277 m

5.382 m

+2.0%

KM L

123.92 m

105.57 m

-14.81%

RW P

73.2 kN

37.1 kN

-49.38%

RW P trans

51.6 kN

21.2 kN

-58.87%

RW P div

21.6 kN

15.8 kN

-26.73%

RW

107.2 kN

49.0 kN

-54.31%

RW P /

2.624 103

1.384 103

-47.24%

RW /

3.840

103

2664.2

m3

1.829

103

-52.37%

Tab. 6.1: Results of the Wigley hull optimization.


mization study. The shape of the Wigley hull is defined in closed mathematical form, see e.g. Harries
(1998). It yields a double symmetric hull form with a rectangular center line contour and parabolic
sections and waterlines below the design waterline. The design waterline is vertically extended above
1

According to the SHIPFLOW coordinate system, LCB and LCF are measured from FP, VCB is related to DW L, and a
positive trim angle S indicates trim by the stern.
2 Note: The optimization bounds for the integral form parameters were applied to the wetted hull portion under the wavy
free surface of the advancing ship, whereas the form parameter values tabulated in column 2-4 do refer to the ship at rest.

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

123

6 Applications
initial ~ optimized | | | deviation | |

Sectional area A (x) / AM [%]

100
80
60
40
20
0
-20
0
FP

0.2

0.4

0.6
x / LPP [-]

0.8

1
AP

Fig. 6.2: Sectional areas of the initial and the optimized Wigley hull (related to the midship sectional
area AM of the initial hull). Note that the fore perpendicular FP is located at x/LPP = 0.
the water level resulting in a vertical freeboard from the design waterline to the deck, see ship lines
in figure 6.1. The theoretical block coefficient is CB = 4/9. The most widely used Wigley hull is
given by L/B = 10 and L/T = 16. For this hull good accuracy of SHIPFLOW in computing both wave
resistance and total resistance was reported by Larsson (1993).
A selection of geometrical and hydrostatical values of the Wigley hull are tabulated in table 6.1. For the
sake of a clear data representation a fictive length of LPP = 100 m was adopted. The tabulated values
were subsequently scaled according to Froudes law at the postprocessing stage. The geometry related
values in table 6.1, the ship lines, figure 6.1, the sectional area curves, figure 6.2, and the hull surface
renderings, figures 6.3 to 6.5, were generated from full surface representations. The surface representation is obtained at the postprocessing stage by interpolation of the point mesh, i.e., the coherent input
point mesh from the joint relocation scheme, using an automated B-Spline lofting procedure within the
CASHD system MultiSurf, see section 5.4. MultiSurf s hydrostatics module was applied to deduce the
metacentric heights.
A thorough optimization study was conducted by means of MinSWASH for the Wigley hull. The effect
of the optimization parameters was tested and a balanced parameter set was deduced. This includes the
provision of a suitable hull and free surface panelization.
For the optimization runs themselves a suitable panelization was generated and kept during the whole
process. The hull panelization together with a part of the wavy free surface is shown in figure 6.12.
Since the flow is symmetric to the center plane only one symmetric half of the hull and free surface
domain needed to be panelized. First order flat quadrilateral panels with a constant source density
distribution along each panel were applied both for the hull and the free surface. In total 605 panels
were employed on the hull, 55 lengthwise with a slight stretching (hyperbolic tangent) towards bow
and stern and 10 uniformly spaced panels girthwise with a single extra panel row above the still water
level. The extra panels are used to improve the convergence of the CFD simulation and to include a

124

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

6.2 Wigley hull

Fig. 6.3: Birds view of the initial (top) and the optimized Wigley hull (bottom).
part of the hull above the water level in the optimization. The hull panelization was specifically adapted
to suit both the CFD simulation and the optimization purpose. A rectangular free surface panelization
was generated which extended 0.5 LPP upstream, 2.4 LPP downstream of the stern and 1.4 LPP to the
side. 25 panels per fundamental wavelength (0 /LPP = 0.565) were placed uniformly in longitudinal
direction. Laterally 27 panels were employed with a slight stretching (hyperbolic tangent) towards the
hull. The free surface panel row next to the hull had a lateral size of 0.03 LPP. In total 4482 panels
were utilized for the free surface.
All flow simulations were conducted from the beginning (no SHIPFLOW restarts). In each suboptimization task the offset point mesh was prepared by means of the joint relocation scheme and
trimmed back to the initial floating position before the CFD analysis. The nonlinear free surface
SHIPFLOW simulations were conducted with free trim and sinkage and carried to firm convergence.
The sensitivity analysis proceeded around the converged solution. Throughout the optimization process
the initial still waterline was adopted for all SHIPFLOW computations. However, the displacement, the
waterline area and the centroids are self-adjusting within the bounds. Infinite water depth was assumed
throughout the optimizations.
For the longitudinal wave cut analysis a lateral wave cut position of y/LPP = 0.25 was used throughout the optimizations. The evaluation of the free surface elevations including the perturbation terms,
their bidirectional interpolation resulting in a longitudinal wave cut and the wave cut analysis itself

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

125

6 Applications

Fig. 6.4: Fishs view of the initial (top) and the optimized Wigley hull (bottom).

Fig. 6.5: Bottom view of the initial (top) and the optimized Wigley hull (bottom).

were conducted outside SHIPFLOW within the optimization environment MinSWASH. This way, a
consistent evaluation was ensured throughout the process. The discrete wave elevations, including the
perturbation effect, at the collocation points of the curved free surface mesh were evaluated and interpolated to yield the longitudinal wave cut. At first the wave elevations were interpolated laterally
in y-direction for constant x-positions by means of a piecewise cubic spline interpolation. The curve
mesh, thus constructed, then was interpolated longitudinally in x-direction for a constant lateral wave
cut position. Linear interpolation was utilized longitudinally. The same interpolation procedure is used

126

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

6.2 Wigley hull

Resistance ratios [%]

100
RWP

90

RWP (predict)
RW

80

RF ITTC`57
70
60
50
0

6
8
Optimization loop

10

12

Fig. 6.6: Optimization history of selected resistance components of the Wigley hull (related to the
respective values of the initial hull).

Main hull form parameter ratios [%]

102
100
98
96

LCB

94

VCB
92

AWL
LCF

90
0

6
8
Optimization loop

10

12

Fig. 6.7: Optimization history of selected main hull form parameters of the Wigley hull (related to the
respective values of the initial hull).
within SHIPFLOW to determine longitudinal wave cuts internally, compare Janson (1997). For the
truncation correction a reliable fitting length of 5/4 times the fundamental wavelength was selected
throughout the optimizations. The singularity of the correction function was allowed to self-adjust,
promising an optimal fit, compare section 3.2.3.
The perturbation magnitudes and n were selected according to section 4.2.2 with = ~n = 0.1.
The adjacent relative hull panel normal relocations were limited to 10% of the maximum possible hull
panel relocation per sub-optimization loop. This condition was introduced to synchronize the predicted

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

127

6 Applications
initial ~ optimized | | | deviation | |

Wave pattern resist. coeff. CWP 1000 [-]

2
transverse

diverging

1.5
1

0.5
0

-0.5
-1
-1.5
0

20
40
60
Component wave direction Q []

Fig. 6.8: Wave pattern resistance coefficient CW P over the component wave direction of the initial
and the optimized Wigley hull.
initial ~ optimized | | | deviation | |

Wave pattern resistance fraction [%]

100
transverse

diverging

75
50
25
0
-25

-41.5%

-49.4%

-50
0

20
40
60
Component wave direction Q []

Fig. 6.9: Wave pattern resistance fraction over of the initial and the optimized Wigley hull (related to
the total wave pattern resistance of the initial hull).
optimization result with the result of the correction step. The correction step is composed of the joint
relocation scheme which recovers a coherent input point mesh from the predicted optimal hull panel
relocations and a validation run with SHIPFLOW followed by a longitudinal wave cut analysis. If the
adjacent relative hull panel relocations were not kept within certain bounds within the optimization
then the inclination changes of the hull panels, inherently imposed by the joint relocation scheme,
were spoiling or, at worst, even reversing the optimization trend. Moreover, restricting the adjacent

128

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

6.2 Wigley hull


initial ~ optimized | | |

Wave elevation / LPP [-]

0.006
0.004
0.002
0
-0.002
-0.004

x / LPP [-]

Fig. 6.10: Longitudinal wave cut at y/LPP = 0.25 of the initial and the optimized Wigley hull.
initial ~ optimized | | |
Wave pattern resist. coeff. CWP 1000 [-]

Wave elevation / LPP [-]

0.006
0.004
0.002
0
-0.002
-0.004

initial ~ optimized | | | deviation | |

2
1.5
1
0.5
0
-0.5
-1
-1.5
0

x / LPP [-]

initial ~ optimized | | | deviation | |

0.15

0.15

0.1

0.1

Free wave spectra AC [-]

Free wave spectra AS [-]

initial ~ optimized | | | deviation | |

0.05
0
-0.05
-0.1
-0.15

20
40
60
Component wave direction []

0.05
0
-0.05
-0.1
-0.15

20
40
60
Component wave direction []

20
40
60
Component wave direction []

Fig. 6.11: Summarized results of the longitudinal wave cut (y/LPP = 0.25) analysis. Comparison of
the initial and the optimized Wigley hull.

relative hull panel relocations introduced a certain smoothness to the hull variation, however, at the
expense of a limited scope of the hull variation. The latter restraint was less serious since a sequence

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

129

6 Applications

Fig. 6.12: Panelization of the initial Wigley hull (fishs view with part of the wavy free surface).

Fig. 6.13: Wave contours of the initial (top) and the optimized Wigley hull (bottom).
of sub-optimization tasks were iterated so that all prominent (and less prominent) hull variations were
allowed to evolve. During the optimization the initial panel topology was retained by controlling the
panel alignment, i.e., limiting the maximum allowed panel normal relocations. The initial rectangular
centerline of the Wigley hull was maintained by freezing the panels attached to the centerline during
optimization. All other hull panels were included in the optimization. The local influence segment
of a single optimization variable was limited to the adjacent hull panels, i.e., up to level 1, compare
section 4.5.
Additional bounds for the maximum hull extensions and selected main hull form parameters are given
in table 6.1. For the maximal hull dimensions just slight variations were allowed. The displacement
variation was relaxed to 5% of the initial displacement. The displacement bounds were applied

130

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

6.2 Wigley hull

l0

l0 /2

Fig. 6.14: Wave contours (enhanced resolution) in close proximity to the initial (top) and to the optimized Wigley hull (bottom).
to the wetted hull portion under the wavy free surface of the advancing ship. The same as for the
displacement applies to the waterline area. The location of the center of buoyancy and the longitudinal
center of flotation were kept unbounded.
In total 11 optimization loops with alternating prediction and correction steps were performed. The
wave pattern resistance was minimized over the whole wave directions range. The process terminated
by violation of criterium (6.2). This often happens when the optimization process reaches one or more
form parameter bounds. In this case the bound for the water plane area became active. When a form
parameter bound is reached differences in the predicted and corrected results frequently become large
enough to even reverse the optimization trend in the correction step. The reason is that the panel normal
relocations and, hence, their effects decrease when approaching a bound, while the unavoidable errors
imposed by the changes of the hull panel inclinations in the joint relocation scheme are still of some
magnitude. This is considered an undesired feature of the present optimization approach which is
further discussed in the conclusion, chapter 7.
In figure 6.6 and 6.7 the optimization histories for the resistance components and the selected main hull
form parameters are presented. They are related to their respective initial values (i.e., a value of 100%

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

131

6 Applications

Fig. 6.15: Birds view of the wave pattern of the initial (port side) and the optimized Wigley hull (starboard side).
means identical to the initial value). A substantial decrease of the wave pattern resistance by 49.4% and
of the wave-pattern-resistance-displacement ratio by 47.2% was achieved, see figure 6.6. The history
of RW P for the predicted (indicated by predict) and the recomputed wave pattern resistance in the
correction step is of very similar characteristic. The wave resistance RW from hull pressure integration
taken from the correction step shows the same trend but with an offset to even larger gains. Figure 6.7
reveals the form parameter histories. Both the displacement and the water plane area first increase and
finally decrease until the water plane area reaches its lower bound. The centroids are almost unchanged
by the optimization.
In figure 6.8 the wave pattern resistance (given as coefficient CW P ) as a function of the component

132

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

6.2 Wigley hull

Fig. 6.16: Dynamic hull pressure of the initial (top) and the optimized Wigley hull (bottom).
wave direction is compared for the initial and the optimized hull. The deviation of both resistance
curves gives a distinct picture of the gains and losses for the particular wave pattern components.
Substantial gains both in the transverse and in the upper divergent wave range are apparent. A slight
increase in wave energy emerges for the wave components travelling in 40 to the course of the
ship. A remarkable result is the reduction of the wave heights of the transverse waves following the
stern in = 0 by 65%. Figure 6.9 shows a comparison of the wave pattern resistance fraction over the
component wave direction. The graph indicates the wave pattern resistance fraction integrated up to
the particular component wave direction. All values are related to the total wave pattern resistance of
the initial hull. Again the deviation between the optimized and the initial hull is given. It reveals that
already the minimization of the transverse waves up to = 35 16 contributes a reduction of 41.5% of
the wave pattern resistance. Whereas the aggregated contribution of the whole divergent wave range
amounts to another 8%. The magnitude of the slope of the deviation curve at = 0 is an indicator
of the reduction of the transverse waves following the ship. A negative slope indicates a reduction,
whereas a positive slope means an amplification of the following waves. To obtain the reduction of the
wave pattern resistance of a particular wave range one just needs to subtract the value of the deviation
curve at the end of the particular wave range from the respective starting value. This way figure 6.9 in
conjunction with figure 6.8 are particulary suited to assign the effect of locally confined hull variations
to specific wave ranges, as will be shown later in this chapter.
A summary of the optimization results is given in figure 6.10 and 6.11 in terms of the analysed wave
cut, the wave pattern resistance coefficient and the free wave spectra.
A close inspection of the wave contour plots in figure 6.13 and 6.14 as well as the birds views of the
wave pattern in figure 6.15 reveal that both the bow and the stern wave of the Wigley hull and the
transverse waves are considerably reduced, as already seen in the CW P plots in figure 6.8. Furthermore,
at the emerging fore and aft shoulder of the optimized hull 3 wave troughs are developing with a
3

The initial hull neither features shoulders nor inflection points because of the parabolic waterlines.

