Sei sulla pagina 1di 262

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

net/publication/258128071

A non-hydrostatic model for free surface flows

Thesis · January 1999

CITATIONS READS
41 316

1 author:

Jacek A. Jankowski
Bundesanstalt für Wasserbau
16 PUBLICATIONS   215 CITATIONS   

SEE PROFILE

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

No projects currently View project

All content following this page was uploaded by Jacek A. Jankowski on 27 June 2016.

The user has requested enhancement of the downloaded file.


Referent: Prof. Dr.-Ing. Werner Zielke
Korreferent: Prof. Dr.-Ing. habil. Rainer Helmig

Tag der Promotion: 12. November 1998


Institut für Strömungsmechanik
und Elektron. Rechnen im Bauwesen
der Universität Hannover

Bericht Nr. 56/1999

Jacek A. Jankowski

A non-hydrostatic model for free surface flows


ISSN 0177 — 9028

Institut für Strömungsmechanik


und Elektronisches Rechnen im Bauwesen
Universität Hannover
Appelstraße 9A, D – 30167 Hannover

Tel.: +49 – 511 – 762-3568


Fax.: +49 – 511 – 762-3777
URL: http://www.hydromech.uni-hannover.de

Von dem Fachbereich für Bauingenieur- und Vermessungswesen


der Universität Hannover zur Erlangung des Grades eines Doktor – Ingenieurs
genehmigte Dissertation.

Referent: Prof. Dr.-Ing. Werner Zielke


Korreferent: Prof. Dr.-Ing. habil. Rainer Helmig

Tag der Promotion: 12. November 1998


Abstract i

Abstract
An algorithm for solution of three-dimensional Navier-Stokes equations for incompress-
ible free surface flows is developed. A decoupled algorithm based on the fractional step
(operator-splitting) technique is applied. The solution is obtained in subsequent stages
treating equations split into parts having well-defined mathematical properties, so that
the most adequate methods for a given differential operator type can be used. The de-
coupled algorithm structure, which does not use the continuity equation explicitly, allows
application of equal-order linear interpolation functions for all variables. The applied
reference element type, a prism with six nodes and linear interpolation functions, is a
compromise between the exactness of the interpolation, model complexity and compu-
tational cost. The finite difference method is applied for the time discretisation and the
computational domain variability is taken into account by a standard σ-mesh structure
which is well suited to most geophysical applications.
The model is implemented in the framework of the Telemac system, developed in Lab-
oratoire National d’Hydraulique in Chatou, Electricité de France (EDF) for modelling
of free surface flows including transport phenomena. It is a far-reaching further devel-
opment of existing three-dimensional code Telemac3D whose application domain is
limited due to the hydrostatic approximation and because the free surface position is
computed from shallow water equations. Therefore, the vertical acceleration is neglected
and the free surface and bottom gradients must be small.
In the new model, the vertical acceleration is taken into account by decomposing the
global pressure into its hydrostatic (i.e. barotropic as well as baroclinic) and hydrody-
namic components. In consequence, the hydrodynamic pressure is found from a pressure
Poisson equation from fractional formulation, whereby the hydrostatic part is computed
explicitly from the free surface elevation and density field. The final velocity is obtained
under assumption of its incompressibility from the projection of the intermediate result,
which is found without hydrodynamic pressure gradients.
The free surface position can be found using height function methods based on the kine-
matic boundary condition or the free surface conservative equation (vertically integrated
continuity equation). In the numerical solution algorithm the method of characteristics,
streamline upwind Petrov-Galerkin FEM, or explicit formulation in finite elements can
be chosen.
The developed new algorithm steps have been verified separately, using analytical solu-
tions. The verification of the algorithm as a whole has been performed using a number
of benchmark test cases covering the targeted model application domain. Some of these
allowed a formal comparison with an analytical problem solution. They include free
surface and internal waves, sub- and supercritical channel flow over a steep ramp, wind-
and buoyancy-driven currents.
The implementation of the model within the well-validated Telemac system allows im-
mediate application of the newly developed algorithm for practical hydraulic engineering
problems and geophysical flows.
Keywords: hydrodynamic-numeric model, non-hydrostatic, free surface flow.
ii Zusammenfassung

Zusammenfassung
Im Rahmen der Arbeit wurde ein Algorithmus zur Lösung der dreidimensionalen Navier-
Stokes Gleichungen für inkompressible Fluide mit freier Oberfläche entwickelt. Der Algo-
rithmus basiert auf einer fractional step Methode, wobei die Grundgleichungen in Anteile
mit gleichen mathematischen Eigenschaften aufgespalten werden, die dann mit optimal
angepaßten Verfahren schrittweise gelöst werden (operator splitting). Es werden für alle
Variablen die gleichen linearen Ansatzfunktionen verwendet. Das prismatische 6-Knoten
Element stellt einen Kompromiß zwischen Genauigkeit der Interpolation, Komplexität
des Modells und benötigter Rechenzeit dar. Die zeitliche Diskretisierung wird mit einem
Finite-Differenzen Verfahren unter Zugrundelegung der σ-Transformation durchgeführt,
was sich für die meisten geophysikalischen Problemstellungen als geeignete Vorgehens-
weise erwiesen hat.
Der neue Algorithmus ist Bestandteil des Programmsystems Telemac, das von der
Electrité de France (EDF) im Laboratoire National d’Hydraulique in Chatou für die
Simulation der Strömungen mit freier Oberfläche und Stofftransport entwickelt wurde.
Das neue Modell stellt eine Weiterentwicklung des bereits existierenden dreidimensio-
nalen Codes Telemac3D dar, dessen Anwendungsbereich infolge einer hydrostatischen
Druckapproximation beschränkt ist. Bislang konnten keine vertikalen Beschleunigungen
berücksichtigt werden und nur kleine Gradienten der freien Oberfläche und des Bodens
behandelt werden.
Im neuen Algorithmus wird die vertikale Beschleunigung durch eine Zerlegung des glo-
balen Drucks in einen hydrostatischen (d.h. barotropen sowie baroklinen) und einen
hydrodynamischen Anteil bestimmt. Im Rahmen der fractional step Methode wird die
hydrodynamische Komponente aus der Lösung einer Poisson-Gleichung ermittelt, und
der hydrostatische Druck explizit aus der Position der freien Oberfläche und dem Dichte-
feld berechnet. Das Geschwindigkeitsfeld wird unter Annahme der Inkompressibilitätbe-
dingung aus einer Projektion des Zwischenergebnisses bestimmt, welches ohne hydrody-
namischen Anteil berechnet wurde.
Die Lage der freien Oberfläche kann aus der kinematischen Randbedingung oder aus
einer konservativen Formulierung (vertikal gemittelte Kontinuitätsgleichung) ermittelt
werden. Zur numerischen Lösung stehen wahlweise ein Charakteristiken Verfahren, eine
semi-implizite SUPG FEM oder eine explizite Formulierung zur Verfügung.
Alle entwickelten Teile des neuen Algorithmus’ sind getrennt mittels analytischer Lösun-
gen verifiziert worden. Der Gesamtalgorithmus ist anhand zahlreicher Testbeispiele aus
dem erweiterten Anwendungsbereich des Modells verifiziert worden, wozu in vielen Fällen
analytische Lösungen herangezogen werden konnten. Es sind unter anderem interne und
Oberflächenwellen, Kanalströmungen über eine Rampe im strömenden und schießenden
Zustand sowie wind- und dichtegetriebene Strömungen behandelt worden.
Die Implementierung des neuen Algorithmus’ innerhalb des etablierten Programmsy-
stems Telemac ermöglicht einem großem Anwenderkreis die sofortige Nutzung für prak-
tische Problemstellungen des Ingenieurwesens und der Geophysik.
Schlagworte: hydrodynamisch-numerisches model, nicht-hydrostatisch, Freispiegelströmung.
Vorwort iii

Vorwort
Die vorliegende Arbeit ist in den Jahren 1995–1998 während meiner Tätigkeit im Institut
für Strömungsmechanik und Elektronisches Rechnen im Bauwesen entstanden.
Auf die Möglichkeit, ein nicht-hydrostatisches Modell auf der Basis von Bibliotheken
des Programmsystems Telemac zu entwickeln, wurde ich erst im Mai 1995 aufmerk-
sam gemacht. Im Herbst folgten die ersten Fortran-Zeilen und im Januar 1996 gab es
schon eine erste lauffähige Programmversion. Was ich zuerst als eine kleine und sehr un-
systematisch ausgeübte Nebenbeschäftigung bei der Bearbeitung eines BMBF-Projektes
Mesoskalige Stofftransporte im Pazifik als Folge der technischen Eingriffe in die Tiefsee
(1992-98) betrachtet habe, hat sich in 1996-97 mit jedem neuen berechneten Verifikati-
onsbeispiel zu einer ruhigen Passion entwickelt. Der Text dieser Arbeit ist im wesentli-
chen auch in diesen Jahren entstanden. Im Winter und Frühjahr 1998 wurde das Modell
neu programmiert und mit vielen neuen zusätzlichen Verbesserungen ausgestattet, um
anspruchsvollere Verifikationen durchführen zu können. Die Dissertation habe ich am
letzten Tag im Juni 1998 abgegeben.
Im ersten ruhigeren und zwanglosen Rückblick auf meine Arbeit sehe ich als ihren
Schwerpunkt die Eröffnung der Möglichkeit der physikalisch und numerisch korrekten
FEM-Modellierung der dreidimensionalen Freispiegelströmungen mit einem entkoppel-
ten Algorithmus.
Diese Arbeit ist allen denen gewidmet, die mittels einer inneren Überzeugung ihre be-
scheidenen Kräfte sammeln, um trotz des sie umgebenden Skeptizismus’ und Besserwis-
serei etwas Neues und Gutes auf diese Welt zu bringen. In diesem Zusammenhang ist die
Bedeutung der auf der nächsten Seite zitierten uralten östlichen Weisheit zu verstehen.
Obwohl ich die meisten Zeit als selbständiger Einzelkämpfer mit den Problemen, die mir
diese Arbeit bereitete, beschäftigt war, gibt es eine Reihe von Personen, bei denen ich
mich bedanken möchte. Die Danksagungen sind auf den letzten Seiten dieser Arbeit zu
finden.

Hannover, Dezember 1998


Jacek A. Jankowski
iv Kalama Sutra

From Kalama Sutra

The Buddha’s Advice


to the Kalamas of Kesaputta [17]

[...]
3. The Kalamas who were inhabitants of Kesaputta sitting on one side said to the
Blessed One:
There are some monks and Brahmins, Venerable Sir, who visit Kesaputta. They
expound and explain only their own doctrines; the doctrines of others they despise,
revile, and pull to pieces. Some other monks and Brahmins too, Venerable Sir, come
to Kesaputta. They also expound and explain only their own doctrines; the doctrines
of others they despise, revile, and pull to pieces. Venerable Sir, there is doubt, there
is uncertainty in us concerning them. Which of these reverend monks and Brahmins
spoke the truth and which falsehood?

4. It is proper for you, Kalamas, to doubt, to be uncertain; uncertainty has arisen in


you about what is doubtful. Come, Kalamas.
Do not go upon what has been acquired by repeated hearing; nor upon tradition; nor
upon rumour; nor upon what is in a scripture; nor upon surmise; nor upon an axiom;
nor upon specious reasoning; nor upon a bias towards a notion that has been pondered
over; nor upon another’s seeming ability; nor upon the consideration, ’The monk is our
teacher.’
Kalamas, when you yourselves know: ’These things are bad; these things are blamable;
these things are censured by the wise; undertaken and observed, these things lead to
harm and ill,’ abandon them.
[...]
10. Come, Kalamas.
Do not go upon what has been acquired by repeated hearing; nor upon tradition; nor
upon rumour; nor upon what is in a scripture; nor upon surmise; nor upon an axiom;
nor upon specious reasoning; nor upon a bias towards a notion that has been pondered
over; nor upon another’s seeming ability; nor upon the consideration, ’The monk is our
teacher.’
Kalamas, when you yourselves know: ’These things are good; these things are not blam-
able; these things are praised by the wise; undertaken and observed, these things lead
to benefit and happiness,’ enter on and abide in them.
[...]
Contents

Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . i
Zusammenfassung . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ii
Vorwort . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iii
From Kalama Sutra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv

1 Introduction 1
1.1 Motivation and aim of the work . . . . . . . . . . . . . . . . . . . . . . . 1
1.1.1 Introducing remarks . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1.2 Why a non-hydrostatic code? . . . . . . . . . . . . . . . . . . . . 2
1.1.3 Difficulties to be overcome . . . . . . . . . . . . . . . . . . . . . . 4
1.2 Literature review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.2.1 Models for three-dimensional incompressible shallow water flows . 5
1.2.2 Available methods for incompressible free surface flows . . . . . . 6
1.2.3 Non-hydrostatic free surface flow models . . . . . . . . . . . . . . 8
1.3 Structure of the work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

2 Solution methods for three-dimensional free surface flow equations 13


2.1 Formulation of governing equations . . . . . . . . . . . . . . . . . . . . . 14
2.1.1 Basic equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.1.2 Incompressibility . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.1.3 Turbulent stress tensor . . . . . . . . . . . . . . . . . . . . . . . . 16
2.1.4 Boussinesq approximation . . . . . . . . . . . . . . . . . . . . . . 17
2.1.5 Non-inertial co-ordinate system fixed to the earth . . . . . . . . . 18
2.1.6 Equation set for geophysical, incompressible flows with Boussinesq
approximation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.2 Hydrostatic approximation . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.2.1 Scaling the vertical equation of motion . . . . . . . . . . . . . . . 19
2.2.2 Hydrostatic approximation and its consequences . . . . . . . . . . 21
2.2.3 Three-dimensional shallow water equations . . . . . . . . . . . . . 21
2.3 Algorithm with the hydrostatic approximation . . . . . . . . . . . . . . . 22
2.3.1 Operator splitting . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.3.2 Advection and diffusion steps . . . . . . . . . . . . . . . . . . . . 23
2.3.3 Pressure – free surface – continuity step . . . . . . . . . . . . . . 24
2.4 Numeric considerations concerning the solution method . . . . . . . . . . 27
2.4.1 General solution methods for 3D Navier-Stokes equations . . . . . 27
2.4.2 Overview of the solution methods with FEM . . . . . . . . . . . . 30

v
vi Contents

2.4.3 Equal and mixed interpolation . . . . . . . . . . . . . . . . . . . . 37


2.4.4 Method applied in the developed model . . . . . . . . . . . . . . . 41
2.5 Pressure treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
2.5.1 Deriving of the conventional pressure Poisson equation . . . . . . 41
2.5.2 General idea of the projection method . . . . . . . . . . . . . . . 42
2.5.3 Pressure equation from fractional step formulation . . . . . . . . . 45
2.5.4 Poisson equation for the hydrodynamic pressure in free surface flows 47
2.5.5 Treatment of pressure and buoyancy terms . . . . . . . . . . . . . 48
2.6 Free surface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
2.6.1 Marker-and-cell method . . . . . . . . . . . . . . . . . . . . . . . 50
2.6.2 Volume of fluid method . . . . . . . . . . . . . . . . . . . . . . . . 51
2.6.3 Height function methods . . . . . . . . . . . . . . . . . . . . . . . 54
2.6.4 Line segment method . . . . . . . . . . . . . . . . . . . . . . . . . 55
2.6.5 Combinations of different methods . . . . . . . . . . . . . . . . . 56
2.6.6 Treating variable domain extents . . . . . . . . . . . . . . . . . . 56
2.6.7 Implemented free surface algorithms . . . . . . . . . . . . . . . . 57
2.7 Initial and boundary conditions for the flow . . . . . . . . . . . . . . . . 57
2.7.1 Initial conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
2.7.2 Dynamic boundary conditions . . . . . . . . . . . . . . . . . . . . 59
2.7.3 Impermeability: normal velocity at solid boundaries . . . . . . . . 61
2.7.4 Tangential velocity at solid boundaries . . . . . . . . . . . . . . . 62
2.7.5 Rigid lid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
2.7.6 Bottom and free surface stresses . . . . . . . . . . . . . . . . . . . 63
2.7.7 Open boundary conditions . . . . . . . . . . . . . . . . . . . . . . 63
2.7.8 Non-reflecting boundary conditions . . . . . . . . . . . . . . . . . 65
2.7.9 Absorbing boundary conditions . . . . . . . . . . . . . . . . . . . 69
2.7.10 Boundary conditions for the tracer . . . . . . . . . . . . . . . . . 69
2.7.11 Boundary conditions for the hydrodynamic pressure equation . . . 70
2.8 Algorithm for the non-hydrostatic equations . . . . . . . . . . . . . . . . 71
2.8.1 Equation set to be solved . . . . . . . . . . . . . . . . . . . . . . 71
2.8.2 Introducing remarks . . . . . . . . . . . . . . . . . . . . . . . . . 72
2.8.3 Operator splitting . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
2.8.4 Advection step . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
2.8.5 Diffusion step . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
2.8.6 Continuity step . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
2.8.7 Free surface step . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
2.8.8 Free surface from vertically integrated momentum equations . . . 78
2.8.9 Projection-2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78

3 Numeric algorithm description 81


3.1 Computational domain . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
3.1.1 σ-mesh and σ-transformation . . . . . . . . . . . . . . . . . . . . 81
3.1.2 Comments on the σ-transformation . . . . . . . . . . . . . . . . . 82
3.1.3 Reference element and interpolation functions . . . . . . . . . . . 83
3.1.4 Two-dimensional elements . . . . . . . . . . . . . . . . . . . . . . 86
Contents vii