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

133

6 Applications
pronounced wave trough at the fore shoulder. This is reflected also by the underpressure regions in the
dynamic hull pressure distribution in figure 6.16. An explanation for this can be found by investigating
the hull form changes. Examining the sectional area curves in figure 6.2 and the waterline contours
in figure 6.1 and 6.5 shows that a considerable amount of volume is shifted both from the entrance
and the run towards the emerging fore and aft shoulders. As a consequence the waterline angles at
the entrance and the run are narrowed, whereas the waterlines at the shoulders are pushed outwards.
This in combination introduces inflection points to the waterlines both at the entrance and the run and
amplifies the flow deflections at the shoulders. Moreover, the waterline contour between the shoulders
becomes almost parallel to the center line with a slight contraction around the midship section. The
midship section is enlarged which is expressed by an increased midship section coefficient CM , whereas
the prismatic coefficient CP decreases (related to the ships floating position at rest).
Summarizing the results the following optimization achievements led to a substantial reduction of the
bow and stern waves and to a considerable reduction of the transverse waves, compare figure 6.14:
Small waterline angles at the entrance and the run considerably reduce the bow and stern wave
crests.
Emerging shoulders are located nearly equidistantly to the midship section and their distance is
exactly half the fundamental wavelength. Hence, both shoulder wave trains are opposite in phase
which results in a mutual cancelation.
Given a constraint minimization task, this is the expected result of the wave pattern resistance minimization for the Wigley hull.
Relaxing the bounds, particulary relaxing the breadth bound, even amplifies the above outlined hull
variation trend. However, the flow deflection at the shoulders eventually becomes to large which may
cause adverse viscous effects. Furthermore, the favourable interferences of waves as predicted by
potential flow theory will be less pronounced in real flow.

6.3 FantaRoRo ferry


An application case of practical relevance is that of a state of the art twin screw ferry typical of
fast short-sea shipping. The ferry features a bulbous bow and a tunnelled transom stern. The ferry
hull is well suited since it was devised as an elaborate test case within the European R&D project
FANTASTIC Functional Design and Optimization of Ship Hull Forms (EC Growth project GRD11999-10666) and extensively studied both numerically and experimentally, see Maisonneuve et al.
(2003) for an overview on FANTASTIC. The so called FantaRoRo test case was set up by the consortium to study the potential of the tools developed or improved throughout the project. Several
alternative optimization schemes were applied by various partners, for details on the FantaRoRo study
see Valdenazzi et al. (2003).
At the Technical University Berlin a sophisticated shape optimization approach has been developed
and successfully applied to the FantaRoRo test case. In this approach advanced parametric modelling
and shape manipulation applying the FRIENDSHIP-Modeler by FRIENDSHIP SYSTEMS GmbH,
CFD based hydrodynamic analysis utilizing SHIPFLOW and detailed performance assessment using
the wave cut analysis software SWASH were highly integrated and controlled via the state of the art
optimization environment modeFRONTIER by Esteco. Within this study the optimization range was

134

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

20

AP

transom

0
-

initial

FP

optimized

6.3 FantaRoRo ferry

Fig. 6.17: Ship lines of the initial and the bulbous bow optimized FantaRoRo.
limited by the consortium solely to the fore body and the bulbous bow. For details of the optimization
approach see Abt et al. (2003) and Heimann and Harries (2003).
The present optimization method aims at the same objective, i.e., the improvement of the wave-making

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

135

6 Applications
Initial

Optimized

Fn

0.311

VS

21 kn

1.025 t/m3

Gains &
Changes

Optimization bounds

122.74 m

122.74 m

130.12 m

131.17 m

+0.81%

[+7% LPP forward FP, +4% LPP aft AP]

4.235 m

5.28 m

+24.68%

+7% LPP forward FP

19.2 m

19.2 m

kept within tolerance

5.032 m

+0.278%

kept within tolerance

+0.66%

[-5%, +5%]5

LPP
LOS
LBBow

m3

5.046 m

6570.5

LCB 4

64.754 m

64.327 m

-0.66%

[-6%, +6%]5

VCB 4

-2.122 m

-2.122 m

[-6%, +6%]5

AW

1836.17 m2

1845.16 m2

+0.49%

[-5%, +5%]5

LCF 4

71.223 m

70.865 m

-0.5%

[-8%, +8%]5

2466.34 m2

2483.81 m2

+0.71%

CB

0.540

0.523

-3.15%

CP

0.606

0.588

-2.97%

CM

0.890

0.890

CW P

0.760

0.736

-3.16%

S 4

0.038

0.011

-0.027

zS

-0.282 m

-0.292 m

-3.55%

KM

9.93 m

9.898 m

-0.32%

KM L

249.44 m

255.25 m

+2.33%

RW P

124.1 kN

111.7 kN

-10%

RW P trans

95.9 kN

85.9 kN

-10.45%

RW P div

28.2 kN

25.8 kN

-8.47%

RW

168 kN

155.5 kN

-7.42%

RW P /

1.879 103

1.68 103

-10.59%

RW /

2.543

103

6613.6

m3

2.34

103

-8.03%

Tab. 6.2: Results of the bulbous bow optimization of the FantaRoRo ferry.
characteristics. However, both optimization approaches substantially differ in the way they tackle this
task. Below the capabilities of the present scheme are elaborated on the basis of two FantaRoRo
4

According to the SHIPFLOW coordinate system, LCB , LCF and the bulbous bow length LBBow are measured from FP, VCB
is related to DW L, and a positive trim angle S indicates trim by the stern.
5 Note: The optimization bounds for the integral form parameters were applied to the wetted hull portion under the wavy
free surface of the advancing ship, whereas the form parameter values tabulated in column 2-4 do refer to the ship at rest.

136

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

6.3 FantaRoRo ferry


initial ~ optimized | | | deviation | |

Sectional area A (x) / AM [%]

100
80
60
40
20
0

0
FP

0.2

0.4
0.6
x / LPP [-]

0.8

1
AP

Fig. 6.18: Sectional areas of the initial and the bulbous bow optimized FantaRoRo (related to the midship sectional area AM of the initial hull). Note that the fore perpendicular FP is located at
x/LPP = 0.
initial ~ optimized | | | deviation | |

Sectional area A (x) / AM [%]

20

15

10

-0.05

-0.025

0
FP

0.025
x / LPP [-]

0.05

0.075

0.1

Fig. 6.19: Sectional areas of the initial and the bulbous bow optimized FantaRoRo, detail magnification
of the entrance.

optimizations which are picked out as representative results in place of a thorough optimization study
which was conducted by means of MinSWASH. The first study focused on the bulbous bow and the
transition region to the entrance. The second study extended the optimization to the entire FanatRoRo
hull.

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

137

6 Applications

Resistance ratios [%]

100
90
80
70
RWP
RWP (predict)

60

RW
50

RF ITTC`57
0

6
8
10
Optimization loop

12

14

Fig. 6.20: Optimization history of selected resistance components of the bulbous bow optimized
FantaRoRo (related to the respective values of the initial hull).

Main hull form parameter ratios [%]

102
100
98
96

LCB

94

VCB
92

AWL
LCF

90
0

6
8
10
Optimization loop

12

14

Fig. 6.21: Optimization history of selected main hull form parameters of the bulbous bow optimized
FantaRoRo (related to the respective values of the initial hull).
The initial FantaRoRo does not possess a parallel midbody and the maximum section is located slightly
aft of the midship section. A selection of geometrical and hydrostatical values is tabulated in table 6.2
for the bulbous bow optimized and in table 6.3 for the entirely optimized FantaRoRo. The hull was
scaled to the adopted length in the FANTASTIC optimizations of LPP = 122.74 m. In the optimization
applications, however, all quantities were utilized in normalized form. The tabulated values were
calculated according to Froudes law at the postprocessing stage. The geometry related values in the
tables, the ship lines, see figure 6.17 and 6.26, the sectional area curves, see figure 6.18 and 6.27, and
the hull surface renderings, see figures 6.35 to 6.39, were generated automatically at the postprocessing

138

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

6.3 FantaRoRo ferry

Wave pattern resist. coeff. CWP 1000 [-]

initial ~ optimized | | | deviation | |


1.5

transverse

diverging

0.5

-0.5

20
40
60
Component wave direction Q []

Fig. 6.22: Wave pattern resistance coefficient CW P over the component wave direction of the initial
and the bulbous bow optimized FantaRoRo.
initial ~ optimized | | | deviation | |

Wave pattern resistance fraction [%]

100
transverse

75

diverging

50
25
+

+0.3%

-0.6%

-8.1%

-10.0%

-10.6% -10.0%

-25
-50
0

20
40
60
Component wave direction Q []

Fig. 6.23: Wave pattern resistance fraction over of the initial and the bulbous bow optimized
FantaRoRo (related to the total wave pattern resistance of the initial hull).
stage by means of MultiSurf from a full surface representation, compare section 6.2. Hydrostatic values
were calculated by means of MultiSurf s hydrostatics module.
For the optimization runs a suitable panelization was generated and kept during the whole process. The
hull panelization for the initial FantaRoRo together with a part of the wavy free surface is shown in
figure 6.40. Only one symmetric half of the hull and free surface domain is panelized. First order flat

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

139

6 Applications
initial ~ optimized | | |

0.006

Wave elevation / LPP [-]

0.004
0.002
0
-0.002
-0.004
-0.006
0

x / LPP [-]

Fig. 6.24: Longitudinal wave cut at y/LPP = 0.25 of the initial and the bulbous bow optimized
FantaRoRo.
initial ~ optimized | | | deviation | |

initial ~ optimized | | |
Wave pattern resist. coeff. CWP 1000 [-]

0.006

Wave elevation / LPP [-]

0.004
0.002
0
-0.002
-0.004
-0.006
0

1.5

0.5

-0.5

x / LPP [-]
initial ~ optimized | | | deviation | |

0.05

-0.05

20
40
60
Component wave direction []

initial ~ optimized | | | deviation | |

0.1

Free wave spectra AC [-]

Free wave spectra AS [-]

0.1

20
40
60
Component wave direction []

0.05

-0.05

20
40
60
Component wave direction []

Fig. 6.25: Summarized results of the longitudinal wave cut (y/LPP = 0.25) analysis. Comparison of
the initial and the bulbous bow optimized FantaRoRo.
quadrilateral panels with a constant source density distribution along each panel were applied both for
the hull and the free surface. In total 1166 panels were employed on the hull, 110 of them in the bul-

140

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

FP

10

9
8

6 7
3

20

AP

16
15
14

17

18

19

0
transom
20
-

1 2

initial

optimized

6.3 FantaRoRo ferry

Fig. 6.26: Ship lines of the initial and the entirely optimized FantaRoRo.
bous bow group (extending up to FP), 752 in the main hull group (extending from FP up to 75% LPP)
and 304 panels in the remaining group for the run. The dry transom is not panelized. This way, the
hydrostatic pressure difference due to the dry transom is implicitly considered. The bulbous bow panelization is stretched (hyperbolic tangent) longitudinally towards the bow tip and the hull panelization

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

141

6 Applications
Initial

Optimized

Fn

0.311

VS

21 kn

1.025 t/m3

Gains &
Changes

Optimization bounds

122.74 m

122.74 m

130.12 m

131.13 m

+0.78%

[+7% LPP forward FP, +4% LPP aft AP]

4.235 m

5.25 m

+23.97%

+7% LPP forward FP

19.2 m

19.222 m

+0.11%

kept within tolerance

5.032 m

5.049 m

+0.34%

kept within tolerance

-4.83%

[-5%, +5%]7

LPP
LOS
LBBow

m3

6570.5

LCB 6

64.754 m

63.596 m

-1.79%

[-6%, +6%]7

VCB 6

-2.122 m

-2.15 m

+1.32%

[-6%, +6%]7

AW

1836.17 m2

1721.16 m2

-6.26%

[-5%, +5%]7

LCF 6

71.223 m

69.846 m

-1.93%

[-8%, +8%]7

2466.34 m2

2373.88 m2

-3.75%

CB

0.540

0.5

-7.41%

CP

0.606

0.552

-8.91%

CM

0.890

0.906

+1.8%

CW P

0.760

0.695

-8.55%

S 6

0.038

-0.065

-0.103

zS

-0.282 m

-0.3 m

-6.38%

KM

9.93 m

9.752 m

-1.79%

KM L

249.44 m

222.1 m

-10.96%

RW P

124.1 kN

74.6 kN

-39.9%

RW P trans

95.9 kN

43.2 kN

-54.98%

RW P div

28.2 kN

31.4 kN

+11.40%

RW

168 kN

113.5 kN

-32.44%

RW P /

1.879 103

1.187 103

-36.85%

RW /

2.543

103

6253.3

m3

1.806

103

-29.01%

Tab. 6.3: Results of the entire hull optimization of the FantaRoRo ferry.
is slightly stretched (hyperbolic tangent) towards FP and the transom edge, respectively. On the hull
66 panel columns were utilized lengthwise with 13 almost uniformly spaced panels girthwise includ6

According to the SHIPFLOW coordinate system, LCB , LCF and the bulbous bow length LBBow are measured from FP, VCB
is related to DW L, and a positive trim angle S indicates trim by the stern.
7 Note: The optimization bounds for the integral form parameters were applied to the wetted hull portion under the wavy
free surface of the advancing ship, whereas the form parameter values tabulated in column 2-4 do refer to the ship at rest.

142

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

6.3 FantaRoRo ferry


initial ~ optimized | | | deviation | |

Sectional area A (x) / AM [%]

100
80
60
40
20
0

0
FP

0.2

0.4
0.6
x / LPP [-]

0.8

1
AP

Fig. 6.27: Sectional areas of the initial and the entirely optimized FantaRoRo (related to the midship
sectional area AM of the initial hull). Note that the fore perpendicular FP is located at
x/LPP = 0.
ing three extra panel rows above the still water level. The hull panelization was specifically adapted
to suit both the CFD simulation and the optimization purpose. A rectangular free surface panelization
was generated which extended 0.5 LPP upstream, 2.4 LPP downstream of the stern and 1.4 LPP to the
side. 25 panels per fundamental wavelength (0 /LPP = 0.609) were placed uniformly in longitudinal
direction. Laterally, 27 panels were employed with a slight stretching (hyperbolic tangent) towards the
hull. The free surface panel row next to the hull had a lateral size of 0.03 LPP . In addition a free surface
transom group was applied detaching the transom stern to account for the tangential flow separation of
the trailing edge of the transom. In other words, special transom panels were adopted in order to satisfy the Kutta condition at the transom edge. The free surface transom group introduces 300 additional
panels. In total 4674 panels were utilized for the free surface discretization.
For the initialization of the optimization a complete SHIPFLOW flow simulation was conducted with
free trim and sinkage and iterated until full convergence. In order to accelerate convergence intermediate restart runs were conducted in addition (alternating) from previously converged solutions. In
restart runs trim and sinkage usually were fixed to the latest dynamic floating position. In each suboptimization task the offset point mesh for the CFD flow simulation was prepared by means of the
joint relocation scheme. The nonlinear free surface SHIPFLOW simulations were conducted either
with or without free trim and sinkage and carried to firm convergence. The sensitivity analysis proceeded around the converged solution. Throughout the optimization process the initial still waterline
was adopted for all SHIPFLOW computations. However, the displacement, the waterline area and
the centroids were self-adjusting within the bounds. Infinite water depth was assumed throughout the
optimizations.
For the longitudinal wave cut analysis a lateral wave cut position of y/LPP = 0.25 was used throughout
the optimizations. In order to attain the longitudinal wave cut, the discrete wave elevations at the
free surface collocation points were evaluated and interpolated both for the free surface and the free

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

143

6 Applications
initial ~ optimized | | | deviation | |

Sectional area A (x) / AM [%]

20

15

10

-0.05

0
FP

-0.025

0.025
x / LPP [-]

0.05

0.075

0.1

initial ~ optimized | | | deviation | |

Sectional area A (x) / AM [%]

15

10

-5

-10
0.94

0.96

0.98
1
x / LPP [-] AP

1.02

1.04

Fig. 6.28: Sectional areas of the initial and the entirely optimized FantaRoRo, detail magnification of
the entrance (top) and run (bottom).
surface transom group. This was conducted outside SHIPFLOW within the optimization environment
MinSWASH, thus, always ensuring a consistent evaluation up to the wave cut analysis. Piecewise
cubic spline interpolation was used laterally and linear interpolation longitudinally. For the truncation
correction a reliable fitting length of 5/4 times the fundamental wavelength was settled throughout
the optimizations. The singularity of the correction function was allowed to self-adjust, promising an
optimal fit, compare section 3.2.3.
The perturbation magnitudes and n were selected according to section 4.2.2 with = ~n = 0.1.
The adjacent relative hull panel normal relocations were limited to 10% of the maximum possible hull

144

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

6.3 FantaRoRo ferry

Resistance ratios [%]

100
RWP

90

RWP (predict)
RW

80

RF ITTC`57
70
60
50
0

10
Optimization loop

15

20

Fig. 6.29: Optimization history of selected resistance components of the entirely optimized FantaRoRo
(related to the respective values of the initial hull).

Main hull form parameter ratios [%]

102
100
98
96

LCB

94

VCB
92

AWL
LCF

90
0

10
Optimization loop

15

20

Fig. 6.30: Optimization history of selected main hull form parameters of the entirely optimized
FantaRoRo (related to the respective values of the initial hull).
panel relocation per sub-optimization loop. This condition was necessary to synchronize the predicted
optimization result with the result of the correction step. However, a sequence of sub-optimization tasks
was iterated so that all prominent hull variations were allowed to evolve, see also section 6.2. During
the optimization the initial panel topology was maintained by controlling the panel alignment.
Generally, the flat of bottom of the FantaRoRo and the stem profile were left untouched by freezing
the respective panels during the optimization. In case of the bulbous bow optimization all bulb panels
and the panels covering the transition region to the entrance up to approximately 5% LPP aft of FP and

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

145

6 Applications

Wave pattern resist. coeff. CWP 1000 [-]

initial ~ optimized | | | deviation | |


1.5

transverse

diverging

0.5

+
+

-0.5

20
40
60
Component wave direction Q []

Fig. 6.31: Wave pattern resistance coefficient CW P over the component wave direction of the initial
and the entirely optimized FantaRoRo.
initial ~ optimized | | | deviation | |

Wave pattern resistance fraction [%]

100
transverse

diverging

75
50
25
0

-25

-39.9%
-42.5%

-50
0

-44.2%

20
40
60
Component wave direction Q []

Fig. 6.32: Wave pattern resistance fraction over of the initial and the entirely optimized FantaRoRo
(related to the total wave pattern resistance of the initial hull).
up to the bulb height at FP girthwise were included in the optimization. A smooth transition to the
remaining entrance panelization is attained by an appropriate choice of the panels utilized for the hull
panel normal relocation scheme.
For the entire FantaRoRo optimization all hull and extra panels (above the still waterline), except the
flat of bottom and the stem panels, were utilized in the optimization. The local influence segment of

146

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

6.3 FantaRoRo ferry


initial ~ optimized | | |

0.006

Wave elevation / LPP [-]

0.004
0.002
0
-0.002
-0.004
-0.006
0

x / LPP [-]

Fig. 6.33: Longitudinal wave cut at y/LPP = 0.25 of the initial and the entirely optimized FantaRoRo.
initial ~ optimized | | | deviation | |

initial ~ optimized | | |
Wave pattern resist. coeff. CWP 1000 [-]

0.006

Wave elevation / LPP [-]

0.004
0.002
0
-0.002
-0.004
-0.006
0

1.5

0.5

-0.5

x / LPP [-]
initial ~ optimized | | | deviation | |

0.05

-0.05

20
40
60
Component wave direction []

initial ~ optimized | | | deviation | |

0.1

Free wave spectra AC [-]

Free wave spectra AS [-]

0.1

20
40
60
Component wave direction []

0.05

-0.05

20
40
60
Component wave direction []

Fig. 6.34: Summarized results of the longitudinal wave cut (y/LPP = 0.25) analysis. Comparison of
the initial and the entirely optimized FantaRoRo.

a single optimization variable was limited to the direct neighbourhood, i.e., up to level 1, compare
section 4.5.

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

147

6 Applications

Fig. 6.35: Birds view of the initial (top), the bulbous bow optimized (middle) and the entirely optimized FantaRoRo (bottom).
The same additional bounds for the maximum hull extensions and selected main hull form parameters
were defined both for the bulbous bow and the entirely optimized FantaRoRo, see table 6.2 and 6.3.
The bulbous bow was allowed to double its initial length forward of FP which conforms with 7% LPP
forward of FP. The maximum aft extension of the hull was limited to 4% LPP aft of AP which allows
a slight extension of the hull astern. Both breadth and draft were kept (within a small tolerance). The
displacement variation was relaxed to 5% of the initial displacement. The displacement bounds were
applied to the wetted hull portion under the wavy free surface of the advancing ship. The same as for
the displacement applies to the waterline area. The longitudinal and vertical location of the center of
buoyancy were bounded to 6% of the respective initial values. The longitudinal center of flotation
was free to change 8% of the initial value.
In total 15 optimization loops with alternating prediction and correction steps were performed for the
bulbous bow optimization of the FantaRoRo. The wave pattern resistance was minimized over the

148

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

6.3 FantaRoRo ferry

Fig. 6.36: Fishs view of the initial (top), the bulbous bow optimized (middle) and the entirely optimized FantaRoRo (bottom).

whole range of wave directions. The process was terminated when the maximum number of optimization loops were reached according to the criterium (6.3). However, marginal improvements were
achieved within the last three loops, so the result of optimization loop 12 was selected as representative, compare the optimization histories of the resistance components in figure 6.20. For the entire hull
optimization of the FantaRoRo in total 17 loops were performed, whereas the result of optimization
loop 15 was selected as representative for the same reason as in case of the bulbous bow optimization.
The wave pattern resistance was minimized over the whole wave directions range. In fact the entire
FantaRoRo optimization was split into two successive optimization runs. In the first run the entire hull
was optimized until the lower displacement bound was reached in loop nine with only marginal improvements in the following loops (which are not presented in the graphs), compare the optimization
histories in figure 6.29 and 6.30. Then in a second successive optimization run (loop 10 to 17) it was
focused solely on the bulbous bow and a part of the entrance region up to approximately 17% LPP aft

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

149

6 Applications

Fig. 6.37: Views of the bulbous bow and entrance of the initial (top), the bulbous bow optimized (middle) and the entirely optimized FantaRoRo (bottom).
of FP. This was to investigate how further improvements are possible by focusing the optimization to
a particular confined hull region.
In figure 6.20 and 6.21 the optimization histories for the resistance components and the selected main
hull form parameters are presented for the bulbous bow optimized FantaRoRo. They are related to
their respective initial values (i.e., a value of 100% means identical to the initial value). In view
of the relatively small hull variations confined to the bulbous bow area a substantial decrease of the
wave pattern resistance by 10% and of the wave-pattern-resistance-displacement ratio by 10.6% was
achieved, see figure 6.20. Naturally, the resistance gain is substantially smaller compared to the entire
hull optimization. The histories of RW P for the predicted (indicated by predict) and the recomputed
wave pattern resistance in the correction step as well as the wave resistance RW from hull pressure
integration taken from the correction step are all in good agreement. Figure 6.21 reveals the form
parameter histories. The displacement increases almost linearly due to an increase of the bulb length
and volume, whereas, at the same time, the longitudinal center of buoyancy LCB is shifted towards the
bow. The remaining main hull form parameters stay almost unchanged. Figure 6.21 shows that all
form parameter constraints were inactive throughout the optimization.
In figure 6.22 the wave pattern resistance (given as coefficient CW P ) as a function of the component
wave direction is compared for the initial and the bulbous bow optimized FantaRoRo. The deviation
of both resistance curves gives a distinct picture of the gains and losses for the particular wave pattern components. The transverse waves following the stern in = 0 were slightly amplified and no
effective improvement of the wave pattern resistance was attained up to = 20 . However, tangible improvement of the wave pattern resistance is present in the upper transverse waves range and at
the transition to the diverging waves. The upper diverging wave range is characterized by alternating

150

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

6.3 FantaRoRo ferry

Fig. 6.38: Side view of the initial (top), the bulbous bow optimized (middle) and the entirely optimized
FantaRoRo (bottom).
gains and losses in RW P which almost balance in the sum. Figure 6.23 shows a comparison of the
wave pattern resistance fraction over the component wave direction for the initial and the bulbous bow
optimized FantaRoRo. The graph indicates the wave pattern resistance fraction integrated up to the
particular component wave direction. All values are related to the total wave pattern resistance of the
initial hull. The deviation curve reveals that only a marginal reduction of the wave pattern resistance
was attained up to = 20 . Moreover, the positive slope of the deviation curve at = 0 indicates an
amplification, even though slightly, of the transverse waves following the stern. This may be explained
by the fact that the bulbous bow wave crest and the wave crest aft of the transom in the stern flow
are positioned approximately two fundamental wavelengths apart, i.e., both wave trains are in phase.
This can be seen nicely from the wave contour plots in figure 6.41 and 6.42. Hence, superimposing an
amplified bulbous bow wave crest in phase onto an almost unchanged wave flow at the run and stern
might explain the slightly amplified stern waves. However, the effective improvement of the wave
pattern resistance by 9.4% is associated with a tangible reduction of the wave components in the range
20 40 . Whereas the aggregated contribution of the upper divergent wave range is negligible.
This above picture is very descriptive to explain why the application of large protruding hook-shaped

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

151

6 Applications

Fig. 6.39: Bottom view of the initial (top), the bulbous bow optimized (middle) and the entirely optimized FantaRoRo (bottom).
bulbous bows is often promising in cases of medium speed and fast ships in the range of 0.2 Fn
0.4. At these Froude numbers a considerable amount of wave pattern resistance is dedicated to the
wave components range 20 40 which, apparently, can be effectively reduced by a suitable
bulbous bow, as shown in the present example. Meanwhile, at very low and very high speeds the wave
pattern resistance peaks are shifted towards the lower transverse and the upper divergent wave ranges,
respectively. Hence, either differently shaped bulbs are required or it is even found advantageous to
omit a bulbous bow completely. In conclusion, the graphs 6.22 and 6.23 are thought to be particulary
valuable to assign the effect of a locally confined hull variation to a specific wave range and vice
versa.
A summary of the optimization results is given in figure 6.24 and 6.25 in terms of the analysed wave
cut, the wave pattern resistance coefficient and the free wave spectra for the initial and the bulbous bow
optimized FantaRoRo.
A close inspection of the wave contour plots in figure 6.41 and 6.42 reveals that a favourable interaction
of the bulb with the bow wave and the entrance region leads to a wave height reduction in the vicinity
of the bow up to the middle part. However, the wave pattern astern is almost unchanged since hull
modifications were conducted only at the bulbous bow and at the transition to the entrance. These
observations are in agreement with the CW P plots in figure 6.22.
The bulbous bow optimization generated a hook-shaped protruding bulbous bow, penetrating the still
water level, with a considerable volume increase forward and aft of FP and an increased bulb length
forward of FP by 25%. The bulb top raised above the still water level. A smooth transition of the
variation region to the unaltered entrance part at approximately 5% LPP aft of FP was accomplished.
For a graphical impression see the ship lines, figure 6.17, the sectional area curves, figure 6.18 and
6.19, and the hull renderings, figures 6.35 to 6.39. In consequence of the volume increase and the
bulb elongation a pronounced underpressure region originates from the bulb top towards its back,

152

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

6.3 FantaRoRo ferry

Fig. 6.40: Hull panelization of the initial FantaRoRo (fishs view with part of the wavy free surface).
as the dynamic hull pressure distribution in figure 6.44 implies. The underpressure counteracts the
overpressure which emerges due to the waterline expansion at the bow. In other words, the down-flow
at the back of the bulb counteracts the up-flow at the bow stagnation, compare the velocity vector plots
in figure 6.45.
In figure 6.29 and 6.30 the optimization histories for the resistance components and the selected main
hull form parameters are presented for the entirely optimized FantaRoRo. A substantial decrease of
the wave pattern resistance by 39.9% and of the wave-pattern-resistance-displacement ratio by 36.9%
was achieved, see figure 6.29. Extrapolating the results gives a reduction of the total resistance of 16%
and of the total resistance-displacement ratio by 11.5%. 8
The histories of RW P for the predicted (indicated by predict) and the recomputed wave pattern resistance in the correction step are in good agreement. The wave resistance RW from hull pressure integration features smaller gains. However, the RW resistance curve still declines within the last optimization
loops, whereas RW P only reduces slightly. This discrepancy might be explained by the sensitivity of
RW to panelization changes, particulary, at the bulbous bow close to the stagnation point. From loop 10
on the optimization focused on a continuous elongation of the bulb tip area which was accompanied by
a tangible change of the bulb panelization. As a result, the final bulbous bow optimization contributes
only by an improvement in RW P of further 1.5%. Figure 6.30 presents the form parameter histories.
The displacement , the waterline area AW L and the longitudinal center of buoyancy LCB decreased,
whereas the longitudinal center of flotation LCF stayed unchanged. The displacement approached its
lower bound. All other form parameter constraints were inactive throughout the optimization.
In figure 6.31 the wave pattern resistance (given as coefficient CW P ) as a function of the component
wave direction is compared for the initial and the entirely optimized FantaRoRo. A significant reduction of the wave heights of the transverse waves following the stern at = 0 was achieved, the value
being 25%. The transverse waves, ranging up to = 35 16 , contribute substantially to the reduction of
the total wave pattern resistance. In contrast, the diverging wave range is characterized by alternating
gains and losses in RW P even impairing the overall result. Figure 6.32 shows a comparison of the wave
pattern resistance fraction over the component wave direction for the initial and the entirely optimized
FantaRoRo. The deviation curve reveals that already the minimization of the transverse waves range up
to = 35 16 contributes by a reduction of 42.5%. Moreover, the negative slope of the deviation curve
at = 0 indicates a significant reduction of the transverse waves which directly follow the ship.
8

The viscous resistance component was calculated according to the ITTC formula with a fixed form factor of 1 + k = 1.085,
compare Valdenazzi et al. (2003).