3.2 Operator splitting and time discretisation . . . . . . . . . . . . . . . . . . 87


3.3 Continuity step . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
3.3.1 Poisson equation solver . . . . . . . . . . . . . . . . . . . . . . . . 88
3.3.2 Formulation of the pressure Poisson equation in the finite element
method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
3.3.3 Resulting divergence-free velocity field . . . . . . . . . . . . . . . 90
3.3.4 Source term of the pressure Poisson equation . . . . . . . . . . . . 91
3.4 Diffusion step . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
3.4.1 Diffusion step: formulation in finite elements . . . . . . . . . . . . 91
3.4.2 Boundary terms in the diffusion step . . . . . . . . . . . . . . . . 94
3.4.3 Boundary conditions for the diffusion step . . . . . . . . . . . . . 96
3.4.4 Computation of the source terms . . . . . . . . . . . . . . . . . . 96
3.5 Advection step . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
3.5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
3.5.2 Advection with the method of characteristics . . . . . . . . . . . . 98
3.5.3 Advection with SUPG method . . . . . . . . . . . . . . . . . . . . 100
3.5.4 Boundary conditions for advection step . . . . . . . . . . . . . . . 102
3.6 Free surface step . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
3.6.1 Free surface with the method of characteristics . . . . . . . . . . . 102
3.6.2 SUPG formulation for the kinematic boundary condition . . . . . 102
3.6.3 Explicit formulation for the conservative free surface equation . . 105
3.6.4 Semi-implicit formulation for the conservative free surface equation 106
3.6.5 Free surface stabilisation . . . . . . . . . . . . . . . . . . . . . . . 107

4 Verification of algorithm stages 109


4.1 Diffusion step test: Ekman profiles . . . . . . . . . . . . . . . . . . . . . 109
4.1.1 Slope current . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
4.1.2 Drift current . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
4.1.3 Numerical test: drift and slope current . . . . . . . . . . . . . . . 113
4.2 Tests of the Poisson equation step . . . . . . . . . . . . . . . . . . . . . . 115
4.2.1 Testing the boundary conditions influence . . . . . . . . . . . . . 116
4.2.2 Testing the source term influence . . . . . . . . . . . . . . . . . . 117
4.3 Tests of the free surface step . . . . . . . . . . . . . . . . . . . . . . . . . 119
4.3.1 Kinematic boundary condition . . . . . . . . . . . . . . . . . . . . 119
4.3.2 Conservative free surface equation . . . . . . . . . . . . . . . . . . 128

5 Model verification 131


5.1 Waves reflection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
5.2 Wind-driven circulation . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
5.3 Interfacial internal waves . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
5.3.1 Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
5.3.2 Test realisation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
5.4 Lock exchange flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
5.4.1 Hydrostatic interpretation . . . . . . . . . . . . . . . . . . . . . . 144
5.4.2 Lock exchange flow speed . . . . . . . . . . . . . . . . . . . . . . 147
viii Contents

5.4.3 Lock exchange flow – further considerations . . . . . . . . . . . . 148


5.4.4 Test realisation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
5.5 Standing wave in a closed basin . . . . . . . . . . . . . . . . . . . . . . . 155
5.5.1 Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156
5.5.2 Deep water – short waves . . . . . . . . . . . . . . . . . . . . . . 157
5.5.3 Shallow water – long waves . . . . . . . . . . . . . . . . . . . . . 160
5.5.4 Comparison of different free surface algorithms . . . . . . . . . . . 161
5.6 Subcritical and supercritical flow over a ramp . . . . . . . . . . . . . . . 165
5.7 Channel with a bump . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168
5.8 Waves travelling over an underwater channel . . . . . . . . . . . . . . . . 173
5.9 Solitary wave in a long channel . . . . . . . . . . . . . . . . . . . . . . . 178
5.9.1 Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178
5.9.2 Solitary wave propagation over a flat bottom . . . . . . . . . . . . 179
5.9.3 Solitary wave in a long channel with varying depth . . . . . . . . 185
5.9.4 Collision between two solitons . . . . . . . . . . . . . . . . . . . . 191
5.10 Limits of application . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 192
5.10.1 Free surface breaking . . . . . . . . . . . . . . . . . . . . . . . . . 192
5.10.2 Bottom gradients for supercritical flow . . . . . . . . . . . . . . . 193
5.10.3 Breaking of a solitary wave . . . . . . . . . . . . . . . . . . . . . . 195

6 Summary and Conclusions 199


6.1 Realised new developments . . . . . . . . . . . . . . . . . . . . . . . . . . 199
6.2 Result discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201
6.3 Model applications and limitations . . . . . . . . . . . . . . . . . . . . . 203
6.4 Recommended further developments . . . . . . . . . . . . . . . . . . . . . 204

A Various derivations 207


A.1 Incompressibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207
A.2 Coriolis terms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208
A.3 Small amplitude free surface waves in a homogeneous fluid . . . . . . . . 211
A.4 Equations of state for seawater . . . . . . . . . . . . . . . . . . . . . . . 215
A.4.1 EOS–80: UNESCO equation of state . . . . . . . . . . . . . . . . 215
A.4.2 Simple forms of the equation of state . . . . . . . . . . . . . . . . 216
A.4.3 Suspended sediment influence on the water density . . . . . . . . 219
A.5 Meshes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
A.6 Systematic approach to model development . . . . . . . . . . . . . . . . . 219

Bibliography 223

Nomenclature 235

Danksagung 244
Chapter 1

Introduction

1.1 Motivation and aim of the work

1.1.1 Introducing remarks

The physical modelling of free surface flows in a hydraulic laboratory has a very long
tradition, but in the recent 30 to 40 years the numerical simulation of free surface flows
gained more and more in importance. As numerical methods of computational fluid
dynamics (CFD) developed to a considerable level and the computational power of com-
puters compared to their price grew as well, it seems that the physical models are getting
too expensive and outdated. Indeed, their role transformed to the provision of verifica-
tion, calibration and support data sets for numerical modelling, especially for specific
flow configurations or aspects. For modelling flows on the geophysical scale, such as
whole rivers or lakes, or for assessing the results of anthropogenic activities in nature, the
hydrodynamic-numerical models are nowadays indispensable. Supported by numerous
advanced pre- and post-processing programs, the commercially available computational
fluid dynamics codes became standard engineering tools.
There are also some negative aspects of this situation. The rapid development of user-
friendly interfaces makes even the most complex codes available also to non-specialists.
Colourful animations help to obtain a wider public approval of the results. In effect,
the risk of misuse and error has grown considerably, since the limitations of the numeri-
cal algorithms and other uncertainties are not adequately taken into consideration and
clearly presented. The currently available tools are still far from being entirely satisfac-
tory. There are still numerous flow cases, especially three-dimensional free surface flows,
which cannot be simulated without a non-standard approach.
The trend to take advantage of the possibilities of new computers yielding higher and
higher accuracy, larger and larger storage and ever-growing computation speed is very
important, but one cannot rely on them as the only source of improvements. Progress
will be achieved when a considerable effort is also invested in improving, extending or
even replacing basic algorithms by newer, better ones.

1
2 1.1 Motivation and aim of the work

Another major issue in computational fluid dynamics is securing the quality of the codes.
The modern programs consist of large procedure or module libraries concerning various
disciplines (e.g. hydrodynamics, water quality, sedimentology) and controlling the overall
quality may be very difficult. Guidelines for widely acknowledged methodology for vali-
dation processes, which should be absolved by the computer codes, are presently being
prepared by national and international associations connected with hydraulic research.

1.1.2 Why a non-hydrostatic code?

At present, one- and two-dimensional numerical models are widely used in engineering
practice. As computer capacities grew, the application of computationally more intensive
three-dimensional models became feasible. Some of these models are thought of as an
extension of two-dimensional (vertically integrated) models in order to describe selected,
but not all, three-dimensional properties of the flow. In most cases, where the vertical
acceleration component is small, sufficient accuracy for simulating most geophysical free
surface flows is obtained by using numerical models, which assume the hydrostatic pres-
sure approximation. It simplifies the coupled three-dimensional set of equations to be
solved and therefore allows development of computationally efficient three-dimensional
codes. The hydrostatic models are very successful, for example, in describing horizon-
tal velocity profiles in vertical direction or stratified currents in mostly horizontal flows.
Nevertheless, these models cannot be pushed beyond their limits of application.
There are situations, where the vertical acceleration, and therefore the non-hydrostatic
pressure component cannot be neglected. This is the case when, for example, flows
over abruptly changing bottom topography, orbital movements in short wave motions,
or intensive vertical circulations are to be considered. Generally, the application of
hydrostatic approximation is disputable, when the ratio of vertical to horizontal motion
scales (velocities and distances) is not small. This approximation is neither appropriate
in flows around obstacles, when the pressure at the boundary is not hydrostatic, nor
when the flow is determined by stronger density gradients enhancing vertical movements.
The motivation of this work is to describe properly these free surface flows, where the
hydrostatic approximation is no longer valid.
In many cases assumptions due to the basic equations form or to the applied numerical
techniques (especially the mesh structure) cause limitations on the domain boundary
geometry. A classical example is the severe constraint on the bottom or free surface
steepness in models using shallow water equations. Other geometrical assumptions in-
fluence strongly the treatment of the moving domain boundaries or surfaces separating
fluids with different properties. An additional consequence of using the shallow water
equations is that they provide correct solutions for long waves only.
The aim of this work is to develop a three-dimensional numerical model for these incom-
pressible geophysical free surface flows, where the solution of entirely three-dimensional
Navier-Stokes equations is required. The limitations of hydrostatic codes are to be over-
come by taking the unchanged momentum equation for the vertical velocity component
into consideration. The free surface movement computation should not be constrained
1.1 Motivation and aim of the work 3

to the long waves only. Development should take place in the framework of an existing,
advanced, and well-validated model system, allowing immediate applications in complex
geometries not only for the hydrodynamics, but also for e.g. transport phenomena and
taking turbulence models of various complexity into consideration.

Therefore, an existing finite element three-dimensional hydrostatic code, Telemac-3D,


was taken as the starting point [74]. Precisely, it solves the three-dimensional, shallow-
water, Reynolds-averaged, Navier-Stokes equations for incompressible, turbulent, free
surface flows with Boussinesq approximation. To the typical application domain of this
code belong geophysical free surface flows with complex geometry, where the full ad-
vantage of the finite element approach can be taken. Telemac system was developed
by Electricité de France (Laboratoire National d’Hydraulique) using strict quality pre-
cautions. It has been well validated in numerous practical, mostly fluvial and marine
applications including some advanced ones, as for example, the sediment transport in
tidal estuaries [103, 110, 111].

Historically, Telemac3D code was developed from a two-dimensional code


Telemac2D [52]. For instance, its two-dimensional routines are applied in order to
compute the free surface movements in the three-dimensional code. Telemac3D is
characterised by the operator-splitting technique of solving the hydrodynamic equations
and the modular structure of the algorithm based on libraries of numerical procedures of
a broad application scope which include modern equation solvers. This open structure
allows efficient development of a new algorithm by replacing or building in additional
steps to the existing algorithm frame.

The main limitations of the Telemac3D code, encountered by applications at Institut


für Strömungsmechanik und ERiB were connected with hydrodynamics of areas with
steep bottom gradients and buoyancy-driven currents, for example, in estuaries [110].
Similar, though less serious problems, were found when forecasting sediment transport by
bottom-near deep ocean currents over variable bathymetry, where neglecting the vertical
acceleration is not appropriate [84, 85].

Particular situations, where the models using the hydrostatic approximation are inad-
equate, are numerous. Examples include: interaction of currents with structures (e.g.
pillars, pipelines); flows over larger bottom gradients; various aspects of sediment trans-
port (erosion and consequent deposition, e.g. travelling bottom dunes or erosion behind
bridge pillars); complex three-dimensional currents (e.g. secondary currents in river
meanders) and finally a wider spectrum of internal and surface wave problems (short,
non-hydrostatic waves, e.g. in connection with deposition and erosion of sediments in
harbours).

Most of the problems mentioned above concern flows in complex geometries, where the
finite element method is the most effective one. A need to open the possibility of mod-
elling these flows in the framework of the existing finite element models is clearly seen.
The real challenge, however, is to provide a hydrodynamic-numeric model which could
simulate the phenomena mentioned above simultaneously and still remain computation-
ally feasible for practical applications.
4 1.2 Literature review

1.1.3 Difficulties to be overcome

The difficulties encountered by the development of fully three-dimensional numeric-


hydrodynamic codes with a variable free surface are manifold. But three of them are the
most important.
First of all, in the past the prohibitive computation times reduced the application of
the three-dimensional models only to academic test cases, hampering broader interest in
them. Nowadays, new efficient algorithms for solving large equation systems and, above
all, the available computational power allow application of these very complex models
for scientific purposes. However, engineering practice also requires for those special cases
that cannot be dealt with without an entirely three-dimensional approach, that the codes
remain computationally feasible.
Secondly, the complexity of these models is greater than awaited by a simple extension to
a third dimension or by taking another equation into consideration. The hydrodynamic
equations describing flows in domains with variable geometries are coupled in a complex
way. Especially the coupling of the movements of the free surface and the interior
solution brings a lot of difficulties. The changeable free surface not only defines the new
computational geometry, but also influences the solution through boundary conditions
for all flow variables. Last, but not least, the barotropic pressure gradients caused by
free surface slopes are the main driving force for the shallow water flows. Additionally,
the large diversity of presently available numerical methods makes it difficult to choose
the optimal ones.
The third source of difficulty is the fact, that not only equations, but also boundary
conditions for flow variables are complex and coupled as well, and must be treated with
the greatest care in order to secure a stable and physically meaningful solution. The
boundary conditions for the pressure belong to the most difficult to be properly posed.
Additionally, the boundary conditions must be applied on the moving boundaries. Last,
but not least, the initial conditions must fulfil important mathematical constraints.

1.2 Literature review

In contrast to the flows with fixed boundaries, a basic requirement for analysing free sur-
face flows is the determination of both velocity and pressure fields as well as the variable
position of the free surface simultaneously. The shape of the deformable domain is a part
of the solution to be achieved. The literature on the subject is large and a selection of
relevant domains of interest as well as concentration on ideas and good examples, not a
complete and chronological development description is necessary. Leaving specific details
for next chapters, the following literature review concentrates on the most important and
relevant aspects of the model presented in this work. For clarity and briefness, the vast
field of internal (incompressible or compressible) flows and turbulence modelling is left
(almost) untouched.
1.2 Literature review 5

1.2.1 Models for three-dimensional incompressible shallow wa-


ter flows
One-dimensional (1D) and two-dimensional (2D, e.g. depth-integrated) models used
in most geophysical and engineering applications are very well advanced and achieve
satisfactory results in their domain of application [1, 166]. Their advantage is and will
be computational efficiency and easy implementation with satisfactory accuracy for most
(for one dimension, usually fluvial) practical applications. However, an important aspect
that must be taken into consideration while applying these models is how to parameterise
three-dimensional effects appropriately. As illustrative examples of well validated modern
models for two-dimensional shallow water equations, the finite difference Trim2D model
developed by Casulli [19], and finite element model Telemac2D developed by Galland,
Goutal and Hervouet [52] or Tisat developed by Lang [97] should be mentioned.
The first applicable three-dimensional (3D) models were developed as early as 1970 [34].
From this time on, 3D models found their way into various geophysical and engineering
disciplines, but their application remained limited due to their usually high computa-
tional needs.
For free surface flow, a well established type of model for three-dimensional shallow water
flow (i.e. with hydrostatic approximation) exists [68, 120]. The majority of these models
are developed as an extension of two-dimensional models in order to describe some chosen
three-dimensional properties of the flow. The attractivity and strength of these models
lies in the possibility of taking various boundary conditions at the free surface (wind,
thermal radiation) and at the bottom (shear) into consideration in a natural way, so that
bottom and free surface layers are resolved. They also allow modelling of stratified flows
(baroclinic pressure gradients).
A good example of a three-dimensional, finite difference (FD) model with σ-co-ordinates,
turbulence modelling, the hydrostatic approximation and applied to coastal ocean prob-
lems was presented by Blumberg and Mellor (Princeton Ocean Model, POM) [11]. A
special strategy, allowing separation of baroclinic and barotropic motion modes (mode
splitting) is available to reduce computational cost. Another example of such a model
is the one used to compute tidal flows with wind influence in the Irish Sea by Davies and
Lawrence [33].
An extension of the Trim2D model to three dimensions, Trim3D by Casulli at al.
[23, 27, 22] is probably the best example of an optimally developed FD model for three-
dimensional shallow water equations. It uses a semi-implicit formulation – the gradients
of the free surface and velocity divergence are finite-differenced implicitly and the re-
maining terms explicitly. The convective terms are treated by a Langrangian-Eulerian
method. This combination of semi-implicit formulation for propagation of gravity waves
and Lagrangian methods yields an unconditionally stable model, well-suited for engi-
neering purposes, especially for coastal regions with tidal flats. However, when baro-
clinic forcing (stratification) must be taken into consideration, there are limits to this
model’s stability. The assumption of extremely simple, structured and rectangular mesh
of equally-spaced squares allows a formulation of a very efficient and fully vectorisable
code, but requires care when modelling flows in complex geometries.
6 1.2 Literature review

Telemac3D, developed in Laboratoire National d’Hydraulique in Chatou, France,


(EDF), represents a state-of-the-art finite element (FE) model for three-dimensional
shallow water equations [83]. It is based on the operator-splitting scheme, where the
governing equations are split into parts having clear mathematical properties and treated
using the most appropriate methods: e.g. semi-implicit FEM for diffusive parts, or a
Lagrangian approach (method of characteristics) for advective parts. For computational
efficiency, the simplest prismatic 3D finite elements with linear functions and unstruc-
tured mesh based on the σ-co-ordinate system are applied. For treatment of the resulting
matrices, a fully vectorisable element-by-element method is implemented. For additional
treatment of non-linearities iterative procedures are provided. The overall stability of
this method is guaranteed, when the consecutive algorithm steps are also stable. The
computational effort compared to the FDM-based models is surely larger, but FEM uses
all the advantages of an unstructured mesh for resolving complex flow geometries.
The 3D shallow water models use various formulations for finding free surface position
– mostly from the vertically integrated continuity equation (conservative free surface
equation), as Trim3D or POM, or from two-dimensional shallow water equations, as
Telemac3D.

1.2.2 Available methods for incompressible free surface flows

The application limits of the three-dimensional shallow water models mentioned above
are twofold. Firstly, the vertical accelerations in flows are neglected. Secondly, especially
for the free surface schemes, the bottom and wave slopes in the main direction of the
flow must be small.
In order to overcome these limitations, an approach free of hydrostatic approximation is
needed for the internal flow. Techniques as independent as possible from the bottom and
free surface gradients are required for tracking the free surface movements (which also
determine the variability of the computational domain). For this purpose, on the one
hand the techniques developed for arbitrary incompressible three-dimensional internal
flows and, on the other hand, schemes developed for more advanced free surface wave
problems become attractive.
Reviews of the present development and state-of-the-art in modelling incompressible
flows are provided for example by Löhner [106] and by Gresho [58]. As examples of
recent outstanding computations, the simulations of flows around cylinders in a duct by
Glowinski and Pironneau [55], simulations of free surface flow extruding from a duct by
Wambersie and Crochet [165] and flows in complex duct geometries by Tu and Fuchs
[158] and Kwak et al. [93] can be mentioned. A discussion of the intensive computational
demands of 3D-codes for fluid dynamics is provided by Nieuwstadt [119].
Four general methods of solving three-dimensional Navier-Stokes equations exist. First of
all, from a direct discretisation of the governing equations a large global equation system
for the primitive variables can be obtained. The resulting equation system cannot be
solved using standard methods, and non-linear, iterative (e.g. Newton–Raphson) solvers
1.2 Literature review 7

must be applied [49, 50, 80, 168]. This direct solution method can be computationally
very intensive, especially for three-dimensional problems.
In a second group of methods, a decoupled approach is preferred, where instead of one
large global equation system a few smaller, linear equation systems are to be solved.
This is achieved by solving the governing equations in consecutive steps. Various related
methods of decoupling exist; operator-splitting [52], pressure correction methods (SIM-
PLE) [124, 123], pressure methods [18, 24], projection methods [29, 57, 143], fractional
step method [129], and finally Taylor methods [37, 38, 86].
A third group of methods, mostly for stationary flows, is based partially on developments
in the domain of compressible flows and introduces artificial coefficients. In the artificial
compressibility method, the set of equations for compressible flows is applied; when
a steady state is achieved, the partial derivative of pressure in the compressible form
of the continuity equation disappears [78, 148]. In the penalty function methods it
is assumed that the pressure changes proportionally to the velocity divergence. The
artificial proportionality parameter, bulk viscosity, must assume very large values. In
theory, the numerical solution converges to the steady state for an infinite bulk viscosity
[28, 133].
The last group of methods uses non-primitive variables, as a stream function or vorticity.
Introducing them into the governing equations one obtains a set of transport and Poisson
equations [3, 47]. The number of equations to be solved is actually reduced, but the form
of boundary conditions may be very complex.
For dealing with flows in stratified systems consisting of layers of water of different
density and physiochemical properties (oceans, lakes) a very promising approach based
on isopycnic co-ordinates has been developed. A model of this type consists of a set
of interior layers of constant density, giving it the character of a stacked shallow water
model. Special techniques are provided for lenses and outcrops of layers and for taking
diapycnal mixing between the layers or up-welling/down-welling into account. An
example is the MICOM model developed by Bleck et al. [9, 10] or a model by Casulli
[21].
For simulations in a time-variable domain, the Arbitrary Lagrangian-Eulerian (ALE)
Method has found greater popularity, especially for FEM applications. The main idea of
this method is to adapt the mesh to the movements of the domain boundaries using an
arbitrarily chosen mesh velocity [36]. The σ-transformation is a simple form of the ALE
scheme, where the mesh spacing varies only in one direction (vertically) and is dictated
by the free surface movements. Another approach is to use a general, deforming space-
time (FE) formulation in which the mesh movement is taken into account automatically
[149].
Tracking of the free surface movements (or movements of an interface between fluids) can
be performed using two general approaches: by balancing or tracking the fluid volume,
or by controlling the free surface movements itself. Mixed methods are also available.
Two methods for volume tracking exist. In principle, both allow arbitrary movements of
the free surface. Massless marker particles moving passively with the fluid are applied to
trace the fluid volume movements in a fixed mesh in the marker-and-cell (MAC) method,
8 1.2 Literature review

developed by Harlow and Welch [65]. The volume-of-fluid (VOF) method, introduced by
Hirt and Nichols [79], uses a function which describes the filling of a given computational
cell with a fluid. This function is advected with the fluid velocity. Both methods require
additional geometrical algorithms to find the actual approximate shape of the interface
in the partially filled, i.e. boundary cells. Method-specific additional techniques must
be noted: an algorithm to redistribute the markers when required (MAC) or special
bookkeeping adjustment to avoid over-filled or over-emptied cells (VOF).
The surface-tracking algorithms use the kinematic boundary condition or the conserva-
tive free surface equation (being a vertically integrated continuity equation, and taking
therefore also the fluid volume into consideration), e.g. [18, 22]. The searched variable
is a height function, i.e. the position of the fluid surface (or interface) relative to a
defined level plane.
The main limitations of these methods are due to the mathematical requirement for the
height function – it must be a single-valued function. In consequence, these methods
cannot simulate breaking waves or very steep wave slopes, but allow tracking of a much
sharper interface than VOF or MAC.

1.2.3 Non-hydrostatic free surface flow models

Not surprisingly, the first efforts to obtain free surface flow models which were free of
limitations due to hydrostatic and/or shallow-water approximation, concerned entirely
three-dimensional circulating flows, wave problems and larger movements of the free
surface.
Leschziner and Rodi computed in 1979 the 3D helicoidal secondary currents in a channel
bend using k − ε turbulence model [104]. They applied a horizontal fictitious plane rep-
resenting a free surface and applied known pressure gradients there in order to simulate
free surface slopes in an indirect way. Frederiksen and Watts [50] developed in 1981
a finite element model with elements having dimensions in time and space to deal with
transient free surface incompressible flow with examples for entrainment and circulating
flows. Kawahara and Miwa [89] presented in 1984 a two-dimensional FEM model able to
simulate travel and run-up of waves on a breakwater. Bulgarelli, Casulli and Greenspan
[18] published in 1984 the pressure finite difference method for incompressible and com-
pressible free surface flows with examples of stratified circulating flows.
A very typical domain, where the non-hydrostatic conditions combined with a Coriolis
effect are very important, are convective flows in the 0geophysical scale, for example
down-welling and up-welling in oceans and lakes. Local convection events in the benthic
ocean caused by episodic hydrothermal discharges from an oceanic ridge were modelled
by Lavelle and Baker [98]. They used a primitive-equation, non-hydrostatic, finite
difference model taking advantage of the axial symmetry of these plumes (confirmed
by measurements), allowing the use of cylindrical co-ordinates (r, z). Therefore, the
equations were integrated in a rectangular (r, z)-grid, reducing the problem to two di-
mensions.
1.2 Literature review 9

Deep water renewal in temperate lakes triggered by a storm surge (when some of the
relatively cold upper water column is forced downwards through its compensation depth
so that it becomes unstable and sinks) was simulated by Walker and Watts [163] with
a high resolution, three-dimensional non-hydrostatic FD Boussinesq model. In order to
model the density driven spiralling (rotating) plumes of the sinking cold surface water,
not only the non-hydrostatic conditions had to be taken into consideration, but also the
fully three-dimensional Coriolis force, i.e. both vertical and horizontal components of
the earth rotation vector. The code was based on a revised semi-implicit method for
pressure linked equations, as in the SIMPLE algorithm family. A similar approach in
FEM was applied by Forkel [48] for investigation of material transport in lakes with
complex bathymetry.
Open ocean deep convection in convective cells due to strong cooling of the ocean surface
were modelled by Sander et al. [136]. They also used a non-hydrostatic, staggered-grid,
finite difference model with both vertical and horizontal Coriolis coefficients taken into
consideration. The algorithm was, as in the previous case, based on pressure-linked
equations.
In almost all these applications, however, the movements due to the gravitational waves,
treated as unimportant for the problem, were eliminated by ‘freezing’ the free surface.
The rigid lid approximation was used, so that only the internal movement was resolved.
The rigid lid approximation restricts therefore the applicability of these models to the
flows where the barotropic component of the hydrostatic pressure can be neglected. For
example, this is not valid in tidal flows.
An algorithm using the idea of pressure decomposition into hydrostatic (i.e. including
barotropic and baroclinic) and hydrodynamic parts for the geophysical flows in a finite
difference implementation was presented by Casulli and Stelling [24]. In their code the
free surface is found from the conservative free surface equation using provisional velocity
fields resulting from the hydrostatic algorithm step and from the previous time-step
solution. The hydrodynamic pressure is found from a Poisson equation. Verification
results of models based on pressure methods were presented recently by Casulli [20],
Casulli and Stelling [25] and Gaarthuis [51].
Developments using FEM-based projection (i.e. decoupled) methods have been reported
very recently by Schröder [140] and Chen [26]. They apply mixed-interpolation finite ele-
ments and use rigid-lid approximation or obtain the free surface from kinematic boundary
condition. The hydrodynamic equations are decoupled by application of an equation for
the global pressure.
A fully three-dimensional oceanographic development for modelling ocean basins with
movements of the free surface was presented by Mahadevan et al. [108]. They introduced
the movements of the free surface in order to capture properly up-welling and down-
welling in the oceanic mesoscale flows (models with open boundaries). They did this
in order to study the thermohaline structure of the ocean that is sensitive to vertical
circulations and bathymetry variability. In conclusion, it was found that it is essential
to retain the commonly neglected component of the Coriolis acceleration in the vertical
momentum equation. The application case for the Gulf of Mexico basin has shown that
10 1.3 Structure of the work

a non-hydrostatic model without the rigid-lid approximation usual in the oceanography


and allowing faster waves performs well in reproducing a flow strongly influenced by
the bathymetry [109]. Additionally, it was demonstrated that a non-hydrostatic model
is well-posed and does not require a special treatment of the open boundaries like a
hydrostatic one (e.g. enhanced viscosities at the inflow and outflow).
Another method of modelling free surface flows with non-hydrostatic effects is to use a
fully coupled method, which solves Navier-Stokes, continuity and free surface (kinematic
boundary condition) equations. This method, implemented by Yost [168] as a finite
element model, uses higher order interpolation functions and is associated with solution
of large equation systems, making it computationally intensive.
An important branch of non-hydrostatic models are two-dimensional models (in the
vertical plane) aimed at wave problems and large free surface movements. Huerta and
Liu [80] apply the the ALE method and mixed SUPG FEM for computation of flows
with large surface movements – tsunami and dam breaks. For solving the resulting
coupled equation system, they apply a predictor-multi-corrector scheme. Ramaswamy
and Kawahara [131, 130] and Ramaswamy [132] develop a model based on the ALE
technique which is able to deal with various wave problems and density currents with a
free surface as well. They use a velocity correction method based on the Poisson pressure
equation and find the free surface from the kinematic boundary condition, so that their
algorithm belongs to the category of decoupled methods.
In FDM, one of the most interesting applications of a decoupled method for numerical
studies of breaking waves and hydraulic jumps was presented by Lemos [102]. The
continuity is secured by iteration between pressure and velocity performed locally in
each computational cell after obtaining an intermediate solution from the momentum
equations, until the residual divergence disappears. The free surface is updated and
defined by the VOF method and the turbulent processes are described by k − ε–model.
Tezduyar, Behr, et al. apply the general, deforming space-time (FE) formulation method
for free surface flows and flows with an interface between two fluids [149, 150, 151].
Behr and Tezduyar also present special FE strategies allowing a feasible application
of this intensively coupled method for larger scale applications using massively parallel
computers [6].

1.3 Structure of the work


In the particular case of the model described in this work, the physical systems to be
simulated are geophysical free surface flows in complex geometries. The model variables
of interest are three or two-dimensional fields: velocity, free surface elevation and con-
centrations of transported substances. Due to the model assumption that the water in
most geophysical flows can be treated as an incompressible, Boussinesq fluid, the pres-
sure can be treated as a subsidiary variable. The transported substances can be divided
into passive tracers, which move passively with the fluid, active tracers, which influence
the fluid density and in consequence the flow, and particles with special properties, as
sediments having their own settling velocity.
1.3 Structure of the work 11

The applied general numerical method, the finite element method, is one of the most
appropriate methods to treat complex flow geometries. The algorithm structure allows
a proper treatment of various boundary conditions appearing in nature. Due to the
assumed structure of the mesh (σ-mesh) another important model assumption is met
corresponding to the free surface (and the bottom as well). It must be represented by a
single value function, excluding, for example, breaking waves problems from the model
application domain. Additionally, the lateral boundaries are represented by vertically
oriented surfaces.
The conceptual model is highly influenced by the numerical method applied to solve the
governing hydrodynamic equation set describing the physical system in a complex, time-
variable geometry and with various boundary conditions. The algorithm of Telemac3D
is based on the operator-splitting technique, resulting in a modular and open structure
of the solution procedure. Parts or stages of the algorithm can be switched on or off
according to the simulation requirements. When new methods are available, these stages
can be consequently replaced, modified or even totally abandoned. The only disadvan-
tage of this approach is a large number of algorithmic or software parameters to steer the
complex program and forcing the user to perform test runs in order to find an optimal
set for best numerical properties for a particular case.
The hydrostatic Telemac3D algorithm assumed additionally that the hydrostatic ap-
proximation is valid and that the free surface can be obtained from shallow water equa-
tions. It limited the model application domain to shallow water flows without larger free
surface and bottom gradients. In consequence, the development of the non-hydrostatic
algorithm can be understood as an extension of the model application domain to more
complex physical systems without changing the basic conceptual model properties (e.g.
concerning the geometry or fluid properties) or algorithm structure.
The existing previously hydrostatic algorithm has been thoroughly verified using various
verification test cases covering the whole targeted application domain for this model, e.g.
an inviscid channel flow with a bump, the Ekman spiral [41] test, an interfacial standing
internal wave, and wind-driven currents, etc. These tests are very formally documented
in a normalised way so that they can be also be used for example as implementation
tests by installing the software on a new computer. Similar tests must vindicate all new
developments to assure model quality.
Therefore, the new algorithm steps have been carefully verified using analytical solutions.
The verification of the algorithm as a whole has been done using test cases, which cover
the targeted model application domain. The benchmark tests allow in most cases a
formal comparison with an analytical problem solution. The tests include free surface
and internal waves, sub- and supercritical channel flow over a steep ramp, wind- and
buoyancy-driven currents.
Accordingly, this work follows the strict development rules of a hydrodynamic-numeric
model:

• Chapter 1 (i.e. this chapter) defines the aim of the work, provides a literature
review of the relevant state-of-the-art methods of modelling free surface flows and
12 1.3 Structure of the work

presents the structure of this work in the light of the systematic approach to a
quality-ensuring hydrodynamic-numeric model development.

• Chapter 2 reviews the basic theory of the incompressible flows as well as discusses
methods of solving the Navier-Stokes equations in more detail with an emphasis on
the finite element method. The techniques for representing and tracking the free
surface and implementing of various boundary and initial conditions are presented.
It supplies a description of the concepts and the algorithms of the previously ex-
isting model and the currently developed one abstracting from their numerical
implementation. This chapter can be treated as a formal definition of the physical
system and description of the conceptual model with all model parameters and
data required.

• Chapter 3 provides more detailed documentation of the applied numerical meth-


ods and software developments.

• Chapter 4 consists of tests of the developed algorithm stages treated as separate


units, due to the modular algorithm structure. It provides a strict software and
algorithm verification with an heuristic investigation of the stability constraints.

• Chapter 5 performs the verification of the model using tests showing its possibili-
ties in reproducing non-hydrostatic free surface and internal flows and investigates
application limits. Additionally, the test cases provide assessments of model accu-
racy and reliability.

• Chapter 6 provides a summary and conclusions, as well as proposals for ap-


plications which would allow formal model validation. Finally, future research
possibilities are sketched.

• Appendix consists of a few sections providing additional information.