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

153

6 Applications

Fig. 6.41: Wave contours of the initial (top), the bulbous bow optimized (middle) and the entirely optimized FantaRoRo (bottom).
A summary of the optimization results is given in figure 6.33 and 6.34 in terms of the analysed wave
cut, the wave pattern resistance coefficient and the free wave spectra for the initial and the entirely
optimized FantaRoRo.
A close inspection of the wave contour plots in figure 6.41 and 6.42 as well as the birds views of the
wave pattern in figure 6.43 reveals that favourable interactions of the particular wave trains lead to a
substantial reduction of the bow and stern waves and to a reduction of the transverse waves regime.
Otherwise, a pronounced wave trough emerges at the augmented fore shoulder and a less developed
wave trough at the aft shoulder. The latter is shifted towards the fore shoulder compared to the initial
hull. The question arises: Why does the optimization lead to this wave pattern formation?
The answer is found by measuring the distances of the wave crests and troughs along the hull and
balancing them with the hull shape variations. The hull shape variations are depicted in the plots

154

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

6.3 FantaRoRo ferry

l0

l0 /2

l0

Fig. 6.42: Wave contours (enhanced resolution) in close proximity to the initial (top), to the bulbous
bow optimized (middle) and to the entirely optimized FantaRoRo (bottom).
of the ship lines, figure 6.26, the sectional area curves, figure 6.27 and 6.28, and the hull surface
renderings, figures 6.35 to 6.39. The optimization developed a hook-shaped protruding bulbous bow,
penetrating the still water level, with a considerable volume increase forward of FP and an elongation
by 24% of the initial size. Thus the bulb wave is amplified allowing an accentuated reduction of

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

155

6 Applications

Fig. 6.43: Birds view of the wave pattern of the initial (port side) and the entirely optimized FantaRoRo
(starboard side).
the bow wave. However, the bulbous bow of the entirely optimized hull is less pronounced than the
bulb generated by the bulbous bow optimization alone. This is, because the entire hull optimization
narrowed the waterline entrance angles accompanied by a tangible reduction of the entrance volume
which in turn decreases the bow and entrance waves and, hence, only calls for a comparatively moderate
accentuation of the bulbous bow. Opposite to the narrowed waterline entrance angles the waterlines
were pushed outwards at the fore shoulder which is accompanied by a volume increase reaching from
35% to 60% LPP aft of FP. Furthermore, the maximum section is shifted by nearly 5% LPP towards the

156

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

6.3 FantaRoRo ferry

Fig. 6.44: Dynamic hull pressure of the initial (top), the bulbous bow optimized (middle) and the entirely optimized FantaRoRo (bottom).

bow, but still is located slightly aft of the midship section. In consequence of the narrowed waterline
entrance angles and a widening of the hull at the fore shoulder an inflection point is introduced to
the entrance waterlines and, hence, the flow deflection at the fore shoulder is amplified. As a result,
a pronounced wave trough emerges aft of the fore shoulder, which is, compared to the initial hull,
shifted by approximately 10% LPP downstream. Moreover, further astern at around 75% LPP aft of FP
a slight contraction is introduced to the waterlines which extends vertically from approximately half
the draft to the still water level and above. This generates a second, however much less pronounced,
wave trough at the aft shoulder. Compared to the initial hull the aft shoulder wave trough is effectively
shifted by more than 10% LPP upstream. At the run the optimization considerably reduced the volume
accompanied by an increased stern rise which can be viewed in figure 6.26 by the raised buttocks at
the run. Moreover, the stern tunnels were accentuated by the optimization. Compared to the initial hull
the transom of the entirely optimized FantaRoRo is completely emerged at the ship at rest. Forward
of the transom the concave buttocks shape is amplified due to an upward movement of the run. This
in turn allows a horizontal run out of the buttocks towards the transom edge. As a result the flow
separates distinctly more horizontally at the transom edge of the entirely optimized FantaRoRo. This
then reduces the stern wave and, hence, the transverse waves which directly follow the ship. The effect
of the optimized run geometry is shown in the velocity vector plots in figure 6.46. Regarding the hull
form coefficients (related to the ship at rest) a decrease of the prismatic coefficient CP and the waterplane area coefficient CW P by 8% to 9% is apparent. However, the midship section is enlarged which
is expressed by an increased midship section coefficient CM .

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

157

6 Applications

Fig. 6.45: Velocity vectors on the entrance of the initial (top), the bulbous bow optimized (middle) and
the entirely optimized FantaRoRo (bottom).
In conclusion, the following main optimization achievements led to a substantial reduction of the bow
and stern waves and to a reduction of the transverse waves, compare figure 6.42:
The down-flow at the back of the bulb counteracts the up-flow at the bow stagnation, reducing
the bow wave.
Furthermore, the narrowed waterline entrance angles reduce the bow and entrance waves.
The bulbous bow wave crest and the fore shoulder wave trough are positioned slightly less than
one fundamental wavelength apart from each other. Both wave trains are thus opposite in phase
which results in a beneficial interference.
The less pronounced wave trough at the evolving aft shoulder is slightly less than half the fundamental wavelength from the fore shoulder wave trough. Hence, both wave trains are opposite
in phase which results in a positive cancelation.
The fore shoulder wave trough and the wave crest aft of the transom in the stern flow are positioned one fundamental wavelength from each other, which again results in a mutual cancelation.
The nearly horizontal flow separation at the transom edge substantially reduces the stern wave
and, hence, the transverse waves which directly follow the ship.
Further relaxing of the bounds even amplifies the above outlined variation trends. However, the flow
deflection at the fore shoulder and at the transition from the flat of bottom to the stern rise eventually
becomes too large which may cause adverse viscous effects. Furthermore, the favourable interferences
of waves as predicted by potential flow theory will be less pronounced in real flow.

158

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

6.3 FantaRoRo ferry

Fig. 6.46: Velocity vectors in the run and flow separation at the transom of the initial (top), the bulbous
bow optimized (middle) and the entirely optimized FantaRoRo (bottom).
A comparison of the present optimization achievements with the results of the FantaRoRo optimization
study conducted within the FANTASTIC project at Technical University Berlin shows good agreement
on a qualitative basis for the fore body hull shape. Both optimization approaches yield a pronounced
and elongated bulbous bow, narrow entrance waterlines and a widening of the fore shoulder. For the
FANTASTIC FantaRoRo optimization a reduction of the wave pattern resistance by 15% and a gain
in the extrapolated total resistance by 6% were prognosticated, see for instance Heimann and Harries
(2003). The achieved resistance gain in the present optimization study is about 2.5 times higher. However, within the FANTASTIC FantaRoRo optimization only the fore body hull shape up to the midship
section was allowed to change and the displacement was kept strictly.

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

159

6 Applications

160

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

7 Conclusion
With the present work a novel hull form optimization approach has been introduced, which substantially differs from other approaches which address the same objective, i.e., the optimization of the
wave-making characteristics and the wave (pattern) resistance of ships. The optimization process is
fully automated requiring no user interaction, however, permitting it.
All aspects of the proposed optimization approach were presented. The application examples clearly
show the merits of the method. Naturally, further research is needed to refine the approach, to make
use of the full potential and study its ultimate limitations.

7.1 Main achievements and contributions


The present work is designated to contribute to the field of hydrodynamic optimization of the wavemaking characteristics and the wave (pattern) resistance of ship hulls. The following has been shown:
Hull variations can be driven directly by hydrodynamics. Only a few prerequisites to the hull
geometry representation are inevitably imposed due to the discretization scheme.
The free wave spectra and the wave pattern resistance distribution from wave cut analysis can
be favourably utilized in terms of the objective function which considerably improves the system identification, compared to pure integral resistance values, and increases the capabilities
of the optimization due to the additional information. As an additional benefit computed wave
pattern and hence wave cut analysis based on them prove to be less sensitive to the flow field
discretization than hull pressure integration.
Hull optimization may be better carried out directly for the effective wetted hull portion of the
advancing ship, including all effects due to dynamic trim and sinkage and wave formation along
the hull.
The optimization process can be established in terms of an iterative marching scheme of successive sub-optimization loops, each mapping the solution space to a simplified convex quadratic
image, the sub-optimization space, which merely possesses a single minimum determined by the
active constraints.
Moreover, linear perturbation of the boundary value problem proved to be a valuable method of accessing nonlinear relations. The simplifications substantially accelerate the gradient evaluations. Consequently, a simultaneous treatment of a large number of locally acting optimization variables becomes
manageable, which in turn introduces a high degree of freedom to the hull variation.
Following conclusions can be drawn from the optimization applications:

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

161

7 Conclusion
The wave-making characteristics of the test cases are greatly improved, particulary due to a
reduction of the adverse effects in the transverse waves regime.
The wave pattern resistance and the wave resistance from pressure integration is reduced by up
to 50% for an academic test case. A similar scale of reduction is not likely for a reasonable initial
design but highlights the potential of the scheme. Tangible improvements could be brought about
for a contemporary ferry which shows the practicality of the work.
The hull variations are driven by the optimization, within the given constraints, to the optimum
hull shapes. The hull geometry is fully self-adjusting according to the optimization requirements.
As a consequence of directed local hull variations, optimal interferences of the local wave trains
are achieved both in amplitude and phase which results in a beneficial cancelation of the wave
trains.
The bulbous bow optimization reveals the capability of a concerted reduction of a particular wave
range by correlation of cause and effect of local hull variations and their respective fingerprints
in the wave spectrum and wave pattern resistance.
Additional information is gained by the wave spectra and the wave pattern resistance distribution
which is utilized in the optimization in terms of an improved system identification.

7.2 Outlook
Future work may be dedicated to a further improvement and validation of the optimization method and
to a coupling to high-level geometric modelling schemes and to other optimization approaches with a
different focus.

7.2.1 Improvement and validation

In the present implementation of the method the adjacent relative hull panel normal relocations are
limited in order to overcome the discrepancy between prediction and correction by reducing the effect
of undesired panel rotations. However, this introduces a certain coupling to the panel mesh limiting the
variety of variation.
It frequently occurred during the application that the results of the prediction step are spoiled in the
correction step. This is caused by the joint relocation scheme when reconstructing a coherent input
point mesh. Hence, the hull variation in terms of the hull panel normal relocation scheme ought to be
improved to account for the immanent panel rotations in the reconstruction phase. One solution might
be to consider a priori the hull panel rotations about the panel principal axes in the perturbation scheme
and in terms of additional optimization variables. It is also conceivable to exchange the joint relocation
scheme by a different reconstruction approach, for instance by tracing of the limiting streamlines along
the hull, as suggested by Pien and Moore (1963).
A different aspect concerns the shape of the solution space and how it is mapped by the perturbation approach to a convex quadratic image in the sub-optimization problem. Further validation of the
present method is required even though the experience gathered so far is encouraging. An investigation

162

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

7.2 Outlook
of the solution space is reported by Peri et al. (2001) and strongly supports the present perturbation
approach.
Moreover, the effect of the immanent inconsistency between the nonlinear CFD simulation of the free
surface flow and the application of linearized wave cut analysis ought to be further investigated.
Although outside the focus of the present work, a further validation of the optimization results should
involve an investigation of the optimized hulls in terms of their viscous flow properties, e.g., streamlines, boundary layer thickening and separation, viscous wake flow and viscous wave interactions,
with the importance of the latter emphasized by Nowacki and Sharma (1972). Such a study might be
conducted either by a viscous CFD tool or by means of EFD.
Hendrix et al. (2001) and Percival et al. (2001) point out the importance of hull form optimization for
the entire or at least an augmented speed range. Therefore, in practical application the optimization
result must be verified for an extended speed range and different loading conditions.
A knowledge base might be created and utilized by the optimization scheme which could be consecutively extended by the information gained during the optimization. This might lead to an automatic
adaptation of the optimization process during runtime. For instance, a hull panel region which did
not effectively contribute to the overall optimization success in preceding optimization loops might be
automatically frozen by the scheme in the following process, thus saving computational costs.
In the present scheme first order panels with a constant source density distribution are adopted. However, the effect of using higher order curved panels with a linear source density distribution might be
investigated in terms of the optimization performance.
The final optimization outcome is a smooth hull point mesh as constructed by the joint relocation
scheme. However, further processing of the hull geometry, like for ship production, requires the application of suitable surface interpolation or approximation and fairing techniques, see e.g. Nowacki et al.
(1998), Nowacki and Kaklis (1998) and Westgaard (2000).

7.2.2 Coupling to high level geometric modelling and variation schemes

High level geometric modelling and variation schemes like form parameter oriented methods effectively condense the design variables required for the description of the hull shape, thus reducing their
overall number, see Birk (1998), Harries (1998) and Abt et al. (2003). However, in an optimization
application it is not a trivial task to select appropriate form parameters. Moreover, the selection of
appropriate variation bounds is critical.
In this sense the integration of the present optimization approach with a high-level geometric modelling
and variation scheme in a coupled optimization environment appears to be promising. Basically, the
presented optimization scheme may act as an advanced preprocessor to detect favourable hull variations
or rather the trend of the optimization. This information then can be directly utilized to set-up a form
parameter based hull optimization with an appropriate, reduced form parameter set which suits the
optimization demands.

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

163

7 Conclusion
7.2.3 Coupling to seakeeping optimization methods

A crucial measure of the quality of a ship hull is its seakeeping performance. In particular modern
container ship designs are susceptible to parametric rolling with the risk of capsizing, whereas ferries are suffering from short roll periods and high accelerations, see e.g. Cramer et al. (2004) and
Clauss and Hennig (2004). To account for the seakeeping performance the present calm water wave
resistance optimization ought to be accompanied by an assessment of the seakeeping characteristics
of the hull form. Or it could be coupled to a seakeeping optimization, like reported by Grigoropoulos
(2004). Moreover, Birk (1998) and Birk and Clauss (2001) developed a sophisticated seakeeping optimization method which they effectively applied to various types of floating systems including longterm probabilities.
At an advanced stage also aspects of optimal ship routing might be taken into account. An approach
to optimal ship routing is reported, e.g., by Harries et al. (2003), Hinnenthal and Harries (2004) and
Hinnenthal and Saetra (2004). Their method is based on knowledge of the ship performance in calm
water and waves which is coupled to detailed sea state and weather scenarios by a sophisticated route
optimization strategy.