Chapter 2

Solution methods for


three-dimensional free surface flow
equations

The pressure is a somewhat mysterious quantity in in-


compressible flows. [...] It is in one sense a mathe-
matical artefact – a Lagrange multiplier that constrains
the velocity field to remain divergence-free; i.e. incom-
pressible – yet its gradient is a relevant physical quan-
tity: a force per unit volume. It propagates at infinite
speed in order to keep the flow always and everywhere
incompressible; i.e. it is always in equilibrium with a
time-varying divergence-free velocity field. [64]
This chapter provides the definitions of the physical system and the conceptual model,
as well as an outline of numerical algorithms applied to solve the three-dimensional free
surface flow equations. It describes:

• the governing equations with a stress on the approximations leading to the par-
ticular form of the basic equations applied in the model (which also define model
limitations);

• algorithms for solving Navier-Stokes equations and considerations concerning the


relevant aspects of the finite element method for decoupled schemes;

• free surface tracking algorithms;

• required specific initial and boundary conditions;

• algorithms of the hydrostatic (previous) and non-hydrostatic (new) models pre-


sented abstracting from their numeric implementation (described in chapter 3).

13
14 2.1 Formulation of governing equations

2.1 Formulation of governing equations

2.1.1 Basic equations


As the starting point, the conventional set of the governing hydrodynamic equations for
turbulent geophysical flows is considered [5, 53, 92, 128]:
Conservation of mass:
1 d̺
+∇·u= 0 (2.1)
̺ dt
Conservation of momentum:
∂u 1
+ u · ∇u = ∇ · σ + g − 2Ω × u + f (2.2)
∂t ̺

Transport equation for a tracer:


∂T
+ u · ∇T = ∇ · (ν T ∇T ) + qT (2.3)
∂t

Equation of state (usually an empirical one):

̺ = ̺(Ti , p) (2.4)

The equation set above is formulated for geophysical free surface flows in the non-inertial,
orthogonal Cartesian co-ordinate system connected with the surface of the earth (Ω is
the earth’s angular speed). For typical applications, x is directed eastward, y northward,
and z points vertically upward in the direction of −g, the acceleration of gravity, figure
2.1. The velocity vector u has components (u, v, w). T describes a tracer: temperature,
salinity, or passive effluent concentration transported by the flow. Tracers may influence
the water density, as the pressure p does (active tracers, noted Ti in eq. 2.4), or not
(passive tracers). The term f represents forces applied on the fluid, other than pressure
and weight.
The momentum conservation equations and the continuity equation (2.1–2.2) are written
in a well-known form of Reynolds-averaged Navier-Stokes (NS) equations for Newtonian
fluids [118, 146, 134], where the Boussinesq eddy-viscosity concept [12] is introduced.
Strictly speaking, the velocity and pressure are stochastic variables devoid of turbulent
fluctuations. The transport equation for the tracer(s) (2.3) coupled through the equation
of state (2.4) with Navier-Stokes equations is introduced to deal with flows with variable
density. The transport equation for tracers has undergone similar treatment concerning
the turbulence concepts. In this formulation, the turbulent stress tensor σ has a form
similar to the one valid for laminar flows of Newtonian fluids:

σ = −pI + 2̺νD (2.5)

ν is the turbulent eddy-viscosity tensor and ν T the turbulent diffusion tensor for tracers.
They describe the effective viscosity and diffusion coefficients taking into consideration
2.1 Formulation of governing equations 15

free surface S(x, y, t)


domain Ω 
z, upwards 

=
6
g
?
QQ
k
Q
*
 Q
 lateral boundary

 y,
> North
bottom B(x, y) (vertical)



 -x, East
O
Figure 2.1: The computational domain. z-axis directed upwards, parallel to –g.

the molecular effects and the fluid turbulence as well. D is the Navier-Stokes (de-
formation rate) tensor; p describes the global pressure. νD is a second order tensor
multiplication with a contraction.
The mathematical theory of the Navier-Stokes equations (2.1–2.2) is not complete, espe-
cially concerning the solution uniqueness, regularity and precise definition of boundary
conditions [30]. The non-linear character of the advection term u · ∇u produces and
maintains instabilities, for large Reynolds numbers Re exciting a wide scale of motions.
The equations (2.1–2.4) are formulated for a domain Ω limited by three general types
of boundaries: the bottom, the free surface and the lateral boundary. In the actual
model implementation, they must fulfil a few important constraints. Namely, B(x, y),
describing the bottom and S(x, y, t) representing the fluid free surface, are single-valued
functions with respect to the horizontal surface OXY (2.1). The lateral boundaries are
assumed to be vertical, so that they can be defined by a closed polygon L(x, y). The
vertical projection of the domain on the horizontal surface OXY Ω0 is clearly defined
and:
Ω = {(x, y, z) ∈ R3 , (x, y) ∈ Ω0 ∧ B(x, y) ≤ z ≤ S(x, y, t)} (2.6)
The lateral boundary can be a solid wall or a liquid, artificial open boundary introduced
in order to truncate the domain. These constraints allow a clear and simple definition
of the computational domain and the application of boundary conditions.

2.1.2 Incompressibility
One of the most widespread approximations in the geophysical flows is to treat water
as an incompressible fluid. This approximation is based on a physically well-supported
assumption that the water density is independent of pressure. The fluid density varia-
tions due to temperature, salinity and/or suspended matter are taken into account solely
through the equation of state (2.4). A more precise discussion of the non-trivial con-
ditions under which such a simplification is possible is given in the appendix, section
A.1. The mathematical consequences of this approximation are paramount. When the
16 2.1 Formulation of governing equations

fluid is incompressible, the velocity field describing its movement is non-divergent (or
solenoidal):
∇·u=0 (2.7)

Incompressibility implies that fluid volume rather than fluid mass is conserved. As a
further consequence, the hydrodynamic models based on this approximation assume
that the volume must be conserved while the density (and so the mass) changes freely
as a function of active tracers (temperature and salinity), but not pressure.

2.1.3 Turbulent stress tensor


The turbulent stress tensor is assumed to have a form analogous to the one proposed by
Boussinesq (eddy-viscosity principle) [12]. Physically speaking, −σij is the amount of
the i-th-component of the momentum flowing per unit time through a unit area normal
to the jth direction. The equation 2.5 is obtained as follows. For laminar flows:

σ = −pI + 2̺ν f D (2.8)

The Navier–Stokes tensor D is:


1  1
D= ∇u + (∇u)T − (∇ · u)I (2.9)
2 3

The viscosity tensor ν f is assumed to be diagonal. For incompressible flows (∇ · u = 0),


the components of the stress tensor σ are given by:

σij = −pδij + ̺νf j (∂ui /∂xj + ∂uj /∂xi ) (2.10)

When the Navier-Stokes equations are Reynolds-averaged, an additional term appears


in (2.10) due to turbulent fluctuations – the Reynolds stress tensor Rij = −̺u′ i u′ j . In
the Boussinesq eddy-viscosity concept the Reynolds tensor is approximated to:

Rij = −̺u′ i u′ j = +̺νtj (∂ui /∂xj + ∂uj /∂xi ) (2.11)

where ν t is the turbulent viscosity coefficient which varies in time and space and is
obtained from a turbulence model. Adding (2.11) to (2.10) yields:

σij = −pδij + ̺νj (∂ui /∂xj + ∂uj /∂xi ) (2.12)

where ν = ν f + ν t is the effective eddy-viscosity coefficient. For simplification, the


turbulent eddy viscosity tensor ν is assumed to be diagonal, with three components
νx , νy , νz . (The turbulent diffusion tensor for tracers assumes the same form.) It can
be shown by introducing the stress tensor in the form (2.12) to the Reynolds-averaged
incompressible Navier-Stokes equations (2.2) that for constant density a simplification
equivalent to setting ̺ν(∇u)T = 0 is possible. Eventually, the stress tensor assumes a
simple form:
σij = −pin δij + ̺νj (∂ui /∂xj ) (2.13)
2.1 Formulation of governing equations 17

The term describing stresses in the momentum conservation equation (2.2) can therefore
be written as follows:
1 1
∇ · σ = − ∇p + ∇ · (ν∇u) (2.14)
̺ ̺

2.1.4 Boussinesq approximation


This approximation is based on the fact that in natural water bodies such as rivers, lakes
and seas the variations of the density are relatively small. The variations of the relative
density in the oceans are less than ±3% of the average density ̺0 . The concept of this
approximation, formulated by Boussinesq [13, 145] is to take into account the variations
of density only in those terms of the momentum conservation equation, where they give
rise to buoyancy forces and to ignore them in those terms, where density appears as
a parameter describing fluid inertia. It means that the variations of fluid mass are
neglected, but their effect on fluid weight is retained.
In the momentum conservation equation:
∂u 1
+ u · ∇u + 2Ω × u = − ∇p + g + ∇ · (ν∇u) (2.15)
∂t ̺

the density is assumed to be a sum of a constant reference density (average density, ̺0 )


and a variable density variance ∆̺ in the form ̺ = ̺0 + ∆̺. The reciprocal of the
density in the pressure term −̺−1 ∇p can be linearised taking ∆̺/̺0 << 1 to:
!
1 1 1 1 ∆̺
= = ∆̺ ≈ 1− (2.16)
̺ ̺0 + ∆̺ ̺0 (1 + ̺0
) ̺0 ̺0

The pressure can also be written as the sum of the hydrostatic pressure pH resulting
from the average density ̺0 :
pH = ̺0 g(S − z) (2.17)

where S(x, y) is the position of the free surface and the pressure variation ∆p, p =
pH + ∆p. Therefore the pressure term in (2.15) can be transformed as following:
!
1 1 ∆̺ 1 ∆̺
− ∇p = − 1− ∇p = − ∇p + 2 ∇p = (2.18)
̺ ̺0 ̺0 ̺0 ̺0
1 ∆̺ ∆̺
− ∇p + 2 ∇pH + 2 ∇(∆p) =
̺0 ̺0 ̺0
1 ∆̺ ∆̺ ∆̺
− ∇p + g− g∇S + 2 ∇(∆p)
̺0 ̺0 ̺0 ̺0

The idea of the Boussinesq approximation is to neglect the terms of the second order,
namely the last two terms in (2.18):
1 1 ∆̺
− ∇p = − ∇p + g (2.19)
̺ ̺0 ̺0
18 2.1 Formulation of governing equations

The momentum conservation equation (2.2) transforms after applying the Boussinesq
approximation to the following form:
∂u 1 ̺
+ u · ∇u + 2Ω × u = − ∇p + g + ∇ · (ν∇u) (2.20)
∂t ̺0 ̺0

It should be strongly stressed that while ̺0 is a constant, average fluid density in a given
flow, ̺ is the variable, local density.

2.1.5 Non-inertial co-ordinate system fixed to the earth


The equation set (2.1–2.4) is formulated for a rotating Cartesian co-ordinate system fixed
locally to the earth’s surface, as described in section 2.1.1. For this system the earth’s
angular rotation velocity vector Ω at the geographic latitude ϕ is inclined towards the
y-direction (North). While the centripetal acceleration appearing in this system can
be treated as a small correction of the gravitational acceleration g, the Coriolis terms
cannot be neglected in general:
   
2wΩ cos ϕ − 2vΩ sin ϕ fH w − fV v
2Ω × u =  2uΩ sin ϕ  =  fV u (2.21)
   

2uΩ cos ϕ fH u

fH = 2Ω cos ϕ and fV = 2Ω sin ϕ are horizontal and vertical Coriolis coefficients. The
gravitational influence of other heavenly bodies (the moon, the sun) is neglected. A more
detailed discussion of the aspects connected with the non-inertial co-ordinate system is
provided in appendix section A.2.
For simulation of larger water bodies, such as seas and ocean parts using the rectangular
co-ordinate system, the earth’s curvature is neglected. For instance, the upper limit for
the model dimension using a Cartesian system is set in oceanography to one tenth of
the earth’s radius [92]. Additionally, simplifying assumptions concerning the latitudinal
variability of the Coriolis parameters can be made when required (appendix section A.2).

2.1.6 Equation set for geophysical, incompressible flows with


Boussinesq approximation
The hydrodynamic equation set, considered as the basis for all further non-hydrostatic
model developments consists of: (1) the three-dimensional equations of motion, with
all directions equally treated and Boussinesq approximation, (2) continuity equation
for incompressible fluids, (3) the transport equation for a tracer (temperature, salinity,
passive effluent concentration), (4) equation of state. It can be written as follows:
∂u 1 ∂p
+ u · ∇u = − − wfH + vfV + ∇ · (ν∇u) (2.22)
∂t ̺0 ∂x
∂v 1 ∂p
+ u · ∇v = − − ufV + ∇ · (ν∇v) (2.23)
∂t ̺0 ∂y
2.2 Hydrostatic approximation 19

∂w 1 ∂p ̺
+ u · ∇w = − − g + ufH + ∇ · (ν∇w) (2.24)
∂t ̺0 ∂z ̺0
∇·u=0 (2.25)
∂T
+ u · ∇T = ∇ · (ν T ∇T ) + qT (2.26)
∂t
̺ = ̺(T, s, c) (2.27)

In the equation of state (2.27) the density is a function of temperature T , salinity s or


suspended matter concentration c. An internationally acknowledged standard equation
of state is the UNESCO EOS 80 equation of state for seawater [160], appendix section
A.4. For the free surface flows, the equations (2.22–2.27) are associated with an equation
for the free surface. A turbulence model can be applied in order to obtain ν and ν T .

2.2 Hydrostatic approximation


The main aim of this work is to develop a model for free surface flows which avoids
the hydrostatic approximation. Therefore, this approximation and its consequences are
discussed in detail in the following section.

2.2.1 Scaling the vertical equation of motion


The various terms of the vertical equation of motion can be examined applying di-
mensional analysis in order to make rough estimates of their magnitude. The vertical
equation of motion (2.24) for constant and equal horizontal (νH ) and constant vertical
(νz ) turbulent viscosities can be written in full as:

dw ∂w ∂w ∂w ∂w 1 ∂p ̺
= +u +v +w =− − g + fH u (2.28)
dt ∂t ∂x ∂y ∂z ̺0 ∂z ̺0
2
∂ w ∂2w ∂2w
!
+νH + + νz
∂x2 ∂y 2 ∂z 2

Let us introduce the typical time, horizontal distance (length), depth and velocity com-
ponent scales T , L, H, U, W and define the following non-dimensional variables (with
asterisk) as:

x y z Ut u v w
x∗ = y∗ = z∗ = t∗ = u∗ = v∗ = w∗ = (2.29)
L L H L U U W

As the characteristic pressure ̺U 2 is taken:


p
p∗ = (2.30)
̺U 2
20 2.2 Hydrostatic approximation

Setting (2.29) and (2.30) into (2.28) and dividing the result by the factor appearing in
the pressure gradient term (U 2 /H) yields:
W 2 ∂w∗
!
HW ∂w∗ ∂w∗ ∂w∗ ∂p∗ H ̺ H
+ u∗ + v∗ + 2
w∗ =− − 2 g + fH u∗ (2.31)
L U ∂t∗ ∂x∗ ∂y∗ U ∂z∗ ∂z∗ U ̺0 U
2 2
W ∂ 2 w∗
!
HW ∂ w∗ ∂ w∗
+νH 2 2 + + νz
LU ∂x2∗ ∂y∗2 HU 2 ∂z∗2

Due to the continuity ∇ · u = 0 the following equation is approximately valid:


W U
= (2.32)
H L

Introducing this dependence into the viscous and advective terms in (2.31) one obtains:
W 2 ∂w∗ H 2 dw∗ W 2 dw∗
!
H W ∂w∗ ∂w∗ ∂w∗
+ u∗ + v∗ + 2 w∗ = 2 = 2 (2.33)
L U ∂t∗ ∂x∗ ∂y∗ U ∂z∗ L dt∗ U dt∗
HW ∂ 2 w∗ ∂ 2 w∗ νz ∂ 2 w∗
!
∂p∗ gH ̺ fH H
=− − 2 + u∗ + νH 2 2 + +
∂z∗ U ̺0 U LU ∂x2∗ ∂y∗2 LU ∂z∗2

Equation (2.33) is in the form, which allows formulation of conditions under which
various terms are negligible compared to the vertical gradient of global pressure p.

1. The vertical acceleration dw/dt can be neglected when the horizontal scale of mo-
tion dominates the vertical one, which can be interpreted as H << L or W << U.
Fortunately, in most of these geophysical flows, where hydrostatic approximation
can be applied (shallow water, long waves), both these conditions are usually ful-
filled. However, a very often overlooked consequence of the condition W << U
is that steep free surface and bottom slopes must be avoided, with a probably
acceptable upper limit of 1:10.
2. In the gravitational and buoyancy force term for the geophysical flows ̺/̺0 ≈ 1.
The value of the ratio (gH/U 2 ) varies greatly. If the motion to be described has
a character of a long wave, gH can be interpreted as squared celerity of the long
waves c2 . Setting for U 2 the amplitude of the velocity in the long wave from
equation (A.44) yields gH/U 2 = H/η0 , where η0 is the amplitude of the surface
wave. As a consequence this term cannot be neglected.
3. The Coriolis force can be neglected when the coefficient U/fH H, playing the role
of a Rossby number, is large. The horizontal Coriolis coefficient varies with the
latitude, fH = 2Ω cos φ, and has its greatest value on the equator, 1.45842×10−4 s−1
and is equal to zero at the poles. For larger horizontal velocity and shallow waters
the Coriolis term can be neglected.
4. The viscous terms can be neglected, when the coefficients L2 U 2 /νH HW and LU/νz ,
playing the role of Reynolds numbers, are large, which is the case for turbulent
flows with dominating horizontal motions.
2.2 Hydrostatic approximation 21

2.2.2 Hydrostatic approximation and its consequences


Scaling the vertical equation of motion (2.24) shows that when the conditions listed
above are fulfilled, the vertical acceleration component dw/dt, as well as Coriolis and
dissipative terms can be neglected. The equation (2.24) reduces then to the simple and
well-known hydrostatic equation:
∂p
= −̺g (2.34)
∂z
As a consequence, when the hydrostatic approximation is valid, the pressure in a free
surface flow can be obtained directly from (2.34). The set of hydrodynamic equations
is much simplified, because the pressure can be eliminated. When horizontal velocity
components are known from solving (2.22–2.23) the vertical velocity component can be
found directly by integrating the continuity equation (2.25).
The hydrostatic pressure is not appropriate in cases where the vertical accelerations are
important or the pressure field cannot be approximated by hydrostatic pressure, e.g.
when:

1. the vertical accelerations caused in deeper waters by influence of the steep bottom
slopes on the velocity field result in changing the flow direction (these influences are
damped in shallow water by gravitational influence of the free surface movements);

2. the description of vertical orbital movements of the fluid particles in the wave
problems is important (short waves);

3. the flows around structures are considered, where, e.g., the pressure field before
the obstacle in a flow cannot be assumed hydrostatic;

4. flows in fluids with variable density, as buoyancy-driven or stratified flows, thermal


convection, internal or lee waves are to be modelled.

2.2.3 Three-dimensional shallow water equations


When the hydrostatic approximation is valid, the global pressure can be obtained from
equation (2.34):
S S S ∆̺
Z Z Z
p= ̺gdz = (̺0 + ∆̺)gdz = ̺0 g(S − z) + ̺0 g dz (2.35)
z z z ̺0

where a variation of density from its average value ̺0 for a given flow ∆̺ = ̺ − ̺0 is
introduced. The gradients of this hydrostatic pressure can be set into the horizontal
momentum equations (2.22–2.23) yielding a sum of terms representing gradients of the
free surface and fluid density. For instance, in the x-direction:
"Z #
1 ∂p ∂S ∂ S ∆̺
− = −g −g dz (2.36)
̺0 ∂x ∂x ∂x z ̺0
22 2.3 Algorithm with the hydrostatic approximation

The resulting terms are often called barotropic (as independent of density and depth,
gradients of a 2D function S(x, y)) and baroclinic (as depth- and density-dependent,
gradients of a 3D function ∆̺(x, y, z)) parts of the hydrostatic pressure gradients.
As the consequence, the pressure can be found from the density field, influenced by the
concentration of active traces through the equation of state (2.4) and from the position of
the free surface. The fields of the active tracers are described by the transport equations
(2.3); for the free surface a separate (or derived) equation is required.
Additionally, for reasons discussed in section A.2, the terms containing horizontal Coriolis
parameters fH can be neglected for flows with dominating horizontal scales. Coriolis
terms with the vertical coefficient fV remain unchanged.
Finally, under incompressibility, Boussinesq and hydrostatic approximations, the follow-
ing set of three-dimensional shallow water equations can be formulated as follows:
"Z #
∂u ∂S ∂ S ∆̺
+ u · ∇u = −g −g dz + vfV + ∇ · (ν∇u) (2.37)
∂t ∂x ∂x z ̺0
"Z #
∂v ∂S ∂ S ∆̺
+ u · ∇v = −g −g dz − ufV + ∇ · (ν∇v) (2.38)
∂t ∂y ∂y z ̺0
∂u ∂v ∂w
+ + =0 (2.39)
∂x ∂y ∂z
∂T
+ u · ∇T = ∇ · (ν T ∇T ) + qT (2.40)
∂t
̺ = ̺(s, T, c) (2.41)

They are applicable for those shallow water flows, where some aspects of their vertical
structure, e.g. stratification or vertical profiles of the horizontal velocity have to be
resolved. The equivalency of shallow water and long-wave approximations are discussed
in appendix section A.3. For closure, they need an additional equation for the free surface
and a turbulence model yielding ν and ν T . The 3D shallow water equations (2.37–2.41)
are the governing equations of the hydrostatic code Telemac3D.

2.3 Algorithm with the hydrostatic approximation


This section describes briefly the algorithm of the code Telemac3D, developed in Lab-
oratoire National d’Hydraulique in Chatou, Electricité de France (EDF). The Telemac
modelling system was developed for the two-dimensional case by Galland, Goutal and
Hervouet [52] on the base of previous considerations of Benqué et al. [8]. The descrip-
tion of the three-dimensional, hydrostatic algorithm is given by Janin, Lepeintre and
Pechon [83]. As an example of the applications of this model LeNormant et al. [103] and
Malcherek [110] are to be mentioned. Leaving the numerical details to the chapter 3, let
us concentrate on the main ideas of this algorithm, outlined in figure 2.2. The treatment
of the free surface and pressure gradients in this model is characteristic for the techniques
based on hydrostatic approximation.
2.3 Algorithm with the hydrostatic approximation 23

2.3.1 Operator splitting


The algorithm is based on the operator-splitting method, described in more detail in
section 3.2. The leading idea is to split the differential operators appearing in equations
into parts with respect to their mathematical and/or physical properties and treat them
separately in consecutive fractional steps by applying the most appropriate numerical
methods. The global algorithm is explicit in time, and consists of the sum of partial
(intermediate) solutions for a variable f after advection f a , diffusion f d , and finally
after pressure – free surface – continuity f n+1 steps. The explicitly formulated time
derivative of the variable f is:

∂f f n+1 − f d f d − f a f a − f n
= + + (2.42)
∂t ∆t ∆t ∆t

f represents the horizontal velocity components u, v, or tracer concentrations T .

2.3.2 Advection and diffusion steps


In order to deal with the variability of the free surface, the σ-mesh with tiers of prismatic
elements is used, described in full detail in section 3.1.1. It allows tracking the time-
variable position of the free surface while formulating equations in the time-independent
σ-transformed mesh.
In the first step of the algorithm, the advection terms in the horizontal momentum
conservation and transport equations (2.22-2.26) are treated:
fa − fn
+ u · ∇f = 0 (2.43)
∆t

This hyperbolic equation is treated using the method of characteristics (i.e. a Lagrangian
method) in a separate step, or together with the diffusion step, basing on the streamline
upwind Petrov-Galerkin (SUPG) method (section 3.5).
In the diffusion (also parabolic) step the diffusive terms are treated:

fd − fa
= ∇ · (ν f ∇f ) + qf (2.44)
∆t

where qf represents the source terms. The diffusion step is solved in the σ-mesh by ap-
plying the 3D finite element method. In treatment of the time derivative the advantages
of the Crank-Nicolson formulation (semi-implicit) are used.
In this step of the algorithm the solution of the transport equations ends. In this version
of the algorithm the source terms qf for the velocity components are equal 0 for the
velocity components; for tracers they represented mass or heat sources. The next step
(pressure – free surface – continuity) influences the solution of the transport equation
only through the actualisation of the vertical node co-ordinates after finding the new
free surface position.
24 2.3 Algorithm with the hydrostatic approximation

2.3.3 Pressure – free surface – continuity step


2.3.3.1 Free surface and forcing

After the advection and diffusion steps, the following equations must be solved:
un+1 − ud
"Z #
∂S ∂ S ∆̺
= −g −g dz + fV v n (2.45)
∆t ∂x ∂x z ̺0
v n+1 − v d
"Z #
∂S ∂ S ∆̺
= −g −g dz − fV un (2.46)
∆t ∂y ∂y z ̺0
∂u ∂v ∂w
+ + =0 (2.47)
∂x ∂y ∂z
The treatment of this set of equation is based on the fact that the hydrostatic approx-
imation is equivalent to the long (shallow water) waves approximation (section A.3).
Long waves can be described by vertically integrated Navier-Stokes equations, i.e. St.-
Venant equations [92, 53]. This step of the algorithm takes full advantages of this fact.
However, all constraints resulting from the shallow water (hydrostatic) approximation
should be born in mind.
In the pressure – free surface – continuity step both barotropic and baroclinic parts of
the hydrostatic pressure gradient as well as the Coriolis terms must be treated. Finally,
the new free surface position has to be found. The equations to be solved in this step
are integrated vertically between the bottom B(x, y) and the free surface S(x, y, t) to
obtain Saint-Venant equations without diffusion and advection terms. They consist of
the conservative free surface equation which is the vertically integrated continuity equa-
tion (section 2.6.3.2) and of two momentum conservation equations for mean horizontal
velocity components:
un+1 − ud ∂h ∂B
+g = −g + Fx + fV v n (2.48)
∆t ∂x ∂x
v n+1 − vd ∂h ∂B
+g = −g + Fy − fV un
∆t ∂y ∂y
∂h ∂(uh) ∂(vh)
+ + = FH
∂t ∂x ∂y
where h(x, y, t) = S(x, y, t) − B(x, y) is the water depth. The mean velocity components
are computed as follows (for the numerical integration the trapezium rule is applied):
1 S
Z
ui = ui dz (2.49)
h B

The terms representing the buoyancy forces Fx and Fy are computed in an analogical
way:
1 S g S ∂ S ∆ρ
Z Z Z
Fx = Fx dz ′ = dzdz ′ (2.50)
h B h B ∂x z ρ0
1 S g S ∂ S ∆ρ
Z Z Z
Fy = Fy dz ′ = dzdz ′ (2.51)
h B h B ∂y z ρ0
2.3 Algorithm with the hydrostatic approximation 25

The source term FH is:


h
FH = [w∗ (z∗ ) − w∗ (0)] (2.52)
z∗
where w∗ is the σ-transformed vertical velocity component. It is equal to zero at the
surface and at the bottom due to the impermeability or kinematic boundary condition.
Non-zero values appear e.g. in the case of discharges at the bottom or surface.
In order to solve the equation system (2.48) the continuity equation is split into hy-
perbolic and parabolic parts, analogically to the global 3D-algorithm. The hyperbolic
part:
∂h ∂h ∂h
+u +v =0 (2.53)
∂t ∂x ∂y
can be solved separately using the 2D method of characteristics, or treated using stream-
line upwind Petrov-Galerkin formulation. The remaining equations:
∂h ∂u ∂v
+h +h = FH (2.54)
∂t ∂x ∂y
un+1 − ud ∂h ∂B
+g = −g + Fx + fV vn
∆t ∂x ∂x
v n+1 − v d ∂h ∂B
+g = −g + Fy − fV un
∆t ∂y ∂y

are solved simultaneously in a coupled way for u, v and h using the 2D finite element
method (on the basic mesh) with Crank-Nicolson (semi-implicit) treatment of the time
derivative. For this purpose parts of the two-dimensional algorithm of the Telemac2D
[52, 77] are applied.
The pressure – free surface – continuity step yields the new (i.e. for tn+1 ) position of
the free surface and new horizontal velocity components. The 3D horizontal velocity is
finally computed as follows:

un+1 = ud + un+1 3D − ud 3D − ∆tF u3D + ∆tFu (2.55)


v n+1 = v d + v n+1 3D − v d 3D − ∆tF v3D + ∆tFv

where Fu = Fx + fV v and Fv = Fy − fV u, i.e. three-dimensional source terms which


were vertically integrated in the Saint-Venant equations to obtain Fu and Fv (eq. 2.50).
Subscript 3D for a two-dimensional field means, that this field is three-dimensionalised
by taking a value for a given 2D-node (x, y) for each σ-mesh level.

2.3.3.2 Vertical velocity component

The vertical velocity is found through integrating the continuity equation taking the
impermeability condition at the bottom (section 2.7.3):
z ∂un+1 z ∂v n+1
Z Z
w n+1(z) = w n+1 (B) − dz − dz (2.56)
B ∂x B ∂y
26 2.3 Algorithm with the hydrostatic approximation

Telemac3D algorithm outline (hydrostatic)

Operator-splitting (f = u, v):

∂f f n+1 − f d f d − f a f a − f n
= + +
∂t ∆t ∆t ∆t
Advection of horizontal velocity (hiperbolic):

fa − fn
+ u · ∇f = 0
∆t
Diffusion of horizontal velocity (parabolic):

fd − fa
= ∇ · (ν∇f ) + qf
∆t
1 RS
Pressure and free surface (h = S − B, ui = h B ui dz):

un+1 − ud ∂h ∂B
+g = −g + Fu
∆t ∂x ∂x
v n+1 − v d ∂h ∂B
+g = −g + Fv
∆t ∂y ∂y
∂h ∂(uh) ∂(vh)
+ + =0
∂t ∂x ∂y
where the vertically integrated source terms are:

g
Z S ∂
Z S ∆ρ
Fu = dzdz ′ + fV v n
h B ∂x z ρ0

g
Z S ∂
Z S ∆ρ
Fv = dzdz ′ − fV un
h B ∂y z ρ0
3D horizontal velocity field:

un+1 = ud + un+1 3D − ud 3D − ∆tF u3D + ∆tFu

v n+1 = v d + v n+1 3D − v d 3D − ∆tF v 3D + ∆tFv


Continuity (vertical velocity component):
z ∂un+1 z ∂v n+1
Z Z
wn+1 (z) = wn+1 (B) − dz − dz
B ∂x B ∂y

Two-step tracer advection-diffusion (hyperbolic-parabolic):

Tn − Td Td − Ta
+ = ∇ · (ν T ∇T ) + qT − u · ∇T
∆t ∆t

Figure 2.2: Telemac3D algorithm outline (hydrostatic)


2.4 Numeric considerations concerning the solution method 27

∂B ∂B
w n+1(B) = un+1 + v n+1 (2.57)
∂x ∂y

Effectively, the vertical component w is adjusted to the horizontal velocity components u,


v so that the continuity equation is fulfilled and the resulting velocity field is divergence-
free. This subordinate treatment is again well-supported by the hydrostatic approxi-
mation, valid when the vertical velocity component is much smaller than the horizontal
ones.

2.4 Numeric considerations concerning the solution


method

2.4.1 General solution methods for 3D Navier-Stokes equations

From the mathematical point of view, the hydrodynamic and transport equations to
be solved are a set of non-linear equations of mixed parabolic-hyperbolic type. Three
methods of numerical treatment of these equations have found widespread acceptance.
Finite difference method (FDM) in the domain of free surface flows was first applied in
oceanography [126]; it was quickly followed by finite element method (FEM) [31] and
finite volume method (FVM) [123, 161], but mostly in engineering applications. Ad-
ditionally, spectral methods exist [47, 68]. All these methods discretise the differential
equation formulated for a continuous problem (with infinite number of degrees of free-
dom) into a discrete problem (with finite number of degrees of freedom) which can be
solved using a computer.
In FDM, the derivatives in the governing partial differential equations are converted
in the process of discretisation into finite differences. Various approaches of different
exactness are available. Consequently, the differential equations are converted into a
system of algebraic equations for nodal values of variables, which can be solved using a
suitable (linear or non-linear) equation solver in order to obtain an approximate solution.
FDM is usually based on structured meshes and is most efficient when the spacing is
equidistant and with quadrangle computational cells, leading to simple forms of algebraic
equations to be solved. Good results are obtained when staggered grids are applied, with
different nodes for velocity components and pressure, making it more similar to the finite
volume method. The disadvantage of this method is a complicated implementation
of boundary conditions in general-purpose codes, and problematic mass conservation
properties.
In FVM the discretisation method is based on balancing the fluxes between arbitrarily
chosen control volumes which are basic computational cells. Instead of formulating the
finite differences for separate nodes, it formulates balances of variable fluxes through
boundaries of finite volumes, operating on nodal values. FVM can be adapted to more
complex geometries in a relatively straightforward way. The resulting formulae and sets
of algebraic equations can be further treated as in FDM. Because this method is based
28 2.4 Numeric considerations concerning the solution method

directly on the flow properties, it is commonly treated as the most natural method for
fluid mechanics and shows much better conservative properties than FDM.
In FEM, the discretisation is achieved by assuming that the value of a variable can be
interpolated locally. The local approximation is substituted into a appropriately weighted
integral of the governing equation (weak formulation), and the integrals evaluated so that
a residue is minimised. This weighted residue method is the most appropriate one for the
hydrodynamic equations (of a mixed type) when solutions by finding a minimum of an
integral in an element (variational method) cannot be formulated in general. Similarly
to FDM, a set of algebraic equations is obtained and solved. While with FDM the
differential operators are approximated, in FEM the resulting function is discretised,
leaving the differential operators unchanged. FEM and FVM show their strength in
cases, where unstructured meshes are needed, allowing flexible adaptation to complex
domain geometries and variable resolutions. The implementation of boundary conditions
is also much easier. However, the mass conservation in FEM can be secured in a global
system only, while in FVM it is guaranteed at the local, control volume level.
In the spectral methods, the global procedure is similar to FEM, except that the assumed
solution for a variable is given in the form of a series of known functions of spatial
co-ordinates with coefficients to be determined. The approximation functions are not
defined locally, as in FEM, but for the whole domain, which puts severe constraints on
the domain geometry. Finally, a set of algebraic equations for the function coefficients
is obtained.
Combinations of various discretisation methods are also available. A very common ap-
proach is to use FEM for spatial, and FDM for time discretisation. This leads to schemes
of an enhanced stability, as e.g. Taylor-Galerkin-FEM [37, 38]. FEM and FVM can be
also coupled in order to obtain schemes, which are mass conservative not only in a global
system, as in FEM, but also locally, as in FVM, i.e. Control Volume FEM [70, 71].
Independently from the used general discretisation method, FDM, FVM or FEM, in
order to solve the equation set (2.22–2.27) four main groups of algorithms exist. They
are distinguished and classified by their overall approach to the numerical solution of the
governing equations.
First, a coupled equation system can be obtained by direct discretisation (using finite
difference or element method) of the governing equations in their primitive variable form
(i.e. the momentum conservation and continuity equation and an equation for the free
surface). The resulting global equation system is usually not easily solved using standard
methods. Non-linear, iterative equation solvers (e.g. Newton-Raphson) or predictor-
corrector methods with convergence being the greatest concern are used [49, 50, 80, 168].
In practice, only implicit formulation is successfully applied. This direct solution method
of Navier-Stokes equations is computationally very intensive, especially for the 3D case.