164

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

Lists

Bibliography
Abt, C.; Harries, S.; Heimann, J.; Winter, H. (2003) From Re-design to Optimal Hull Lines by
Means of Parametric Modelling, Int. Conference on Computer and IT Applications in the Maritime
Industries COMPIT03, Hamburg, Germany.
Baskaya, . (1997) Hydrodynamischer Entwurf von Fahrgastkatamaranen fr Binnenwasserstrassen
der Berliner Region, (in German), Diploma Thesis, Institut fr Schiffs- und Meerestechnik, Technische Universitt Berlin, Germany.
Bertram, V.; Chao, K.Y.; Lammers, G.; Laudan, J. (1992) Entwicklung und Verifikation
Numerischer Verfahren zur Antriebsleistungsprognose Phase 2, (in German), Hamburgische
Schiffbau-Versuchsanstalt GmbH (Hamburg Ship Model Basin) (HSVA), Report No. 1579, BFT
Project 18 S 0014-0, Hamburg, Germany.
Birk, L. (1998) Hydrodynamic Shape Optimization of Offshore Structures, Dissertation (Ph.D. Thesis),
Institut fr Schiffs- und Meerestechnik, Technische Universitt Berlin, Germany.
Birk, L.; Harries, S. (2000) Automated Optimization A Complementing Technique for the Hydrodynamic Design of Ships and Offshore Structures, 1st Int. Conf. on Computer Applications and
Information Technology in the Maritime Industries COMPIT00, Potsdam, Germany.
Birk, L.; Clauss, G.F. (2001) Rational Design Criteria and their Application to Hull Form Optimisation of Floating Systems in Random Seas, Proc. of the 8th Int. Symposium on Practical Design of
Ships and Other Floating Structures PRADS01, Shanghai, China.
Birk, L.; Harries, S. (Eds.) (2003) OPTIMISTIC Optimization in Marine Design, 39th WEGEMT
Summer School 19-23 May 2003, MENSCH & BUCH VERLAG Berlin, Berlin, Germany.
Bjrck, . (Ed.) (1996) Numerical Methods for Least Square Problems, Society for Industrial and
Applied Mathematics (SIAM), Philadelphia, USA.
Bristow, D.R. (1974) A Solution of the Inverse Problem for Incompressible Axisymmetric Potential
Flow, AIAA 7th Fluid and Plasma Dynamics Conf., Palo Alto, USA.
Bristow, D.R. (1980) Development of Panel Methods for Subsonic Analysis and Design, NASA CR
3234.
Brizzolara, S.; Bruzzone, D.; Cassella, P.; Scamardella, A.; Zotti, I. (1998) Wave Resistance and

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

165

Lists
Wave Patterns for High-Speed Crafts; Validation of Numerical Results by Model Tests, 22nd Symp.
on Naval Hydrodynamics, ONR, Washington D.C., USA.
Bronstein, I.N.; Semendjajew, K.A.; Musiol, G.; Mhlig, H. (1993) Taschenbuch der Mathematik,
(in German), Verlag Harri Deutsch, Thun, Frankfurt am Main, Germany.
Clauss, G.F.; Hennig, J. (2004) Deterministic Analysis of Extreme Roll Motions and Subsequent Evaluation of Capsizing Risk, Int. Shipbuilding Progress , Vol. 51, Issue 3/4 (Special Issue Stability).
Cole, J.D. (1968) Perturbation Methods in Applied Mathematics, Blaisdell Publishing Comp.,
Waltham, Massachusetts, USA.
Cramer, H.; Reichert, K.; Hessner, K.; Hennig, J.; Clauss, G.F. (2004) Seakeeping Simulations
and Seaway Models and Parameters Supporting Ship Design and Operation, 9th Int. Symposium on
Practical Design of Ships and Other Floating Structures PRADS04, Lbeck-Travemnde, Germany.
Dawson, C.W. (1977) A Practical Computer Method for Solving Ship-Wave Problems, 2nd Int. Conference on Numerical Ship Hydrodynamics, Berkeley, CA, USA.
van Dyke, M. (1975) Perturbation Methods in Fluid Dynamics, The Parabolic Press, Standford, California, USA.
Dumez, F.-X.; Cordier, S. (1996) Accuracy of Wave Pattern Analysis Methods in Towing Tanks, 21st
Symp. on Naval Hydrodynamics, ONR, Trondheim, Norway.
Eggers, K.W.H. (1962) ber die Ermittlung des Wellenwiderstandes eines Schiffsmodells durch
Analyse seines Wellensystems (Teil 1 und 2), (in German), Schiffstechnik Bd. 9, No. 46, pp. 7984/85 and Schiffstechnik Bd. 10, No. 52, pp. 93-106.
Eggers, K.W.H.; Sharma, S.D.; Ward, L.W. (1967) An Assessment of Some Experimental Methods for Determining the Wavemaking Characteristics of a Ship Form, Transactions of the SNAME,
Vol. 75.
Eggers, K.W.H. (1976) Wave Analysis, State of the Art 1975, Int. Seminar on Wave Resistance, The
Soc. of Naval Architects Japan.
Favre, J.-N.; Avellan, F.; Ryhming, I.L. (1987) Cavitation Performance Improvement by Using a 2-D
Inverse Method of Hydraulic Runner Design, Int. Conf. on Inverse Design Concepts and Optimization in Engineering Sciences II, Pennsylvania State Univ., USA.
Grigoropoulos, G.J. (2004) Hull Form Optimization for Hydrodynamic Performance, Marine Technology, Vol. 41, No. 4, pp. 167-182.
Harries, S.; Schulze, D. (1997) Numerical Investigation of a Systematic Model Series for the Design of
Fast Monohulls, 4th Int. Conference on Fast Sea Transportation FAST97, Vol.1, Sydney, Australia.
Harries, S. (1998) Parametric Design and Hydrodynamic Optimization of Ship Hull Forms, Dissertation (Ph.D. Thesis), Institut fr Schiffs- und Meerestechnik, Technische Universitt Berlin, Germany.
Harries, S.; Abt, C. (1999) Formal Hydrodynamic Optimization of a Fast Monohull on the Basis of
Parametric Hull Design, 5th Int. Conf. on Fast Sea Transportation FAST99, Seattle, USA.

166

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

Bibliography
Harries, S.; Valdenazzi, F.; Abt, C.; Viviani, U. (2001) Investigation on Optimization Strategies for
the Hydrodynamic Design of Fast Ferries, 6th Int. Conference on Fast Sea Transportation FAST01,
Southampton, UK.
Harries, S.; Heimann, J.; Hinnenthal, J. (2003) Pareto-optimal Routing of Ships, Int. Conference on
Ship and Shipping Research NAV03, Palermo, Italy.
Havelock, Sir T.H. (1928) The Wave Pattern of a Doublet in a Stream, Procs. of the Royal Society,
London, Vol. 121, Series A., pp. 515-523.
Havelock, Sir T.H. (1932) The Theory of Wave Resistance, Procs. of the Royal Society, London,
Vol. 138, Series A, pp. 339-348.
Havelock, Sir T.H. (1934a) The Calculation of Wave Resistance, Procs. of the Royal Society, London,
Vol. 144, Series A, pp. 514-521.
Havelock, Sir T.H. (1934b) Wave Patterns and Wave Resistance, Summer Meeting of the 75th Session
of the Institution of Naval Architects.
Heimann, J. (2000) Application of Wave Pattern Analysis in a CFD Based Hull Design Process, 3rd
Int. Numerical Towing Tank Symp. NuTTS00, Tjrn, Sweden.
Heimann, J.; Harries, S. (2003) Optimization of the Wave-Making Characteristics of Fast Ferries, 7th
Int. Conference on Fast Sea Transportation FAST03, Ischia (Gulf of Naples), Italy.
Hendrix, D.; Percival, S.; Noblesse, F. (2001) Practical Hydrodynamic Optimization of a Monohull,
SNAME Anaual Meeting01.
Hess, J.L.; Smith, A.M.O. (1962) Calculation of the Non-Lifting Potential Flow about Arbitrary
Three-Dimensional Bodies, Report No. E.S. 40622, Douglas Aircraft Co., Inc., Long Beach, CA,
USA.
Hess, J.L. (1972) Calculation of Potential Flow about Arbitrary Three-Dimensional Lifting Bodies,
Report No. MDC J5679-01, Douglas Aircraft Co., Inc., Long Beach, CA, USA.
Hess, J.L. (1979) A Higher Order Panel Method for Three-dimensional Potential Flow, Report No.
N62269-77-C0437, Douglas Aircraft Co., Inc., Long Beach, CA, USA.
Hinnenthal, J.; Harries, S. (2004) A Systematic Study on Posing and Solving the Problem of Pareto
Optimal Ship Routing, Int. Conference on Computer and IT Applications in the Maritime Industries
COMPIT04, Siguenza, Spain.
Hinnenthal, J.; Saetra, O. (2004) Robust Pareto-Optimal Routing of Ships utilizing Ensemble Weather
Forecasts, Int. Symposium Information on Ships IPIS04, Hamburg, Germany.
Hirayama, A.; Eguchi, T.; Kimura, K.; Fujii, A.; Ohta, M.; (1998) Optimum Hull Form Design
using Numerical Wave Pattern Analysis, 7th Int. Symposium on Practical Design of Ships and Mobile
Units PRADS98, Den Haag, The Netherlands.
Hsiung, C.C. (1981) Optimal Ship Forms for Minimum Wave Resistance, Journal of Ship Research,
Vol. 25, No. 2.

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

167

Lists
Huan, J.; Huang, T.T. (1998) Sensitivity Analysis Methods for Shape Optimization in Non-Linear Free
Surface Flow, Procs. 3rd Osaka Colloquium on Advanced CFD Applications to Ship Flow and Hull
Form Design, Osaka, Japan, 25-27 May.
Huang, C.-H.; Chiang, C.-C.; Chou, S.-K. (1998) An Inverse Geometry Design Problem in Optimizing Hull Surfaces, Journal of Ship Research, Vol. 42, No. 2, pp. 79-85.
Inui, T. (1962) Wave-Making Resistance of Ships, SNAME Transactions, Vol. 70.
Janson, C.E.; Larsson, L. (1996) A Method for the Optimization of Ship Hulls from a Resistance Point
of View, 21st Symposium on Naval Hydrodynamics, Trondheim, Norway.
Janson, C.E. (1997) Potential Flow Panel Methods for the Calculation of Free-surface Flows with Lift,
Ph.D. Thesis, School of Mechanical and Vehicular Engineering, Chalmers University of Technology,
Gteborg, Sweden.
Jensen, G.; Mi, Z.-X.; Sding, H. (1986) Rankine Source Methods for Numerical Solutions of the
Steady Wave Resistance Problem, Procs. 16th Int. Symp. on Naval Hydrodynamics, University of
California, Berkeley, USA.
Jensen, G. (1988) Berechnung der stationren Potentialstrmung um ein Schiff unter Bercksichtigung der nichtlinearen Randbedingung an der Wasseroberflche, (in German), Dissertation (Ph.D.
Thesis), Bericht Nr. 484, Institut fr Schiffbau, Universitt Hamburg, Germany.
Jensen, G.; Bertram, V.; Sding, H. (1989) Ship Wave-Resistance Computations, Procs. 5th Int. Conf.
on Numerical Ship Hydrodynamics, Hiroshima, Japan.
Lord Kelvin alias Thomson, W. (1906) Deep Sea Ship-Waves, Philosophical Magazine, Series 6, Vol.
11, pp. 1-25.
Kracht, A.M. (1978) Design of Bulbous Bows, SNAME Transactions, Vol. 86.
Kraus, A. (1989) Methoden der Strungstheorie als Hilfsmittel fr den interaktiven Entwurf von umstrmten Krpern, (in German), Dissertation (Ph.D. Thesis), Institut fr Schiffs- und Meerestechnik,
Technische Universitt Berlin, Germany.
Lalli, F.; Felice, F.Di.; Esposito, P.; Moriconi, A. (1998) Some Remarks on the Accuracy of Wave
Resistance Determination from Wave Measurements Along a Parallel Cut, 22nd Symp. on Naval
Hydrodynamics, ONR, Washington D.C., USA.
Lalli, F.; Felice, F.Di.; Esposito, P.; Moriconi, A.; Piscopia, R. (2000) Longitudinal Cut Method
Revisited: A Survey on Main Error Sources, JSR, Vol. 44, No. 2, pp. 120-139.
Larsson, L.; Broberg, L.; Kim, K.-J.; Zhang, D.H. (1989) New Viscous and Inviscid CFD Techniques
for Ship Flow, 5th Int. Conference on Numerical Ship Hydrodynamics, Hiroshima, Japan.
Larsson, L.; Broberg, L.; Kim, K.-J.; Zhang, D.H. (1990) A Method for Resistance and Flow Prediction in Ship Design, SNAME Transactions, Vol. 98.
Larsson, L. (1993) Resistance and Flow Predictions Using the SHIPFLOW Code, 19th WEGEMT
School, Nantes, France.

168

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

Bibliography
Larsson, L.; Janson, C.-E.; Broberg, L.; Regnstrm, B. (2004a) SHIPFLOW Homepage,
http://www.flowtech.se/Shipflow/, maintained by FLOWTECH Int. AB, Gteborg, Sweden.
Larsson, L.; Janson, C.-E.; Broberg, L.; Regnstrm, B. (2004b) SHIPFLOW Users Manual, Release
2.4 to 2.8.10, FLOWTECH Int. AB, Gteborg, Sweden.
Lee, Y.-S. (2003) Trend Validation of CFD Prediction Results for Ship Design (Based on Series 60),
Dissertation (Ph.D. Thesis), Institut fr Land- und Seeverkehr, Bereich Schiffs- und Meerestechnik,
Technische Universitt Berlin, Germany.
Letcher, J. et al. (1999) MultiSurf, Userss Manual and Newsletters, AeroHydro Inc., Southwest Harbor, ME, USA.
Lin, W.-C.; Webster, W.C.; Wehausen, J.V. (1963) Ships of Minimal Total Resistance, Procs. Int.
Seminar on Wave Resistance, University of Michigan, Ann Arbor, USA.
Luenberger, D.G. (Ed.) (1984) Linear and Nonlinear Programming, 2nd ed., Addison-Wesley Inc.,
Reading, Massachusetts, USA.
Maisonneuve, J.-J.; Harries, S.; Marzi, J.; Raven, H.C.; Viviani, U.; Piippo, H. (2003) Towards
Optimal Design of Ship Hull Shapes, 8th Int. Marine Design Conf. IMDC03, Athens, Greece.
Marzi, J. (2004) -SHALLO Homepage, http://www.hsva.de/nushallo/index.html, maintained by the
Hamburgische Schiffbau-Versuchsanstalt GmbH (Hamburg Ship Model Basin) (HSVA), Hamburg,
Germany.
Michell, J.H. (1898) The Wave Resistance of a Ship, Philosophical Magazine, Series 5, Vol. 45, No.
272, pp. 106-123, January 1898.
Michelsen, F.C.; Uberoi, S.B.S. (1971) A Study of Wave Resistance Characteristics Through the
Analysis of Wave Height and Slope Along a Longitudinal Track, Report No. HY-15, Hydro- and
Aerodynamics Laboratory, Hydrodynamics Section, Lyngby, Denmark.
Mittelmann, H.D.; Spellucci, P. (2004) Decision Tree for Optimization Software The Plato server,
http://plato.asu.edu/guide.html, maintained by the Dept. of Mathematics and Statistics at Arizona
State University and the Department of Mathematics at Technical University Darmstadt, Germany.
Nakos, D.E. (1991) Transverse Wave Cut Analysis by a Rankine Panel Method, 6th Int. Workshop on
Water Waves and Floating Bodies.
The NEOS Guide (2004) The NEOS Guide for Optimization, http://www-fp.mcs.anl.gov/otc/Guide/,
maintained by the Optimization Technology Center at Argonne National Laboratory and Northwestern University, USA.
Newman, J.N. (1963) The Determination of Wave Resistance from Wave Measurements along a Parallel Cut, Procs. of the Int. Seminar on Theoretical Wave Resistance, Univ. of Michigan, Ann Arbor,
USA.
Newman, J.N. (1976) Linearized Wave Resistance Theory, Int. Seminar on Wave Resistance, Japan.
Newman, J.N. (1977) Marine Hydrodynamics, The MIT Press, Cambridge, USA.

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

169

Lists
Ni, S.-Y. (1987) Higher Order Panel Methods for Potential Flows with Linear and Non-Linear Free
Surface Boundary Conditions, Ph.D. Thesis, Chalmers University of Technology, Gteborg, Sweden.
Nowacki, H.; Sharma, S.D. (1972) Free-Surface Effects in Hull Propeller Interaction, 9th Symp. on
Naval Hydrodynamics, Vol. 2 (Frontier Problems), ACR-203, Paris, France.
Nowacki, H. (1976) Einfhrung in die Methoden der Optimierung (Introduction to Optimization Methods), (in German), Kontaktkurs Rechnergesttzter Schiffsentwurf, IX. Fortbildungskurs, Institut fr
Schiffbau, Universitt Hamburg, Germany.
Nowacki, H.; Jonas, W.; Reese, D.; Schumann-Hindenberg, U. (1984) Fortschritte des rechneruntersttzten Entwerfens in der Schiffstechnik, (in German), Jahrbuch der Schiffbautechnischen
Gesellschaft STG, Vol. 78.
Nowacki, H. (1994) Rechnergesttzter Schiffsentwurf (Computer Aided Ship Design), (in German),
Lecture notes, Institut fr Schiffs-und Meerestechnik, Technische Universitt Berlin, Germany.
Nowacki, H. (1995) Computational Fluid Dynamics im Schiffsentwurf (Computational Fluid Dynamics in Ship Design), (in German), Lecture notes, Institut fr Schiffs-und Meerestechnik, Technische
Universitt Berlin, Germany.
Nowacki, H.; Bloor, M.I.G.; Oleksiewicz, B. (Eds.) (1995) Computational Geometry for Ships, World
Scientific Publ. Co. Pte. Ltd., London, UK, ISBN 9810221398.
Nowacki, H. (1997) Hydrodynamic Design of Ship Hull Shapes by Methods of Computational Fluid
Dynamics, Progress in Industrial Mathematics at ECMI 96 (European Consortium for Mathematics
in Industry), B.G. Teubner, Stuttgart, Germany.
Nowacki, H.; Westgaard, G.; Heimann, J. (1998) Creation of Fair Surfaces Based on Higher Order Fairness Measures with Interpolation Constraints, in Creating FAIR and SHAPE-Preserving
Curves and Surfaces, HCM-FAIRSHAPE Network (EU), published by B.G. Teubner, Stuttgart,
Germany.
Nowacki, H.; Kaklis, P.D. (Eds.) (1998) Creating FAIR and SHAPE-Preserving Curves and Surfaces,
HCM-FAIRSHAPE Network (EU), published by B.G. Teubner, Stuttgart, Germany, ISBN 3-51902636-8.
Nowacki, H.; Harries, S.; Schulze, D.; Stinzing, H.-D. (1999) Verification of Wave Cut Design
Methodology, Brite EuRam III Project CALYPSO, Task 1.8, Deliverable Report D1.8.
Percival, S.; Hendrix, D.; Noblesse, F. (2001) Hydrodynamic Optimization of Ship Hull Forms, Applied Ocean Research, Vol. 23, pp. 337-355.
Peri, D.; Rossetti, M.; Campana, E. (2001) Design Optimization of Ship Hulls via CFD Techniques,
Journal of Ship Research, Vol. 45, No. 2, pp. 145-155.
Pien, P.C.; Moore, W.L. (1963) Theoretical and Experimental Study on Wave-Making Resistance of
Ships, Procs. Int. Seminar on Wave Resistance, University of Michigan, Ann Arbor, MI, USA.

170

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

Bibliography
Pien, P.C.; Strom-Tejsen, J. (1968) A Hull Form Design Procedure for High-Speed Displacement
Ships, Procs. SNAME Diamond Jubilee Meeting, Williamsburg, USA.
Press, H. Teukolsky, S.A. Vetterling, W.T.; Flannery, B.P. (1992) Numerical Recipes in
C The Art of Scientific Computing, 2nd edition, Cambridge University Press, http://libwww.lanl.gov/numerical/bookcpdf.
Raven, H.C. (1996) A Solution Method for the Nonlinear Ship Wave Resistance Problem, Ph.D. Thesis,
Technische Universiteit Delft, The Netherlands.
Raven, H.C. (1997) The Nature of Nonlinear Effects in Ship Wave Making, Schiffstechnik Bd. 44 /
Ship Technology Research Vol. 44, Hamburg, Germany.
Raven, H.C.; Prins, H.J. (1998a) Wave Pattern Analysis Applied to Nonlinear Ship Wave Calculations, 13th Int. Workshop on Water Waves and Floating Bodies, Alphen a/d Rijn, The Netherlands.
Raven, H.C.; Prins, H.J. (1998b) Improving the RAPID Resistance Prediction, Brite EuRam III
Project CALYPSO, Task 2.2, Deliverable Report D2.2.
Raven, H.C. (2004) RAPID Homepage, http://www.marin.nl/, maintained by the Maritime Research
Institute Netherlands (MARIN), Wageningen, The Netherlands.
Schmiechen, M. (1999) Estimation of Spectra of Transient Functions from Finite Sets of Sampled
Values, Schiffstechnik / Ship Technology Research, Vol. 46, No. 2, pp. 111-127, Hamburg, Germany.
Schmidt, E. (1976) Numerische Berechnung und experimentelle Untersuchung des transonischen
Strmungsfeldes in stark umgelenkten Schaufelgittern, (in German), Dissertation (Ph.D. Thesis),
Stuttgart, Germany.
Schulze, D. (1996) Zur Berechnung des Einflusses lokaler Formvariationen auf viskose Profilumstrmungen mit Methoden der Perturbationstheorie, (in German), Dissertation (Ph.D. Thesis), Institut
fr Schiffs- und Meerestechnik, Technische Universitt Berlin, Germany.
Schulze, D.; Harries, S. (1997) Flow Simulation and its Validation for a Model Series of Fast Ship
Hulls, 7th Symposium on Computational Fluid Dynamics, Beijing, China.
Sharma, S.D. (1963) A Comparison of the Calculated and Measured Free-Wave Spectrum of an Inuid
in Steady Motion, Procs. Int. Seminar on Theoretical Wave Resistance, Ann Arbor, Michigan, USA,
pp. 201-257,270.
Sharma, S.D. (1965) Zur Problematik der Aufteilung des Schiffswiderstandes in zhigkeits- und
wellenbedingte Anteile, (in German), Jahrbuch der Schiffbautechnischen Gesellschaft STG, Bd. 59,
pp. 458-504, Hamburg, Germany.
Sharma, S.D. (1966) An Attempted Application of Wave Analysis Techniques to Achieve Bow-Wave
Reduction, 6th Symp. on Naval Hy-drodynamics, ONR/ACR-136, Washington D.C., USA.
Sding, H. (1997) Drastic Resistance Reductions in Catamarans by Staggered Hulls, 4th Int. Conference on Fast Sea Transportation FAST97, Vol.1, Sydney, Australia.
Sding, H. (2000) Program Kelvin: Methods and some results, SVA-Report (Schiffbau-Versuchs-

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

171

Lists
anstalt Potsdam Potsdam Model Basin) No. 2582.
Sding, H. (2001a) Widerstandsminderung durch rechneruntersttzte Variation der Schiffsform, (in
German), 21st Duisburger Kolloquium, Gerhard Mercator Universitt Duisburg, Germany.
Sding, H. (2001b) Widerstand aus dem Wellenbild: Programm Kelvin, (in German), Internal Report,
Technische Universitt Hamburg-Harburg, Germany.
Spinney, D. (2002) Analysis, Comparison amd Improvement of CFD Wave Resistance Predicition
Techniques, Report No. X-02/133, Department of Naval Architecture and Ocean Engineering,
Chalmers University of Technology, Gteborg, Sweden.
Taylor, D.W. (1915) Calculations for Ships Forms and the Light Thrown by Model Experiments upon
Resistance, Propulsion and Rolling of Ships, Paper No. 196, Transactions International Engineering
Congress, San Francisco, CA, USA.
Tuck, E.O.; Scullen, D.C.; Lazauskas, L. (2002) Wave Patterns and Minimum Wave Resistance for
High-Speed Vessels, 24th Symp. on Naval Hydrodynamics, Fakuoka, Japan.
Valdenazzi, F.; Harries, S.; Janson, C.-E.; Leer-Andersen, M.; Maisonneuve, J.-J.; Marzi, J.;
Raven, H.C. (2003) The FANTASTIC RoRo: CFD Optimisation of the Forebody and its Experimental Verification, Int. Conf. on Ship and Shipping Research NAV03, Palermo, Italy.
Valorani, M.; Peri, D.; Campana, E. (2003) Sensitivity Analysis Methods to Design Optimal Ship
Hulls, Optimization and Engineering, Vol. 4, pp. 337-364.
Wehausen, J.V.; Laitone, E.V. (1960) Surface Waves, In Handbuch der Phyisk Bd. IX,
Berlin/Gttingen/Heidelberg, Springer Verlag, pp. 446-814.
Weinblum, G. (1936) Widerstandsuntersuchungen an Schiffen, (in German), Zeitschrift fr angewandte Mathematik und Mechanik, Band 15, Heft 6.
Weinblum, G.; Wustrau, D.; Vossers, G. (1957) Schiffe geringsten Widerstandes, (in German),
Jahrbuch der Schiffbautechnischen Gesellschaft, STG, Band 51.
Wendt, J.F. (Ed.) (1996) Computational Fluid Dynamics - An Introduction, 2nd edition, Wendt, J.F.,
Springer-Verlag.
Westgaard, G. (2000) Construction of Fair Curves and Surfaces, Dissertation (Ph.D. Thesis), Institut
fr Schiffs- und Meerestechnik, Technische Universitt Berlin, Germany.
Wigley, C. (Ed.) (1963) The Collected Papers of Sir Thomas Havelock on Hydrodynamics, Wigley, C.,
Office of Naval Research, Dep. of the Navy, ONR/ACR-103, USA.
Wolfram, S. (2003) MATHEMATICA 5.0 Online Documentation, Wolfram Research Inc., Illinois,
USA.

172

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

Figures

Figures

2.1

Optimization process flow. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

15

3.1

Zonal approach of the CFD system SHIPFLOW (courtesy of FLOWTECH Int. AB.) . .

26

3.2

Schematic sketch of the ship wave system (only one symmetric half is considered). . . .

29

3.3

Linearized wave pattern of an even submerged dipole beneath the origin, k0 = 1 and
0 = 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

39

Related wave pattern resistance of the even submerged dipole, influence of the signal
length xE and the truncation correction. . . . . . . . . . . . . . . . . . . . . . . . . . .

40

Related wave pattern resistance of the even submerged dipole, influence of the wave cut
position yWC and the truncation correction. . . . . . . . . . . . . . . . . . . . . . . . . .

40

Wave pattern resistance of the Hamburg test case determined from computed
(SHIPFLOW) wave cuts and wave probe measurements (HSVA), influence of the signal length xE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

41

Wave pattern resistance of the Hamburg test case determined from computed
(SHIPFLOW) wave cuts and wave probe measurements (HSVA), influence of the wave
cut position yWC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

41

3.4
3.5
3.6

3.7

4.1

Artificial bicubic surface patch and panelization by 3 3 flat quadrilateral panels (right).

4.2

Relocation of the corner points of the center panel in panel normal direction (left), effect
of the equivalent relocation of the four central input points (right). . . . . . . . . . . . .

54

4.3

Homogeneous panel relocation (left), effect of the joint relocation (right). . . . . . . . .

55

4.4

Stepped panel relocation (left), effect of the joint relocation (right). . . . . . . . . . . . .

56

4.5

Generic ferry-type stern hull surface patch (upside down), y-symmetry, and panelization
by 4 4 flat quadrilateral panels (right). . . . . . . . . . . . . . . . . . . . . . . . . . .

57

4.6

Homogeneous panel normal relocation (left), effect of the joint relocation (right). . . . .

58

4.7

Upstream stepped panel normal relocation (left), effect of the joint relocation (right). . .

59

4.8

Downstream stepped panel normal relocation (left), effect of the joint relocation (right). .

60

4.9

SHIPFLOW panelization of a ferry hull, perspective view. . . . . . . . . . . . . . . . . .

61

4.10

Cut-out of the panelization showing the intersection region at the fore perpendicular FP
and the design waterline DW L (upper edge of the transition area from the bulb to the bow). 61

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

53

173

Lists
4.11

4.12

4.13
4.14

4.15
4.16
4.17

4.18
4.19

4.20

Panel overlapping at design waterline as a consequence of a homogenous panel normal


relocation (left), effect of the joint relocation (right) considering symmetry at the midships plane. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

62

Stem panels crossing the midships plane as a consequence of an inward normal relocation of the upper bow panels (left), effect of the joint relocation (right) considering
symmetry at the midships plane. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

63

Lengthwise stepped panel normal relocation (left), effect of the joint relocation (right)
considering symmetry at the midships plane. . . . . . . . . . . . . . . . . . . . . . . .

64

Generic 2-D panel relocation example examining two panels arranged in a blunt angle
(e.g. the transition from FOB to the bilge radius) permitted and prohibited (crossed
out) relocation cases are compared. . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

65

Process flow of hull perturbation and sensitivity analysis (detail magnification of the
process flow, figure 2.1). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

70

Schematic sketch of the hull panel influence segment of a single source density variation
at level 0, edge cases (left) and central location (right). . . . . . . . . . . . . . . . . . .

72

Outward normal shift of a panel segment at the bow (to the right) of the Wigley hull
(top side view). Predicted change of the hull source density distribution by the direct
perturbation approach (middle) compared to the result of a SHIPFLOW recomputation
(bottom). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

79

Panelization of the Wigley hull. Source density perturbations at selected panels (for the
sake of clarity only the panels at the starboard side are marked). . . . . . . . . . . . . .

84

Improvement of the wave making characteristics of the Wigley hull at Fn = 0.3. The
result of the first sub-optimization loop of an optimization process, considering all but
the center line hull source densities as free variables, is shown. . . . . . . . . . . . . . .