Second, decoupled methods for primitive variables exist, where momentum conservation
equations for three velocity components and a (modified) continuity and/or a derived
equation for pressure are solved in consecutive stages. The main idea of these methods is
to solve sequentially a number of smaller, linear equation systems instead of an iterative
2.4 Numeric considerations concerning the solution method 29

solution of a larger, usually non-linear and slowly converging one. The free surface
equation is also usually solved in a separate step. The decoupled methods differ mostly
in their algorithm structure, but also in the in the treatment of the time derivatives,
with an overall aim to avoid the solution of the fully implicit global equation system.
They appear in the literature under various names, as splitting schemes [52], fractional
step schemes [129] or projection methods [29, 57, 143]. Taylor-Galerkin-methods, where a
more precise approximation of time derivatives is used before applying the finite element
method, can be also assigned here [37, 38, 86]. Various pressure methods belong also to
the decoupled methods [18, 24]; they are described in detail in the following sections.
A special distinction between the two overall approaches – the direct, coupled one and
the fractional, decoupled one – lies in this particular case in the treatment of the in-
compressibility constraint. While in the direct method the continuity equation appears
explicitly as a constraint on the velocity field, in the decoupled pressure methods it is
replaced by a derived Poisson equation for pressure. Thus, the momentum and continu-
ity equations are decoupled, at least in the interior of the computational domain. The
coupling appears by boundary conditions for pressure [64, 137].
The third group of methods takes advantage of techniques for compressible flows:

1. The penalty function method, where the continuity equation is replaced by a form
stating that the pressure changes proportionally to the velocity divergence with a
proportionality parameter called bulk viscosity [28]. The solution of a thus modified
set of equations converges to the solution of the Navier–Stokes equations when the
bulk viscosity ν → ∞, so that the scheme is exact enough only for very large ν.

2. The artificial compressibility method [78, 148], where a limited compressibility of


the fluid is assumed and the continuity equation for compressible flows is applied.
The term representing the partial derivative of pressure in the continuity equation
disappears when steady state is achieved.

3. A combination of both [133].

The main disadvantage of these three methods is introduction of artificial, empirical


variables or constants and the fact that they are applicable mainly to stationary flows.
In the fourth general group of methods, Ψ − ω–methods, the stream function Ψ and
vorticity ω are introduced into basic equations, yielding a Poisson equation for Ψ and
a transport equation for ω in the two-dimensional case [47]. Using such non-primitive
variables brings two advantages: it reduces the number of equations to be solved and does
not require any pressure computation. A disadvantage is a complex form of the boundary
conditions. For the 3D case, vector potentials Ψ and ω are defined (u = ∇ × Ψ and
ω = ∇ × u). As a result, three Poisson equations for Ψ and three transport equations
for ω components are obtained [3]. The computational effort grows considerably in
comparison to other methods.
30 2.4 Numeric considerations concerning the solution method

2.4.2 Overview of the solution methods with FEM


While the mathematical theory for the solution of the Navier-Stokes equations with the
finite element method (FEM) leading to a non-linear global algebraic equation system
is well established [31, 129, 54], the theory for the decoupled methods is still matter for
investigation [59, 57, 62, 86]. The main reason for the introduction of the decoupled
methods is the computational efficiency of these codes compared to the cost of implicit,
iterative solution of very large non-linear equation systems for coupled velocity and
pressure in the direct methods. To obtain an algorithm, where this intense coupling
can be replaced by a sequential solution of smaller and linear equation systems, various
methods are applied. While the decoupled methods were successfully applied in the finite
difference method [18, 24], their formal application in FEM has been hampered in the
past by ambiguities caused by boundary conditions for the pressure [64, 137].
In order to approximately solve Navier-Stokes equations using the finite element method,
the following steps must be followed: (1) reformulate the given problem using the
weighted residue method (weak formulation); (2) perform discretisation in finite ele-
ments using finite-dimensional space of approximation functions, (3) solve the discrete
problem through solution of a set of algebraic equations for the nodal values and finally
(4) implement the method as a computer code. In following subsections the relevant
FEM theory is briefly sketched.

2.4.2.1 Weighted residue method

According to the weighted residue method, an approximation of an unknown function


u being a solution of a partial differential equation in the domain Ω(Γ) (with boundary
Γ):
L(u) = 0 (2.58)

is searched for. The function u is approximated in the form of a linear combination


of a finite number of base approximation functions ϕi being the functions of spatial
coordinates:
N
X
uh = ϕ i ci (2.59)
i=1

The base functions ϕi must be chosen in such a way that the boundary conditions at
Γ are satisfied for arbitrary real coefficients ci . When (2.59) is introduced to (2.58), a
residual value ε appears due to the fact that uh is only an approximation of u:
L(uh ) = ε 6= 0 (2.60)

In order to find an approximate solution of (2.58), this residue should be minimised in


Ω. This is obtained by setting the scalar product of the residue and chosen weighting
(test, trial) functions vi to zero, (vi , ε) = 0 (i.e. to assume that they are orthogonal).
This is equivalent to setting (see section 2.4.2.5):
Z
vj ε dΩ = 0 j = 1, ..., N (2.61)

2.4 Numeric considerations concerning the solution method 31

where N is equal to the number of coefficients ci . The functions vi allow different error
weighting in various parts of the domain Ω, and (2.61) must be evaluated only there,
where vj are not equal to zero. Consequently, the approximate solution uh of the following
integral equation is searched for:
Z Z N
X Z
vj ε dΩ = vj L(uh ) dΩ = ϕ i ci vj L(ϕi )dΩ = 0 (2.62)
Ω Ω i=1 Ω

In the general formulation of the weighted residue method nothing is said about the
interpretation of the base ϕi and test functions vi and the constants ci . According to the
way in which the weighting functions are chosen and the general method in which the
domain Ω is subdivided into a finite number of sub-domains (elements) different finite
element methods are achieved. In the standard methods, ci are equal to ui = uh (xi ), i.e.
function values at the defined finite element nodes and ϕi the interpolation (or shape)
functions corresponding to a node i. The Lagrangian interpolation functions ϕ, which
are applied in this work, are equal to 1 at the node i and to 0 at all other nodes. The
required order of the interpolation functions ϕ depends on the order m of the differential
operator L. In general, it is required that the (m − 1) order derivatives of ϕ must be
continuous.

2.4.2.2 Standard Galerkin finite element method

In the standard Galerkin finite element method (called also Bubnov-Galerkin or standard
FEM), which is applied e.g. in the Poisson equation or diffusion step of the presented
algorithm (elliptic, parabolic), the weighting functions are chosen to be equal to the ap-
proximation functions v = ϕ. However, in order to approximate the advection-dominated
problems properly, this method can be successfully applied only when the time derivative
is approximated with Taylor series of higher order. This method, called Taylor-Galerkin
FEM, has been developed by Donea et al. [37, 38].

2.4.2.3 Streamline upwind Petrov-Galerkin finite element method

In Petrov-Galerkin finite element method, the weighting and approximation functions are
different. This approach has been applied by Brooks and Hughes [16] in order to obtain
stable solutions with FEM for strongly advective problems (hiperbolic). In FDM, the
upwind (or upstream) differences are applied for discretisation of the advection equation
in order to obtain stable schemes – only nodal values from the incoming flow direction
are taken (and varying accordingly to the velocity vector direction). In the FEM, the
weighting functions can be modified in such a way that more information is obtained
from the upstream direction. This effect can be reached by application of appropriate
weighting functions of higher order, or, as in the streamline upwind Petrov-Galerkin
method (SUPG), by addition of polynomials of a lower order to the weighting function.
In the SUPG, the weighting function is taken as:
u
vi = ϕi + α ∇ϕi (2.63)
||u||
32 2.4 Numeric considerations concerning the solution method

where α is an upwind parameter which can be set to a constant, or a variable calibrated


according to a given problem [2] (see also section 3.6.2), u is the flow velocity. If linear
approximation functions are applied, ∇ϕi is constant, and vi are equal to the approxi-
mation functions modified by a constant. The velocity norm is usually computed as the
mean velocity in a given element. The SUPG method can be chosen in the advection
step of the described algorithm.

2.4.2.4 Finite elements

In order to find a discrete solution uh of (2.62) in the domain Ω it is necessary to subdi-


vide it into smaller sub-domains. In the process of subdivision of the bounded solution
domain Ω into non-overlapping finite elements (called triangulation or partitioning) it
is advisable to choose as simple geometric forms as possible in order to reduce compu-
tational cost. In 2D, usually triangles or quadrilaterals, in 3D prisms, tetrahedrals or
hexaedrals are selected. The number of nodes should allow application of the chosen
approximation functions, e.g. for polynomials of higher order the necessary number of
nodes (being support points for them) grows. As the nodal values the variables them-
selves (Lagrange nodes) or their derivatives (Hermite nodes) can be chosen. The integrals
resulting from the weak formulation are evaluated in the simplest, reference form of a
given element type. The real values from elements adjusted to the domain geometry are
transformed into the reference one to perform (simplified) computations. If the approx-
imation functions used for approximation of the spatial co-ordinates are identical with
them for the unknowns, the elements are called iso-parametric. A number of finite ele-
ments (which could be defined as a triplet (1) a geometric object, (2) a finite-dimensional
space of functions defined on it and (3) a set of nodes and values on them) are available
with various shape functions and different properties [31, 173]. The finite elements used
in this model are discussed in chapter 3.

2.4.2.5 Hilbert and Sobolev spaces

For the formal analysis of the finite element method a linear function space L2 (Ω) is
defined being a space of all functions f → R, for which Ω f 2 dΩ < 0. Together with a
R

scalar product defined as: Z


(f, g) = f g dΩ (2.64)

(compare with equation 2.62) this linear vector space builds a Hilbert space V , natural
for the problems considered here. The functions f are called square integrable functions.
The norm in the space L2 (Ω) is defined as:
q
||u|| = (u, u) (2.65)

For analysis of the approximate solutions to differential equations, Sobolev spaces H k (Ω)
are introduced. These Hilbert spaces consist of functions u all of whose partial derivatives
2.4 Numeric considerations concerning the solution method 33

up to the order k are square integrable:

H k (Ω) = {u ∈ L2 (Ω) : D α u ∈ L2 (Ω), |α| ≤ k} (2.66)

where the partial differentiation up to an order α is denoted as follows:


k
∂ |α| u
Dαu =
X
|α| = αk (2.67)
∂xα1 1 ...xαk k i=1

and α = (α1 , α2 , ..., αk ).


The scalar product in a Sobolev space is defined as:

(D α u, D α v)L2 (Ω)2
X
(u, v)H k (Ω) = (2.68)
|α|≤k

and the norm as:


q sX
||u||H k (Ω) = (u, u)H k (Ω) = ||(D αu)||L2 (Ω)2 (2.69)
|α|≤k

For example, the scalar product in H 1 (Ω) is:


Z
(u, v) H 1 (Ω) = (uv + ∇u · ∇v) dΩ (2.70)

From the mathematical point of view, the appropriate choice of the function space is
important, because it allows proving the existence of a solution, estimate errors, etc.
One speaks about the conform finite elements, when the interpolation functions are
included in the same Sobolev space in which the weak formulation of the problem is
stated. Therefore, the usual mathematical methodology for these elements is to tailor
an appropriate space in order to investigate the method properties [87].
In FEM, the infinite dimensional space of the solution function V is reduced to a finite
dimensional subspace Vh where (2.59) is valid, due to the limited number of the approx-
imation functions. The conform approximation methods are based on the assumptions,
that Vh ⊂ V and that the properties of V are valid also for Vh .

2.4.2.6 Weak formulation

The classical methodology of the more abstractly formulated FEM can best be illustrated
by a simple example. The following boundary value problem for the Poisson equation
(e.g. for pressure) in a bounded domain Ω(Γ) is to be considered:

−∇2 u = f in Ω (2.71)
u=0 at Γ (2.72)

As a classical solution one defines the solution of a partial differential equation of the
second order which belongs to C 2 (Ω)∩C 0 (Γ) by Dirichlet boundary condition (BC) or to
34 2.4 Numeric considerations concerning the solution method

C 2 (Ω) ∩ C 1 (Γ) by Neumann BC. (C k is the set of functions with continuous derivatives
up to order k).
Let us apply the weighted residue methodology introduced in section 2.4.2.1. By multi-
plying the Poisson equation by an arbitrary test function v ∈ V and integrating it in Ω
yields: Z Z
− v∇2 u dΩ = vf dΩ (2.73)
Ω Ω

Applying Green’s formula (n is the normal vector to the boundary):


Z Z Z
2
v∇ u dΩ = − ∇v · ∇u dΩ + v∇un dΓ (2.74)
Ω Ω Γ

to (2.73) and taking the boundary condition into consideration (the boundary integral
vanishes), the problem considered can be formulated as follows: find such a u ∈ V that:
a(u, v) = (f, v) ∀v ∈ V (2.75)

where a is a bilinear form and:


Z
a(u, v) = ∇v · ∇u dΩ (2.76)
ZΩ
(f, v) = f v dΩ (2.77)

V is a space of continuous functions in Ω with ∂v/∂xi piecewise continuous in Ω and


equal to zero at the boundary Γ. Such a description of V shows that the appropriate
Hilbert space for this problem is H01 (Ω), defined as H01 (Ω) = {v ∈ H 1 (Ω) : v=0 on Γ},
and the weak formulation (2.75) reads now – find u ∈ H01 (Ω) that:
a(u, v) = (f, v) ∀v ∈ H01 (Ω) (2.78)

The formulation (2.78) is named the weak formulation of (2.71) and its solution is named
weak solution of (2.71). If a weak solution u exists, it is not automatically the classical
solution of (2.71), because it must be regular enough that ∇2 u can also be defined in a
classical way.
The apparatus of functional analysis brings methods to find the solution of the weak
formulation, and then show that it is regular enough to be the classical solution of the
given problem. The usual way is to investigate the properties of the bilinear forms a(u, v)
and to use theorems stating the solution existence, continuity and stability.
The most useful for FEM considerations is that part of the theory which assigns linear
operators L : U → V ′ to bilinear forms on elements from Hilbert spaces U, V such that
a : U × V → R through:
(L(u), v) = a(u, v) v∈V (2.79)
V ′ = L(V, R) is a dual space to V , i.e. a set of all linear and continuous functions on V .
The norm of this space is constructed using an operator norm, for l ∈ V ′ :
( )
|l(u)|
||l||V ′ = sup (2.80)
u6=0 ||u||V
2.4 Numeric considerations concerning the solution method 35

For example, a dual space H −m (Ω) for a Sobolev space H m (Ω) has a following norm:
( )
(u, v)L2(Ω)
||u||−m = sup (2.81)
v6=0 ||u||m

The problems considered here can be constructed using scalar products and have the
following form: for a given f ∈ V ′ , a function u ∈ U is searched for, so that:

a(u, v) = (f, v) v∈V (2.82)

The weak formulation (2.82) is therefore equivalent to finding the inverse operator to L,
such that u = L−1 f . One of the most important theorems in the FEM theory states the
conditions under which L is isomorphic, i.e. both L and L−1 are continuous. L : U → V ′
is isomorphic when the assigned form a : U × V → R fulfils the following conditions:

• Continuity and boundedness (stability): for C ≥ 0:

|a(u, v)| ≥ C · ||u||U · ||v||V (2.83)

• Inf-sup-condition: for α > 0 and u ∈ U:


( )
a(u, v)
inf sup ≥α>0 (2.84)
u∈U v∈V ||U||U ||v||V

According to (2.80), (2.84) is equivalent to the statement for u ∈ U:

||L(u)||V ′ ≥ α||u||U (2.85)

The bilinear forms fulfilling (2.84) so that for v ∈ V |a(v, v)| ≥ α||v||2V are called
V -elliptic. For the weak formulation (2.82) one obtains from the inf-sup-condition (2.84)
or (2.85): ( ) ( )
a(u, v) (f, v)
α||u||U ≤ sup = sup = ||f ||V ′ (2.86)
v∈V ||v||V v∈V ||v||V

Therefore, for a bilinear and V -elliptic form a(u, v) the weak formulation (2.82) has a
solution u = L−1 f for which ||u||U ≤ α−1 ||f ||V ′ . For FEM, as U and V finite dimensional
spaces are taken.