85

Change of the wave pattern characteristic due to variation of one single hull panel source
density in its optimum gradient direction until the perturbation bound is reached. Here
panel # 68, located above keel level 10% LPP downstream of the bow, is selected. . . . .

86

4.21

Change of the wave pattern characteristic due to variation of the hull panel source density
# 76, located directly beneath the wavy free surface level 10% LPP downstream of the bow. 87

4.22

Change of the wave pattern characteristic due to variation of the hull panel source density
# 252, located directly beneath the wavy free surface level 40% LPP downstream of the
bow. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

88

Change of the wave pattern characteristic due to variation of the hull panel source density
# 538, located directly beneath the wavy free surface level 10% LPP upstream of the stern.

89

Change of the wave pattern resistance due to variation of the single hull panel source
density # 68 (an abscissa value of +1 means that the source density variation is pushed
to its upper bound allowed in optimization). . . . . . . . . . . . . . . . . . . . . . . . .

90

4.23
4.24

174

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

Figures
4.25

Change of the wave pattern resistance due to variation of hull panel source density # 76. .

91

4.26

Change of the wave pattern resistance due to variation of hull panel source density # 252.

92

4.27

Change of the wave pattern resistance due to variation of hull panel source density # 538.

93

5.1

The hull improvement stage (detail magnification of the process flow, figure 2.1). . . . .

96

5.2

Generalized inequality constraint bounded by a lower (-) and an upper (+) bound. . . . . 101

5.3

Characteristic of the generalized constraint function subject to different weights w1 . . . . 102

5.4

Influence of the weight w1 on the characteristic of an optimization functional by means


of an example featuring a single objective w0 f (x) and a single constraint term w1 g1 (x). . 104

5.5

Schematic sketch of possible occurrences of induced gaps between adjacent panels (exemplary selection) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

6.1

Ship lines of the initial and the optimized Wigley hull. . . . . . . . . . . . . . . . . . . . 122

6.2

Sectional areas of the initial and the optimized Wigley hull (related to the midship sectional area AM of the initial hull). Note that the fore perpendicular FP is located at
x/LPP = 0. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124

6.3

Birds view of the initial (top) and the optimized Wigley hull (bottom). . . . . . . . . . . 125

6.4

Fishs view of the initial (top) and the optimized Wigley hull (bottom). . . . . . . . . . . 126

6.5

Bottom view of the initial (top) and the optimized Wigley hull (bottom). . . . . . . . . . 126

6.6

Optimization history of selected resistance components of the Wigley hull (related to the
respective values of the initial hull). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127

6.7

Optimization history of selected main hull form parameters of the Wigley hull (related to
the respective values of the initial hull). . . . . . . . . . . . . . . . . . . . . . . . . . . 127

6.8

Wave pattern resistance coefficient CW P over the component wave direction of the
initial and the optimized Wigley hull. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128

6.9

Wave pattern resistance fraction over of the initial and the optimized Wigley hull (related to the total wave pattern resistance of the initial hull). . . . . . . . . . . . . . . . . 128

6.10

Longitudinal wave cut at y/LPP = 0.25 of the initial and the optimized Wigley hull. . . . 129

6.11

Summarized results of the longitudinal wave cut (y/LPP = 0.25) analysis. Comparison
of the initial and the optimized Wigley hull. . . . . . . . . . . . . . . . . . . . . . . . . 129

6.12

Panelization of the initial Wigley hull (fishs view with part of the wavy free surface). . . 130

6.13

Wave contours of the initial (top) and the optimized Wigley hull (bottom). . . . . . . . . 130

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

175

Lists
6.14

Wave contours (enhanced resolution) in close proximity to the initial (top) and to the
optimized Wigley hull (bottom). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131

6.15

Birds view of the wave pattern of the initial (port side) and the optimized Wigley hull
(starboard side). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132

6.16

Dynamic hull pressure of the initial (top) and the optimized Wigley hull (bottom). . . . . 133

6.17

Ship lines of the initial and the bulbous bow optimized FantaRoRo. . . . . . . . . . . . . 135

6.18

Sectional areas of the initial and the bulbous bow optimized FantaRoRo (related to the
midship sectional area AM of the initial hull). Note that the fore perpendicular FP is
located at x/LPP = 0. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137

6.19

Sectional areas of the initial and the bulbous bow optimized FantaRoRo, detail magnification of the entrance. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137

6.20

Optimization history of selected resistance components of the bulbous bow optimized


FantaRoRo (related to the respective values of the initial hull). . . . . . . . . . . . . . . 138

6.21

Optimization history of selected main hull form parameters of the bulbous bow optimized
FantaRoRo (related to the respective values of the initial hull). . . . . . . . . . . . . . . 138

6.22

Wave pattern resistance coefficient CW P over the component wave direction of the
initial and the bulbous bow optimized FantaRoRo. . . . . . . . . . . . . . . . . . . . . . 139

6.23

Wave pattern resistance fraction over of the initial and the bulbous bow optimized
FantaRoRo (related to the total wave pattern resistance of the initial hull). . . . . . . . . 139

6.24

Longitudinal wave cut at y/LPP = 0.25 of the initial and the bulbous bow optimized
FantaRoRo. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140

6.25

Summarized results of the longitudinal wave cut (y/LPP = 0.25) analysis. Comparison
of the initial and the bulbous bow optimized FantaRoRo. . . . . . . . . . . . . . . . . . 140

6.26

Ship lines of the initial and the entirely optimized FantaRoRo. . . . . . . . . . . . . . . 141

6.27

Sectional areas of the initial and the entirely optimized FantaRoRo (related to the midship
sectional area AM of the initial hull). Note that the fore perpendicular FP is located at
x/LPP = 0. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143

6.28

Sectional areas of the initial and the entirely optimized FantaRoRo, detail magnification
of the entrance (top) and run (bottom). . . . . . . . . . . . . . . . . . . . . . . . . . . . 144

6.29

Optimization history of selected resistance components of the entirely optimized


FantaRoRo (related to the respective values of the initial hull). . . . . . . . . . . . . . . 145

6.30

Optimization history of selected main hull form parameters of the entirely optimized
FantaRoRo (related to the respective values of the initial hull). . . . . . . . . . . . . . . 145

176

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

Figures
6.31

Wave pattern resistance coefficient CW P over the component wave direction of the
initial and the entirely optimized FantaRoRo. . . . . . . . . . . . . . . . . . . . . . . . 146

6.32

Wave pattern resistance fraction over of the initial and the entirely optimized
FantaRoRo (related to the total wave pattern resistance of the initial hull). . . . . . . . . 146

6.33

Longitudinal wave cut at y/LPP = 0.25 of the initial and the entirely optimized
FantaRoRo. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147

6.34

Summarized results of the longitudinal wave cut (y/LPP = 0.25) analysis. Comparison
of the initial and the entirely optimized FantaRoRo. . . . . . . . . . . . . . . . . . . . . 147

6.35

Birds view of the initial (top), the bulbous bow optimized (middle) and the entirely
optimized FantaRoRo (bottom). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148

6.36

Fishs view of the initial (top), the bulbous bow optimized (middle) and the entirely
optimized FantaRoRo (bottom). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149

6.37

Views of the bulbous bow and entrance of the initial (top), the bulbous bow optimized
(middle) and the entirely optimized FantaRoRo (bottom). . . . . . . . . . . . . . . . . . 150

6.38

Side view of the initial (top), the bulbous bow optimized (middle) and the entirely optimized FantaRoRo (bottom). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151

6.39

Bottom view of the initial (top), the bulbous bow optimized (middle) and the entirely
optimized FantaRoRo (bottom). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152

6.40

Hull panelization of the initial FantaRoRo (fishs view with part of the wavy free surface). 153

6.41

Wave contours of the initial (top), the bulbous bow optimized (middle) and the entirely
optimized FantaRoRo (bottom). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154

6.42

Wave contours (enhanced resolution) in close proximity to the initial (top), to the bulbous
bow optimized (middle) and to the entirely optimized FantaRoRo (bottom). . . . . . . . 155

6.43

Birds view of the wave pattern of the initial (port side) and the entirely optimized
FantaRoRo (starboard side). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156

6.44

Dynamic hull pressure of the initial (top), the bulbous bow optimized (middle) and the
entirely optimized FantaRoRo (bottom). . . . . . . . . . . . . . . . . . . . . . . . . . . 157

6.45

Velocity vectors on the entrance of the initial (top), the bulbous bow optimized (middle)
and the entirely optimized FantaRoRo (bottom). . . . . . . . . . . . . . . . . . . . . . . 158

6.46

Velocity vectors in the run and flow separation at the transom of the initial (top), the
bulbous bow optimized (middle) and the entirely optimized FantaRoRo (bottom). . . . . 159

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

177

Lists

178

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

Tables

Tables

2.1

Wave resistance optimization techniques. . . . . . . . . . . . . . . . . . . . . . . . . . .

6.1

Results of the Wigley hull optimization. . . . . . . . . . . . . . . . . . . . . . . . . . . 123

6.2

Results of the bulbous bow optimization of the FantaRoRo ferry. . . . . . . . . . . . . . 136

6.3

Results of the entire hull optimization of the FantaRoRo ferry. . . . . . . . . . . . . . . 142

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

12

179

Lists

180

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

Symbols, Abbreviations

Symbols, Abbreviations

3.1

CFD simulation

Velocity potential of the known base solution

Total velocity potential

. . . . . . . . . . . . . . . . . . .

23

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

21

Disturbance potential

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

21

Potential of the undisturbed onset flow . . . . . . . . . . . . . . . . . . . . . . .

21

Source density per unit area of panel j

. . . . . . . . . . . . . . . . . . . . . . .

24

Source density per unit area at point q

. . . . . . . . . . . . . . . . . . . . . . .

24

~n

Unit normal vector on the hull, directed into the fluid

~
U

Velocity vector: (Ux ,Uy ,Uz )T

~
U

Velocity vector of the undisturbed onset flow: (Ux ,Uy ,Uz )T

. . . . . . . . . .

21

b
U

Unit velocity vector of the undisturbed onset flow . . . . . . . . . . . . . . . . .

21

Normalized free surface elevation = f (x, y)

22

Free surface elevation at the collocation point i . . . . . . . . . . . . . . . . . . .

25

Ai j

Influence coefficient from panel j to the velocity at panel i

. . . . . . . . . . . .

24

Bi

Inhomogeneous term at panel i . . . . . . . . . . . . . . . . . . . . . . . . . . .

24

Fn

Froude number

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

22

Acceleration due to gravity . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

22

Characteristic length . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

22

LPP

Length between perpendiculars . . . . . . . . . . . . . . . . . . . . . . . . . . .

21

NH

Total number of hull panels . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

25

NFS

Total number of hull and free surface panels . . . . . . . . . . . . . . . . . . . .

25

Mean distance to the hull . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

23

r pq

Distance from point q to p where the potential is to be computed

. . . . . . . . .

24

SD

Boundary surface of the flow domain . . . . . . . . . . . . . . . . . . . . . . . .

24

VS

Ship speed . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

22

. . . . . . . . . . . . . . .

21

. . . . . . . . . . . . . . . . . . . . . . . . . . . .

21

. . . . . . . . . . . . . . . . . . .

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

181

Lists
Xi j Zi j
3.2

Induced velocities from panel j to the velocities at panel i . . . . . . . . . . . . .

25

Wave cut analysis

Wavelength

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Fundamental wavelength

Water density

Component wave direction

Analytical asymptotic extension of the truncated wave cut signal

31

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

30

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

30

. . . . . . . . . . . . . . . . . . . . . . . . . . . . .

30

. . . . . . . . .

35

Normalized wave elevation . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

32

AC

Free wave spectrum, cosine component . . . . . . . . . . . . . . . . . . . . . . .

34

AS

Free wave spectrum, sine component . . . . . . . . . . . . . . . . . . . . . . . .

34

Fourier transform, cosine component . . . . . . . . . . . . . . . . . . . . . . . .

33

c1,2,3

Regression coefficients

35

CW FT

Weighted Fourier transform, cosine component

. . . . . . . . . . . . . . . . . .

36

CW P

Wave pattern resistance coefficient . . . . . . . . . . . . . . . . . . . . . . . . .

35

Water depth

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

30

Wavenumber . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

31

k0

Basic wavenumber

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

30

RW P

Wave pattern resistance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

34

Fourier transform, sine component

. . . . . . . . . . . . . . . . . . . . . . . . .

33

Longitudinal wavenumber . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

31

Sre f

Wetted hull surface area for normalization

. . . . . . . . . . . . . . . . . . . . .

35

SW FT

Weighted Fourier transform, sine component . . . . . . . . . . . . . . . . . . . .

36

TA

Draft at aft perpendicular

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

38

TF

Draft at fore perpendicular

. . . . . . . . . . . . . . . . . . . . . . . . . . . . .

38

Transverse wavenumber . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

31

xE

Tail end of the wave cut signal

35

182

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . .

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

Symbols, Abbreviations
Transverse position of the longitudinal wave cut . . . . . . . . . . . . . . . . . .

yWC

38

Sensitivity analysis

Variation of the wave elevation . . . . . . . . . . . . . . . . . . . . . . . . . . .

80

Variation of the wave elevation at the i-th free surface collocation point . . . . . .

82

AC

Variation of the free wave spectrum, cosine component

. . . . . . . . . . . . . .

88

AS

Variation of the free wave spectrum, sine component . . . . . . . . . . . . . . . .

88

AW L k

Variation of AW L due to the normal relocation of the waterline panel k . . . . . .

77

AW L

Accumulated variation of the waterline area

77

CW FT

Variation of the weighted Fourier transforms, cosine component

LCB k

Variation of LCB due to the normal relocation of hull panel k . . . . . . . . . . . .

74

LCB

Accumulated variation of the longitudinal center of buoyancy . . . . . . . . . . .

75

LCF k

Variation of LCF due to the normal relocation of the waterline panel k

. . . . . .

77

LCF

Accumulated variation of the longitudinal center of flotation

. . . . . . . . . . .

78

MSLCF

Accumulated variation of the 1st -order longitudinal statical moment of flotation

78

MSL

Accumulated variation of the 1st -order longitudinal statical moment of buoyancy .

75

MSV

Accumulated variation of the 1st -order vertical statical moment of buoyancy . . .

75

nk max

Maximum allowed hull panel normal relocation

64

SW FT

Variation of the weighted Fourier transform, sine component

. . . . . . . . . . .

87

Accumulated displacement variation . . . . . . . . . . . . . . . . . . . . . . . .

74

Vk

Displacement variation due to the normal relocation of hull panel k . . . . . . . .

74

VCB k

Variation of VCB due to the normal relocation of hull panel k . . . . . . . . . . . .

74

VCB

Accumulated variation of the vertical center of buoyancy

. . . . . . . . . . . . .

75

Vi

Variation of the volume flow through the i-th hull panel . . . . . . . . . . . . . .

52

Perturbation parameter controlling the magnitude of the source density variation .

51

~nmax

Perturbation parameter controlling the maximum panel relocation . . . . . . . . .

68

~n

Perturbation parameter controlling the panel relocation magnitude

52

. . . . . . . . . . . . . . . . . . . .
. . . . . . . . .

. . . . . . . . . . . . . . . . . .

. . . . . . . .

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

87

183

Lists
AW L

Parameter controlling the variation of the waterline area . . . . . . . . . . . . . .

77

Bow

Slack parameter controlling the longitudinal panel relocation at bow

. . . . . . .

68

Keel

Slack parameter controlling the vertical panel relocation at keel level . . . . . . .

69

MSLCF

Parameter controlling the variation of MSLCF

. . . . . . . . . . . . . . . . . . . .

79

MSL

Parameter controlling the variation of MSL

. . . . . . . . . . . . . . . . . . . . .

76

MSV

Parameter controlling the variation of MSV

. . . . . . . . . . . . . . . . . . . . .

76

MS

Slack parameter controlling the lateral panel relocation at midships plane

. . . .

69

PA

Slack parameter controlling the panel alignment . . . . . . . . . . . . . . . . . .

66

SS

Slack parameter controlling the lateral panel relocation at shipside

. . . . . . . .

69

Stern

Slack parameter controlling the longitudinal panel relocation at stern . . . . . . .

68

Parameter controlling the displacement variation . . . . . . . . . . . . . . . . . .

74

W FS

Slack parameter controlling the vertical panel relocation at the free surface level .

69

Small perturbation parameter . . . . . . . . . . . . . . . . . . . . . . . . . . . .

45

Of order of the argument

47

nk

Unit vector normal of hull panel k

TX TZ

Transfer matrices for the variation of the free surface wave elevation

. . . . . . .

83

T(0)

Transfer matrix of the hull panel perturbations . . . . . . . . . . . . . . . . . . .

49

Panel corner point (1 4) in vector notation . . . . . . . . . . . . . . . . . . . .

77

Panel control point in vector notation . . . . . . . . . . . . . . . . . . . . . . . .

60

CT k j

Vector pointing from the control point of panel k to the control point of panel j . .

66

IP

Panel input point (equivalent to offset point) in vector notation

. . . . . . . . . .

57

Source density per unit area of the n-th panel . . . . . . . . . . . . . . . . . . . .

47

(1)

Perturbation magnitude of the n-th panel source density . . . . . . . . . . . . . .

47

~ (1)

Vector of source density variations at the hull and free surface . . . . . . . . . . .

50

~nk

Shift magnitude of the k-th hull panel in its normal direction

. . . . . . . . . . .

47

~Pk

Any point on the flat quadrilateral panel k

. . . . . . . . . . . . . . . . . . . . .

47

c (0)
CT
kj

Unit distance vector pointing from CT k to CT j . . . . . . . . . . . . . . . . . . .

66

CP

CT

(1)

184

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
. . . . . . . . . . . . . . . . . . . . . . . . .

47

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

Symbols, Abbreviations
e (0)
A

Matrix product A(0) T A(0)

. . . . . . . . . . . . . . . . . . . . . . . . . . . . .

51

e (0)
T

Matrix product T(0) T T(0)

. . . . . . . . . . . . . . . . . . . . . . . . . . . . .

50

AW L init

Initial waterline area . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

77

AW L

Waterline area . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

76

Breadth

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

69

Functional expression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

50

Characteristic length . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

52

LCB init

Initial longitudinal center of buoyancy

. . . . . . . . . . . . . . . . . . . . . . .

75

LCB

Longitudinal center of buoyancy

. . . . . . . . . . . . . . . . . . . . . . . . . .

74

LCF init

Initial longitudinal center of flotation . . . . . . . . . . . . . . . . . . . . . . . .

78

LCF

Longitudinal center of flotation . . . . . . . . . . . . . . . . . . . . . . . . . . .

77

MSL init

Initial 1st -order longitudinal statical moment of buoyancy . . . . . . . . . . . . .

76

MSLCF init

Initial 1st -order longitudinal statical moment of flotation . . . . . . . . . . . . . .

79

MSV init

Initial 1st -order vertical statical moment of buoyancy

. . . . . . . . . . . . . . .

76

Number of hull panel source densities utilized in the optimization . . . . . . . . .

47

NH

Total number of hull panels . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

47

NK

Number of active hull panels applied in the panel normal relocation scheme

. . .

47

NM

Number of hull panel input points involved in the hull perturbation . . . . . . . .

57

Nm

Number of adjacent hull panels associated with the m-th input point

. . . . . . .

57

NFS
f

Number of panels including only a subset of the free surface panels . . . . . . . .

51

NFS

Total number of hull and free surface panels . . . . . . . . . . . . . . . . . . . .

46

Nfs

Number of free surface panels excluding the hull panels . . . . . . . . . . . . . .

83

Nka

Number of active hull panels in the panel normal relocation scheme . . . . . . . .

64

Nk p

Number of passive hull panels in the perturbation scheme . . . . . . . . . . . . .

64

Pk

Center point of the upper panel edge of the waterline panel k . . . . . . . . . . .

78

Si

Area of the i-th hull panel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

52

Sk

Area of hull panel k . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

74

(0)

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

185

Lists
T

Draft . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

69

Characteristic flow velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

51

Displacement volume . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

74

VCB init

Initial vertical center of buoyancy . . . . . . . . . . . . . . . . . . . . . . . . . .

75

VCB

Vertical center of buoyancy . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

74

Vinit

Initial displacement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

74

xBow

x-value of the foremost hull point . . . . . . . . . . . . . . . . . . . . . . . . . .

68

xStern

x-value of the aftmost hull point

. . . . . . . . . . . . . . . . . . . . . . . . . .

68

yB max

y-value of the maximum lateral hull extension . . . . . . . . . . . . . . . . . . .

69

yMS

y-value of the midships plane . . . . . . . . . . . . . . . . . . . . . . . . . . . .

69

zT max

z-value of the maximum vertical extension at keel . . . . . . . . . . . . . . . . .

69

zW FS

z-value of the hull panel control point located in the wavy free surface level

. . .

69

g j (~x)

General inequality constraint related to its bound . . . . . . . . . . . . . . . . .

102

C j max

Maximum allowed tolerance for the j-th constraint . . . . . . . . . . . . . . . .

101

C j org

Origin shift for the j-th form parameter constraint

. . . . . . . . . . . . . . . .

110

C j max

Parameter controlling the max allowed variation of the j-th form parameter . . .

110

C j org

Parameter controlling the origin shift C j org

. . . . . . . . . . . . . . . . . . .

110

Gradients matrix of the constraints terms . . . . . . . . . . . . . . . . . . . . .

109

AC

Gradients matrix of the free wave spectrum, cosine components . . . . . . . . .

107

AS

Gradients matrix of the free wave spectrum, sine components

. . . . . . . . . .

107

Linear system coefficient matrix . . . . . . . . . . . . . . . . . . . . . . . . . .

113

Diagonal matrix of weights governing the constraints terms

. . . . . . . . . . .

109

~
C
shi f t

Vector of the constraints center point shifts . . . . . . . . . . . . . . . . . . . .

109

~,~

Vectors of unknown Lagrange multipliers

. . . . . . . . . . . . . . . . . . . . .

97

Nabla-operator in vector notation . . . . . . . . . . . . . . . . . . . . . . . . . .

98

186

Improvement of the wave-making characteristics

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

Symbols, Abbreviations
~B

Right hand side vector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

~p

General vector of constant optimization parameters

~s

Vector of unknown slack variables

~w

Vector of positive weights . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

~x

General vector of free optimization variables indicating the minimum . . . . . . .

f
W

Diagonal matrix of numerical integration factors . . . . . . . . . . . . . . . .

107

w
em

numerical integration factors . . . . . . . . . . . . . . . . . . . . . . . . . . .

107

cf

Convergence acceleration factor . . . . . . . . . . . . . . . . . . . . . . . . . .

115

C j (~x)

Actual value of the j-th constraint . . . . . . . . . . . . . . . . . . . . . . . . .

101

Cj

Basic value of the j-th constraint

. . . . . . . . . . . . . . . . . . . . . . . . .

101

C j av

Average point of the lower and upper bound of the j-th constraint . . . . . . . .

102

C j org

Initial j-th constraint value, i.e., the origin before optimization . . . . . . . . . .

102

C j shi f t

Center-shift of the j-th constraint

. . . . . . . . . . . . . . . . . . . . . . . . .

101

F(~x,~,~)

Lagrange functional formulation

. . . . . . . . . . . . . . . . . . . . . . . . . .

97

f (~x ,~p)

General objective function of optimization . . . . . . . . . . . . . . . . . . . . .

96

fRW P (~x)

Objective function of hydrodynamic hull form optimization

g j (~x ,~p)

General inequality constraints . . . . . . . . . . . . . . . . . . . . . . . . . . . .

96

hk (~x ,~p)

General equality constraints . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

96

NG

Number of inequality constraints . . . . . . . . . . . . . . . . . . . . . . . . . .

96

NH

Number of equality constraints . . . . . . . . . . . . . . . . . . . . . . . . . . .

96

NM

Number of equidistant samplings used in numerical integrations . . . . . . . .

107

rf

Relaxation factor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

115

RW P /

Wave-pattern-resistance-displacement ratio . . . . . . . . . . . . . . . . . . . .

105

(0)

113

. . . . . . . . . . . . . . . .

96

. . . . . . . . . . . . . . . . . . . . . . . . .

97

. . . . . . . . . . .

101
96

101

Applications

Displacement force =g

. . . . . . . . . . . . . . . . . . . . . . . . . . . .

123

Displacement volume

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

123

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

187

Lists
KM

Transverse metacentric height above keel . . . . . . . . . . . . . . . . . . . . .

123

KM L

Longitudinal metacentric height above keel . . . . . . . . . . . . . . . . . . . .

123

Water density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

123

Static trim angle

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

123

AW

Waterplane area

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

123

AP

Aft perpendicular . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

123

Breadth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

123

CB

Block coefficient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

123

CM

Midship section coefficient

. . . . . . . . . . . . . . . . . . . . . . . . . . . .

123

CP

Prismatic coefficient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

123

CW P

Waterplane area coefficient

. . . . . . . . . . . . . . . . . . . . . . . . . . . .

123

Fn

Froude number . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

123

FP

Fore perpendicular . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

123

LBBow

Length of bulbous bow forward of FP . . . . . . . . . . . . . . . . . . . . . . .

136

LCB

Longitudinal center of buoyancy . . . . . . . . . . . . . . . . . . . . . . . . . .

123

LCF

Longitudinal center of flotation

. . . . . . . . . . . . . . . . . . . . . . . . . .

123

LOS

Length overall submerged . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

123

LPP

Length between perpendiculars

123

Ns max

Maximum number of optimization loops

RW

Wave resistance from hull pressure integration

. . . . . . . . . . . . . . . . . .

123

RW /

Wave-resistance-displacement ratio . . . . . . . . . . . . . . . . . . . . . . . .

123

RW P div

Wave pattern resistance of the diverging wave components . . . . . . . . . . . .

123

RW P trans

Wave pattern resistance of the transverse wave components

. . . . . . . . . . .

123

RW P

Wave pattern resistance from wave cut analysis . . . . . . . . . . . . . . . . . .

123

RW P /

Wave-pattern-resistance-displacement ratio . . . . . . . . . . . . . . . . . . . .

123

Wetted hull surface area . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

123

Draft . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

123

188

. . . . . . . . . . . . . . . . . . . . . . . . . .
. . . . . . . . . . . . . . . . . . . . .

121

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

Symbols, Abbreviations
VS

Ship speed

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

VCB

Vertical center of buoyancy

zS

Mean static sinkage

123

. . . . . . . . . . . . . . . . . . . . . . . . . . . .

123

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

123

Abbreviations
BEM

Boundary Element Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

19

CASHD

Computer Aided Ship Hull Design . . . . . . . . . . . . . . . . . . . . . . . . .

58

CFD

Computational Fluid Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . .

19

EFD

Experimental Fluid Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . .

20

FOB

Flat Of Bottom

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

49

FOS

Flat Of Side . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

49

IGES

Initial Graphics Exchange Specification

. . . . . . . . . . . . . . . . . . . . . .

58

LP

Linear Programming Problem . . . . . . . . . . . . . . . . . . . . . . . . . . . .

97

MoM

Measure of Merit

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

16

NLP

Nonlinear Programming Problem . . . . . . . . . . . . . . . . . . . . . . . . . .

97

NURBS

Non Uniform Rational B-Splines

RANSE

Reynolds-Average Navier-Stokes Equations

WCA

Wave Cut Analysis

WFS

Wetted Free Surface Level

. . . . . . . . . . . . . . . . . . . . . . . . .

118

. . . . . . . . . . . . . . . . . . . .

28

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

29

. . . . . . . . . . . . . . . . . . . . . . . . . . . . .

69

Subscripts and superscripts


(0)

Indicates the zeros order solution, i.e., the basic solution . . . . . . . . . . . . . .

47

(1)

Indicates the 1st -order solution, i.e., the perturbation expressions

47

(1)

Indicates the previous step in a nonlinear free surface SHIPFLOW iteration

correct

Indicates a result of the correction step

. . . . . . . . .
. . .

. . . . . . . . . . . . . . . . . . . . . .

J. Heimann CFD Based Optimization of the Wave-Making Characteristics of Ship Hulls

81
121

189

FRIENDSHIP SYSTEMS News <www.friendship-systems.com/news.php>:

+++ 05/03/02 +++


We congratulate our team member Justus Heimann on his successful doctoral defense. His
thesis is entitled "CFD Based Optimization of the Wave-Making Characteristics of Ship
Hulls" and covers the hydrodynamic design of fast hulls independent of the underlying CAD
modeling technique. Well done Justus!

+++ 04/11/15 +++


FRIENDSHIP SYSTEMS welcomes Dipl.-Ing. Justus Heimann as permanent member of its
staff.
Justus extends our expertise in Computational Fluid Dynamics and optimization. As a
freelancer Justus has already been associated with FRIENDSHIP SYSTEMS for more than
two years while allocating his main resources to his research on wave cut analysis and hull
form optimization. Justus has developed a novel hull optimization technique based on wave
cut analysis and perturbation of the nonlinear free surface waves which nicely complements
FRIENDSHIP SYSTEMS ' established parametric approach.
Justus Heimann looks back at many years as scientist at the Institute of Land and Sea
Transport Systems of the Technical University Berlin. Having graduated in naval architecture
from TU Berlin in 1993, Justus became a scientist with Prof. Dr.-Ing. Gnther Clauss and
later with Prof. Dr.-Ing. Dr. h.c. Horst Nowacki. His experience comprises both teaching and
research. He participated in many national and European research projects with focus on
computational geometry, numerical simulation of calm-water performance and seakeeping.

Vous aimerez peut-être aussi