2.4.2.7 Global equation system

In the finite element method, the solution is searched for in a subspace Vh of V , and the
aim is to find such uh ∈ Vh that:

a(uh , vh ) = (f, vh ) ∀vh ∈ Vh (2.87)


36 2.4 Numeric considerations concerning the solution method

In standard Galerkin FEM, the test functions are equal to the interpolation functions
and:
a(uh , ϕj ) = (f, ϕj ) j = 1..N (2.88)

where N is the number of nodes. Because uh is approximated using Lagrangian interpo-


lation functions so that:
N
X
uh = ϕi u i ui = uh (xi ) (2.89)
i=1

one may write the weak formulation in the form


N
X
ui a(ϕi , ϕj ) = (f, ϕj ) j = 1..N (2.90)
i=1

being a system of N equations with N unknowns being the nodal values ui :

Au = b (2.91)

where uj is the vector of the unknown nodal values and bj = (f, ϕj ) is the load vector.
The elements of the stiffness matrix aij = a(ϕi , ϕj ) are usually computed by summing
the contributions from M different elements:
M
X
a(ϕi , ϕj ) = am (ϕi , ϕj ) (2.92)
m=1

This matrix assembly can therefore be performed by computing element stiffness matrices
separately and then summing them up in a loop over all elements. It must be noted that
for Lagrangian interpolation functions am (ϕi , ϕj ) = 0 unless both nodes i and j are
nodes belonging to a given element m. The load vector b is computed analogically. The
solution of the equation system (2.91) yields the discrete solution of the given problem.
The solution of the linear equation system (2.91) can be performed using various meth-
ods, from (exact and direct) Gaussian elimination for smaller problems to (approximate
and iterative) minimising methods for a very large number of nodes, as, for example, with
the conjugate gradient method. The iterative methods usually require pre-conditioning
of the equation system matrices in order to reduce the number of iterations.
For computations using a vector computer, a special technique for treatment of the
resulting system matrix is advantageous to profit from the computational possibilities
of these machines. By using iterative equation solvers, the operation of multiplying a
matrix by a vector is very common and the most time-consuming computation part
compared to other operations. Therefore, a special storage system can be introduced in
order to make the matrix-vector product as effective as possible. For performing these
products, the assembly of the global matrix is not required. In the element-by-element
(EBE) method [81, 56, 73], the diagonal and extra-diagonal elements of the separate
element stiffness matrices are stored in memory and special procedures for performing
basic algebraic operations are provided. The main advantage is that the matrix-vector
2.4 Numeric considerations concerning the solution method 37

product is vectorisable and that the computationally intensive matrix assembly process
is avoided. It should be noted that this storage method does not yield any adventages
when using a scalar computer.
The vectorisation idea is to work simultaneously in real time on a cluster containing
a number of elements in a loop (e.g. vectorisation length of 64 or 128 on Cray YMP
or 1024 elements on Fujitsu-Siemens VPP). Therefore, a loop over all elements is per-
formed in chunks containing 1024 elements (for Fujitsu) significantly speeding up the
computation process. However, the method requires sorting of the element numbers in
a computational mesh after its generation in order to avoid vector dependencies. These
vector dependencies do not occur, when an element of a processed vector appears only
once in a cluster of 1024 elements. Namely, in a loop over element numbers in a given
vectorisation length the situation must be avoided that two or more elements share the
same node and this node has the same local numbers in the elements considered. The
conditions for the mesh sorting get more severe as the targeted vectorisation length
increases.
The numerical libraries of Telemac3D contain numerous relevant modern iterative
solvers and pre-conditioning procedures for large equation systems [73].

2.4.3 Equal and mixed interpolation

2.4.3.1 Stokes problem

Let us consider the stationary Stokes equations for an incompressible fluid with a con-
stant dynamic viscosity µ contained in a domain Ω(Γ) ⊂ R3 :

−µ∇2 u = −∇p + f in Ω (2.93)


∇ · u = 0 in Ω (2.94)
u = 0 on Γ (2.95)

where f are volume forces. The pressure is determined only up to a real constant C;
if a pair u, p is a solution of (2.93) then a pair u, p + C is a solution as well. In order
to obtain a unique solution for pressure an additional constraint should be posed, e.g.
R
Ω p dΩ = 0.

In order to obtain the weak formulation, (2.93) is multiplied by a test function v ∈ V =


[H01 (Ω)]3 which is divergence-free, ∇ · v = 0 in Ω, integrate it in Ω, and apply Green’s
formula:
Z Z Z
2
vf dΩ = −µ v∇ u dΩ + v∇p dΩ =
ΩZ Ω Z Ω Z Z
− v∇un dΓ + µ ∇v · ∇u dΩ + vpn dΓ − (∇ · v)p dΩ = (2.96)
ΓZ Ω Γ Ω

=µ ∇v · ∇u dΩ

38 2.4 Numeric considerations concerning the solution method

The pressure term disappears from the problem because v = 0 on Γ and the incompress-
ibility condition ∇ · v = 0 in Ω. The weak formulation can be stated as: find u ∈ V
that:
µ(∇u, ∇v) = (f, v) ∀v ∈ V (2.97)

Z
µ(∇u, ∇v) = µ ∇v · ∇u dΩ (2.98)
Z Ω

(f, v) = vf dΩ (2.99)

V = {v ∈ [H01 (Ω)]3 : ∇ · v = 0inΩ} (2.100)

It means that when a finite element method is to be formulated for (2.98), a finite-
dimensional space Vh of V , the incompressibility condition ∇ · v = 0 must be fulfilled
exactly. (It can easily be proved that this is the case, when v = ∇ × ψ, where ψ is
a stream function of the velocity field). It can be shown that under this condition the
(Galerkin) finite element method can yield the classical solution of (2.93). However,
constraints are present – the incompressibility condition must be fulfilled exactly [87].

2.4.3.2 Mixed interpolation

Mixed interpolation methods are applied in order to avoid the restriction to exactly
divergence-free velocity. The solutions for u and p are searched for in different spaces V
and Q, where V is not limited to divergence-free velocity. Formally, taking (2.96) into
consideration, the weak formulation of (2.93) is set now as follows: find (u, p) ∈ V × Q
that:

µ(∇u, ∇v) − (p, ∇ · v) = (f, v) ∀v ∈ V


(q, ∇ · u) = 0 ∀q ∈ Q (2.101)

Z
µ(∇u, ∇v) = µ ∇v · ∇u dΩ (2.102)

Z
(f, v) = v · f dΩ (2.103)

V = {v ∈ [H01 (Ω)]3 } (2.104)
Z
Q = {q ∈ [L2 (Ω)]3 : q dΩ = 0} (2.105)

(Note that (.,.) denotes a scalar product in the L2 space.). The finite element method, as
previously discussed, is obtained by replacing V and P by finite-dimensional subspaces
Vh and Qh . The mixed FEM can be formulated as follows: find (uh , ph ) ∈ Vh × Qh that:

µ(∇uh , ∇vh ) − (ph , ∇ · vh ) = (f, vh ) ∀vh ∈ Vh (2.106)

(qh , ∇ · uh ) = 0 ∀qh ∈ Qh (2.107)


2.4 Numeric considerations concerning the solution method 39

Introducing approximation functions ϕi for Vh (velocity) and ψi for Qh (pressure), the


problem (2.106–2.107) can be written in a compact matrix form:

Du − Cp = f (2.108)
CT u = 0

where D = (dij ), C = (cij ), F = (fj ) and:


Z
dij = a(ϕi , ϕj ), cij = (ψi , ∇ · φj ), fj = f ϕj dΩ (2.109)

In this mixed interpolation method the possibility is opened to work with velocities which
fulfil the incompressibility constraint only approximately, namely through the discrete
formulation (2.107) or the second equation of (2.108). The subspaces Vh and Qh should be
chosen in such a way that a compromise is achieved – the resulting method is stable and
accurate as well. For choosing these spaces, a condition analogous to inf-sup-condition
(2.84) is of paramount importance. The solution stability can be formulated as follows:
there exists such a constant C that when (uh , ph ) ∈ Vh ×Qh is a solution of (2.106–(2.107)
then
||uh||1 + ||ph || ≤ C||f ||−1 (2.110)

where the notation introduced by (2.81) is applied. It can be shown that in order to
fulfil (2.110) the following Ladyzhenskaya-Babuška-Brezzi condition (LBB-condition) for
(2.106–2.107) must be satisfied [15]: there exists such a constant c that:

(qh , ∇ · vh )
sup ≥ c||qh ||0 > 0 ∀qh ∈ Qh (2.111)
vh ∈Vh ||vh ||1

The LBB condition (2.111) provides a means to construct spaces for velocity Vh and
pressure Qh , so that the solution is stable. The most simple method is to choose Vh
to be large enough. Obviously, for computational reasons it is advisable to choose it as
small as possible, but in agreement with Qh .
The LBB condition supplied the appropriate mathematical tool for understanding and
removing the instabilities observed by practitioners when applying finite elements which
do not fulfil this condition [14]. Largest problems were caused by parasitic modes or
oscillations of the pressure [138, 164]. Various other methods to remove these instabilities
have been applied previously, as applying of numerical filters or averaging of resulting
values in a node patch, mass-lumping or setting un-physically high viscosity or diffusion
coefficients [149]. These procedures have a disadvantage in common that a part of the
physical information is simply damped in order to improve numerical stability.
As a rule of thumb, the interpolation order for the velocity should be higher than for
the pressure, however it is not enough to guarantee stability. For example, a quadrilat-
eral element with linear approximation for velocity and piecewise constant functions for
pressure proves to be unstable. Proofs of the LBB-condition for various element types
are not trivial, but already a number of element families has been theoretically checked
40 2.4 Numeric considerations concerning the solution method

[54, 15, 87, 14, 173]. Another method to avoid instabilities is to define the pressure on
a mesh which is coarser as the mesh for velocity [55].
Although the LBB condition has been formulated for elliptic problems, there is strong
evidence of its validity also for hyperbolic problems, Navier-Stokes equations and shallow
water equations [15, 54]. However, a strict mathematical proof of these hypotheses does
not exist at present [2]. Stabilised solutions for Stokes problem can be achieved using the
Petrov-Galerkin formulation [82]. Various other methods allow circumventing the LBB
condition (2.111) for more advanced problems, as for example using enhanced assumed
strain elements [164].
As an example of a 3D free surface model using the coupled, direct approach and where
an element strictly fulfilling the LBB condition is applied (a cube with quadratic approx-
imation functions for velocity, 27 nodes, and linear functions for pressure, 8 nodes) is
the work of Yost [168]. The resulting global equation system is solved using the Newton-
Raphson method. However, even in this case instabilities of the free surface have been
observed. Huerta and Liu [80] in their model for 2D viscous flows with large move-
ments of free surface, use a predictor-corrector method for solving the resulting equation
system. Another example is a model developed by Frederiksen and Watts [50].

2.4.3.3 Equal-order interpolation

For the decoupled methods, where the solution of the Navier-Stokes equations is achieved
in separately performed steps and where a set of smaller equation systems are solved
instead of a global one (section 2.4.1) a consistent FEM theory still does not exist. In
the past, most researchers used elements where the pressure is approximated by functions
of lower order than the velocity ones, as Gresho et al. [63, 57, 60] or Donea et al. [39], and
succeeded even with elements which do not fulfil LBB conditions strictly. Larger efforts
have been invested in better understanding of imposing various boundary conditions
(especially pressure) [64, 137] or in stabilisation of the numerical integration in time
[149, 150, 86, 89].
In search for a better approximation of pressure gradients in distorted elements, and in
order to simplify the algorithms and improve the solution efficiency, equal order approx-
imation for both pressure and velocity (bilinear) has been successfully (re)introduced by
Gresho et al. [61, 62], although the equal-order elements also fail the LBB-test. Further
investigation shows that when a weak formulation of a derived pressure Poisson equation
(with automatically built-in boundary conditions) is solved in a separate step (sections
2.5, 3.3.1), stable and accurate solutions for incompressible flows are obtained [89, 66].
As an explanation why an equal-order approximation is appropriate, it is presumed,
that the LBB condition is circumvented by the fact that the decoupled methods do not
require an incompressibility condition in the form of an explicit equation ∇ · u = 0 oc-
curring in the global equation set. The incompressibility is asymptotically achieved by
the convergence of a pressure equation solution. As mentioned before, stability can be
additionally enhanced by applying proper boundary conditions, improving the methods
of integration in time and appropriate treatment of advection terms (SUPG).
2.5 Pressure treatment 41

For free surface flows, efficient equal-order finite element 2D models (in vertical plane,
aimed for wave applications) have been developed by Ramaswamy and Kawahara [130,
132] and Jiang and Kawahara [86], both containing a solution of a pressure equation in a
separate step. A 3D model with non-equal approximation based on a projection method
for the global pressure [57] has been developed by Chen [26].

2.4.4 Method applied in the developed model


In the non-hydrostatic model described in this work, a decoupled method based on
the previously existing operator-splitting approach is implemented. Linear interpolation
functions in space are applied for all flow variables: for the velocity and the pressure
in a prismatic 3D-element and for the water level in a triangular 2D-element (section
3.1.3). Therefore, this algorithm uses the equal-order interpolation functions in finite
elements. Concerning integration in time, the overall algorithm is explicit. For the
advection-diffusion and free surface steps semi-implicit or fully implicit approaches in
the framework of the overall operator-splitting technique are possible (section 3.2). In
the following sections the details of the applied method are provided, starting with an
overview of the pressure treatment and pressure equations.

2.5 Pressure treatment


The characteristic feature of the implemented solution algorithm of the three-dimensional
set of hydrodynamic equations without the hydrostatic approximation (2.22–2.27) is
treatment of the pressure. Application of the hydrostatic approximation yields an explicit
equation for the pressure (section 2.2) from the third (vertical) momentum conservation
equation. For an arbitrary, three dimensional case without the hydrostatic approxima-
tion it is not possible.
For compressible flows, the pressure appears in the continuity equation (2.1) because
the fluid density is pressure-dependent. For the incompressible flows, in the continuity
equation (2.25) only velocity components appear, and there exists no explicit equation
for pressure. An equation for the pressure (or a potential function related to pressure)
can be derived from the momentum conservation equations; in this case one speaks about
a derived pressure equation.
In the following paragraphs, the decoupled pressure methods for primitive variables using
a derived pressure equation are shortly reviewed.

2.5.1 Deriving of the conventional pressure Poisson equation


Applying the divergence operator on the Navier-Stokes equations in the following general
form:
∂u
+ u · ∇u = −∇P + ν∇2 u + f (2.112)
∂t
42 2.5 Pressure treatment

(where P = p/̺ is the kinematic pressure and f represents other force terms) one obtains:
!
∂u
∇· + u · ∇u = −∇2 P + ν∇ · (∇2 u + f) (2.113)
∂t

Assuming that divergence and ∂/∂t can be commuted and taking ∇ · u = 0, one obtains
the general pressure Poisson equation:

∇2 P = ∇ · (ν∇2 u − u · ∇u) + ∇ · f (2.114)

As a consequence of the incompressibility approximation, the global pressure p can be


found as a solution of an elliptic partial differential equation. Each change of the pressure
propagates at infinite speed in order to keep the flow incompressible everywhere.
The boundary conditions for this separate equation for global pressure should be obtained
from the momentum and continuity equations. The Neumann boundary conditions for
this equation derived from the momentum equations (2.112) are:
!
2 ∂u
n · ∇P = n · ν∇ u − u · ∇u − +f =
∂t
∂un
ν∇2 un − − u · ∇un + fn (2.115)
∂t

The Dirichlet boundary condition can be obtained from the dynamic boundary condition
(2.178), i.e.:
∂un
P =ν − Fn (2.116)
∂n
The numeric solution of the thus stated Poisson–Neumann (2.114–2.115) or Poisson–
Dirichlet (2.114–2.116) problem is not trivial, because in general it contains unknown
velocities on the right hand side of the equation. The boundary conditions cannot be
implemented in a straightforward way and are a source of numerous numerically caused
ambiguities [64, 58].

2.5.2 General idea of the projection method


The large family of projection methods is based on the fact that any vector field v in a
domain Ω with a smooth boundary δΩ can be uniquely decomposed into the following
form [30]:
v = vd + ∇φ (2.117)

where vd is solenoidal (divergence-free) and parallel to the boundary δΩ:

∇ · vd = 0 in Ω
vd · n = 0 at δΩ (2.118)
2.5 Pressure treatment 43

and φ is a scalar function. ∇φ is therefore rotation-free and (2.117) is equivalent to


splitting the vector v into divergence-free and rotation-free components. The momentum
conservation and continuity equations can be written in the following way:
∂u
+ ∇P = S(u) in Ω (2.119)
∂t

where on the left side a sum of the local acceleration (velocity time partial derivative)
and of the kinematic pressure gradient is present. The function of u on the right side
appears as:
S(u) = g + ν∇2 u − u · ∇u (2.120)

Generally, S(u) is neither divergence-free nor rotation-free. Due to the incompressibility


of the flow and from the vector analysis it can be remarked that:
!
∂u ∂
∇· = (∇ · u) = 0
∂t ∂t
∇ × ∇P = 0 (2.121)

Following the method described by Chorin [30], when u is given, then S(u) is also
known and therefore can be projected simultaneously into subspaces of the divergence-
free vectors ∂u/∂t and rotation-free vectors ∇P . This can be formally stated as:
∂u
= P[S(u)] ∇P = Q[S(u)] (2.122)
∂t
Operators defined this way, P and Q = I − P are called projection operators. P projects
any vector v onto the null space of the divergence operator – ∇ · P(v) = 0, and Q
projects any vector onto the null space of the rotation operator ∇ × Q(u) = 0. They
fulfil the following dependencies: P 2 = P, Q2 = Q, PQ = QP = 0. Comparison of
(2.122) and (2.119) brings the following form of these operators [57]:

P = I − ∇(∇2 )−1 ∇ · Q = I − P = ∇(∇2 )−1 ∇· (2.123)

Using these operators, the pressure and local acceleration can be decoupled from the
Navier-Stokes equations.
As the starting point, the numerical solution procedure takes the following approximate
formulation:
∂ũ
= S(u) − ∇P̃ (2.124)
∂t
where P̃ is a guessed or approximate pressure (from initial conditions, previous step).
The resulting intermediate velocity ũ is not divergence-free, because P̃ 6= P in general.
ũ starts from the same initial condition as the final, real velocity u and can be projected
into the divergence-free subspace using operator P. Using this intermediate velocity
given by the first step, a divergence-free solution can be approximated in the next stage
by the projection:
ud = P ũ (2.125)
44 2.5 Pressure treatment

It is assumed that ud is the final solution. However, the operator P is difficult to apply
directly, because the operator ∇2 can be inverted formally only by Green function. In
order to perform this inversion one must have information about the particular boundary
conditions and domain geometry. Therefore, another approximation is met – recalling
the equation (2.117) the projection step can be realised by the following decomposition:

ũ = ud + ∇φ ∇ · ud = 0 (2.126)

where φ plays the role of a Lagrangian multiplier associated with the projection of the
intermediate solution ũ into the space of divergence-free vectors. The main idea of the
projection algorithm can be therefore sketched as follows:

1. Intermediate velocity step. Taking the divergence-free velocity from the initial
condition u0 as well as an approximate initial pressure P0 , solve (2.124):

∂ ũ
= −∇P n + g + ν∇2 un − un · ∇un (2.127)
∂t
where un = u0 and P n = P0 at t = 0.

2. Projection step. Perform the projection as follows: solve for φ and ud from (2.126):

ũ = ud + ∇φ ∇ · ud = 0 (2.128)

in a two-step procedure:

• Solve the pressure Poisson equation (PPE) achieved by applying the diver-
gence operator on the first equation of 2.126:

∇2 φ = ∇ · ũ (2.129)

• Determine the projected velocity, i.e. the final one:

ud = ũ − ∇φ (2.130)

3. Updating. Accepting the projected velocity as the final result un+1 = ud find the
physical pressure P n+1 (e.g. by solving (2.114) – but it can be avoided as discussed
further) and repeat the cycle.

A number of algorithms based on the projection method, especially for internal flows,
have been formulated since the first papers of Chorin [29] and Patankar [124, 123] (SIM-
PLE algorithm). In the description above, the non-trivial problems connected with the
boundary conditions to be applied at different stages of this general algorithm, espe-
cially for the PPE were not mentioned. An extensive reviews of different variations of
this method, various forms of PPE’s and, above all, boundary conditions are given by
Gresho and Sani [64, 57, 58].
2.5 Pressure treatment 45

2.5.2.1 Projection-1 method

In the original method developed by Chorin [29] the intermediate velocity ũ is computed
from momentum equations without taking the pressure gradients into consideration.
From this intermediate result, the global pressure and the divergence-free final velocity
are found in the projection step. The function φ is in this case the physical global
pressure. The realisation of this idea in a numerical scheme is described in section 2.5.3.

2.5.2.2 Projection-2 method

In the second version of the projection algorithm, the influence of the pressure gradients
is not neglected in the solution of the momentum equations. The difference between
the intermediate velocity ũ and the final, divergence-free one un+1 is therefore smaller.
In the intermediate velocity step the pressure gradient from the previous step ∇P n
is applied. In the projection step, a pressure correction φ and the final velocity are
found. φ can be treated as a more abstract potential function related to pressure. In the
original SIMPLE/SIMPLER algorithms [124, 123], in order to deal with the problems
associated with the non-linearity of the momentum equations, an iterative procedure is
implemented. The whole algorithm is repeated until the divergence of the intermediate
velocity is approximately equal to zero. The pressure can be updated by solving the
conventional pressure Poisson equation (2.114) from the known velocity field.

2.5.2.3 Projection-3 method

Gresho [57] proposed a projection-3 method as a possible improvement of the projection-


1 and projection-2 methods. In this scheme the rate of convergence is accelerated by
taking into consideration the rate of change of the pressure gradients ∂∇P/∂t. A strict
mathematical consideration presented by Shen [143] shows that the projection-3 method
is not a proper approximation of the unsteady Navier-Stokes equations.

2.5.3 Pressure equation from fractional step formulation


The pressure Poisson equation used in the numerical realisation of the projection-1
method can be obtained from the time-discretised form of the Navier-Stokes equations.
In the solution algorithm the velocity time derivative is treated explicitly and can be
split into
∂u un+1 − ũ ũ − un
= + (2.131)
∂t ∆t ∆t
where ũ is an intermediate solution for the velocity field, which does not need to satisfy
the incompressibility condition. In this way the equation set (2.22–2.24) can be trans-
formed into a set of equations containing all terms but pressure gradients and a second
set containing them exclusively:
ũ − un
+ u · ∇u = −wfH + vfV + ∇ · (ν∇u) (2.132)
∆t
46 2.5 Pressure treatment

ṽ − v n
+ u · ∇v = −ufV + ∇ · (ν∇v) (2.133)
∆t
w̃ − w n ̺
+ u · ∇w = − g + ufH + ∇ · (ν∇w) (2.134)
∆t ̺0

un+1 − ũ 1 ∂p v n+1 − ṽ 1 ∂p w n+1 − w̃ 1 ∂p


=− =− =− (2.135)
∆t ̺0 ∂x ∆t ̺0 ∂y ∆t ̺0 ∂z

Combining equations (2.135) and taking into consideration that the resulting field un+1
must fulfil the incompressible continuity equation (2.25):

∇ · un+1 = 0 (2.136)

another form of the Poisson equation for the pressure can be obtained:

̺0
∇2 p = ∇ · ũ (2.137)
∆t

which is known as pressure Poisson equation from fractional step formulation. Find-
ing the pressure from (2.137) and substituting it into equations (2.135) one obtains a
divergence-free velocity field, i.e. satisfying the continuity equation.
The appropriate boundary conditions for the different steps of this fractional algorithm
are theoretically discussed in [64, 57, 58, 129]. According to Quartapelle [129] in the first
part of the algorithm the Dirichlet and Neumann boundary conditions for the physical
velocity should be applied. In the second part, the impermeability conditions must be
fulfilled by the resulting field. The Neumann boundary condition in the form n · ∇p = 0
can be derived for this equation. Gresho et al. [64, 57, 58] discuss in detail consequences
of applying various boundary conditions (and their simplifications), obtaining numerous
variations of the projection algorithm.
The advantage of the algorithm using equation (2.137) in comparison to the pressure
Poisson equation (2.114) is based on the fact that the velocity ũ on the right hand of
the equation (2.137) is now known from the first step of the algorithm.
In the projection-2 method, the known pressure gradients from the previous step are
taken into consideration when solving equations (2.132–2.134) and the pressure Poisson
equation is formulated for a pressure correction. This scheme can be sketched as follows:

ũ − un 1 ̺
+ u · ∇u = − ∇pn + g + ∇ · (ν∇u) − 2Ω × u
∆t ̺0 ̺0
n+1
u − ũ 1
= − ∇(pn+1 − pn ) (2.138)
∆t ̺0
n+1
∇·u =0
2.5 Pressure treatment 47

2.5.4 Poisson equation for the hydrodynamic pressure in free


surface flows

In contrast to the internal flows, for a free surface flow the pressure term in the momen-
tum equations −̺−10 ∇p can be separated into terms consisting: (1) that part of pressure,
which can be explicitly computed, and (2) that part, which can be found by solving a
Poisson pressure equation (2.137).
In the free surface flows, the hydrostatic pressure can be easily found. Therefore, in
following, the pressure is decomposed (split) into a sum of the pressures p = pH + π,
where the sum components can be physically interpreted as the hydrostatic pressure pH
and hydrodynamic (dynamic, motion) pressure π.
This fact can be used to develop an algorithm, where the equations (2.22–2.25) will be
split in such a way, that the hydrodynamic pressure is treated as a correction to the
hydrostatic component. A velocity field obtained from the gradients of the pressure
correction field (i.e., the hydrodynamic pressure field) overlapped on the velocity field
obtained using hydrostatic pressure, makes the resulting velocity field solenoidal, i.e.
incompressible. The derivation follows the pattern described in the previous section:

ũ − un 1 ∂pH
+ u · ∇u = − − wfH + vfV + ∇ · (ν∇u) (2.139)
∆t ̺0 ∂x

ṽ − v n 1 ∂pH
+ u · ∇v = − − ufV + ∇ · (ν∇v) (2.140)
∆t ̺0 ∂y
w̃ − w n 1 ∂pH ̺
+ u · ∇w = − − g + ufH + ∇ · (ν∇w) (2.141)
∆t ̺0 ∂z ̺0
̺0
∇2 π = ∇ · ũ (2.142)
∆t
un+1 − ũ 1 ∂π v n+1 − ṽ 1 ∂π w n+1 − w̃ 1 ∂π
=− =− =− (2.143)
∆t ̺0 ∂x ∆t ̺0 ∂y ∆t ̺0 ∂z

This algorithm is analogous to the large projection algorithm family with this significant
difference, that the split pressure components can be physically interpreted. In order to
obtain the intermediate velocity field, the explicitly given hydrostatic pressure gradients
are applied. In the projection step, the gradients of the hydrodynamic pressure obtained
from a Poisson equation are used. Due to the fact that the free surface, as well as the
tracer concentration field determining the fluid density are computed in a decoupled way,
the hydrostatic pressure gradients are given at a time level n.
For those rare cases, when the hydrodynamic pressure in comparison to the hydrostatic
one is large in the whole computational domain, a version analogous to the projection-2
algorithm can be considered. In this case, in the momentum equations (2.139–2.141)
not only the hydrostatic pressure gradients are applied, but also the gradients of the
hydrodynamic pressure from the previous step are taken into consideration. In the second
48 2.5 Pressure treatment

part (projection), a pressure correction φ (treated as a correction to the hydrodynamic


part of the pressure, φ = π n+1 − π n ) and the final velocity field is computed:

ũ − un 1 1 ̺
+ u · ∇u = − ∇pH n − ∇π n + g + ∇ · (ν∇u) − 2Ω × u
∆t ̺0 ̺0 ̺0
un+1 − ũ 1
= − ∇φ (2.144)
∆t ̺0
∇ · un+1 = 0

2.5.5 Treatment of pressure and buoyancy terms


The main idea of the algorithm is to separate the hydrostatic pressure pH from the
hydrodynamic pressure π and to treat them in different parts of the algorithm, whereby
the hydrodynamic pressure is treated as a pressure correction to the hydrostatic one.
The terms to be discussed (from eq. 2.22–2.24) are:

1 ∂p 1 ∂pH 1 ∂π
Πx = − =− −
̺0 ∂x ̺0 ∂x ̺0 ∂x
1 ∂p 1 ∂pH 1 ∂π
Πy = − =− − (2.145)
̺0 ∂y ̺0 ∂y ̺0 ∂y
1 ∂p ̺ 1 ∂pH 1 ∂π ̺
Πz = − − g=− − − g
̺0 ∂z ̺0 ̺0 ∂z ̺0 ∂z ̺0

There are two ways of separating (splitting) the pressures, discussed in following two
sections. For simplicity, atmospheric pressure pa is set to zero and the hydrostatic
pressure is assumed to be equal to zero at the free surface.

2.5.5.1 Hydrostatic pressure resulting from a constant mean density

The hydrostatic pressure can be defined by the following hydrostatic equation:

∂pH
= −̺0 g (2.146)
∂z

i.e. the hydrostatic pressure resulting from the constant mean fluid density for a given
flow ̺0 (a parameter resulting from the Boussinesq approximation).
This formulation is similar to the small perturbation equations. Integration in the water
column yields:
pH = −̺0 g(S − z) (2.147)

pH this is the pressure, which would result if in equations (2.22–2.24) the hydrostatic
approximation has been performed for a fluid with a constant density ̺0 . This is not the
physical hydrostatic pressure, but a part of it resulting from splitting the equations.
2.5 Pressure treatment 49

The pressure and buoyancy terms are (for Πy analogically):


" #
1 ∂p 1 ∂ ZS 1 ∂π ∂S 1 ∂π
Πx = − =− ̺0 gdz − = −g − (2.148)
̺0 ∂x ̺0 ∂x z ̺0 ∂x ∂x ̺0 ∂x

The vertical term is (introducing ∆̺ = ̺ − ̺0 ):


1 ∂p ̺ 1 1 ∂π ̺ ∆̺ 1 ∂π
Πz = − − g = − (−̺0 g) − − g=− g− (2.149)
̺0 ∂z ̺0 ̺0 ̺0 ∂z ̺0 ̺0 ̺0 ∂z

As a result, in the horizontal terms free surface gradients appear independent of the
density (barotropic part) and gradients of the hydrodynamic pressure. The terms de-
scribing horizontal gradients of the density disappear. In the vertical term the buoyancy
term representing density effects remains with the vertical gradient of the hydrodynamic
pressure.

2.5.5.2 Hydrostatic pressure resulting from the local density

In this case the hydrostatic pressure is resulting from the local fluid density ̺(x, y, z) i.e.
in the usual physical way, and defined by equation:
∂pH
= −̺g (2.150)
∂z
Integrating in the water column yields:
S S S ∆̺
Z Z Z
pH = ̺gdz = (̺0 + ∆̺)gdz = ̺0 g(S − z) + ̺0 g dz (2.151)
z z z ̺0

pH is now the physical hydrostatic pressure with taking the variable fluid density ̺ into
consideration. ∆̺ = ̺ − ̺0 is a density variation from an average value for a given flow.
The pressure and buoyancy terms are (for Πy analogically):
" #
1 ∂p 1 ∂ 1 ∂ S ∆̺ 1 ∂π
Z
Πx = − =− [̺0 g(S − z)] − ̺0 g dz − = (2.152)
̺0 ∂x ̺0 ∂x ̺0 ∂x z ̺0 ̺0 ∂x
" #
∂S ∂ Z S ∆̺ 1 ∂π
−g −g dz −
∂x ∂x z ̺0 ̺0 ∂x

The vertical term is:


1 ∂p ̺ ̺ 1 ∂π ̺ 1 ∂π
Πz = − − g= g− − g=− (2.153)
̺0 ∂z ̺0 ̺0 ̺0 ∂z ̺0 ̺0 ∂z

As a result, free surface gradients independent of the density (barotropic part), horizontal
gradients of the pressure resulting from density differences (baroclinic part) and gradients
of the hydrodynamic pressure appear in the horizontal terms. In the vertical term
only the vertical gradient of the hydrodynamic pressure remains. This is an important
simplification and this treatment of the pressure terms is eventually implemented in the
non-hydrostatic algorithm.
50 2.6 Free surface

2.6 Free surface


A free surface is a material surface across which the density is discontinuous. Calculating
the position of the free surface in a hydrodynamic-numeric model provides a unique chal-
lenge. First of all, free surface and interfacial flows can be difficult to model numerically
because the computational domain changes in time. The free surface must be discretely
described (free surface representation) and advanced in time (free surface tracking or ad-
vancing). The application of diverse important boundary conditions at the free surface
(surface tension, external pressure fields, wind stress, thermal radiation and boundary
conditions for variables describing turbulence) is troublesome, because they must be ap-
plied at the eventual location of the free surface. This is difficult, especially when fluid
regions may be coalescing, detaching or breaking up. All these problems are closely
related – the algorithms used to advance or track the free surface depend on its discrete
description or representation, and the boundary conditions can be applied only when
the free surface position is known. Finally, the barotropic pressure gradients caused by
free surface slopes are a driving force of the flow, especially in shallow waters, making
the exact reproduction of free surface movements a crucial point in the modelling.
A few methods of finding free surface position found wider acceptance. Two of them,
marker-in-cell (MAC) and volume-of-fluid (VOF), and various mixed methods belong to
schemes where the fluid regions or volume is tracked (figure 2.3). In the height function
and line segment methods, the free surface itself is modelled. Although only the height
function method has actually been used here, all these methods are described here in
order to deliver a critical overview of the presently available methods in light of the finite
element method.
The position of the free surface is determined by the following equation:

dxs
= us (2.154)
dt

where xs describes a free surface point and us velocity at the free surface. The same
equation applies for the fluid interior, and when there are no specific boundary conditions
to be considered at the surface, its position can be found directly from tracking the fluid
volume.

2.6.1 Marker-and-cell method

The Marker-and-cell method (MAC), developed 1965 by Harlow and Welch [65], uses
massless marker particles distributed in the whole fluid volume to trace the free surface
movements. The free surface is therefore tracked using a Lagrangian approach. In the
original technique the computational grid is fixed in time and space. The method was
improved later in several ways, e.g. by Miyata [115], Tomé and McKee [156, 157] in the
framework of the finite difference method, and Nakayama and Mori in finite elements
[117].
2.6 Free surface 51

In this method the free surface is tracked by positions of fictitious and massless La-
grangian marker particles. Initially, they are equally distributed in the whole fluid vol-
ume (in order to distinguish the computational cells taken by the fluid) and are advected
passively with the local fluid velocity. The fixed computational mesh covers the whole
area of possible fluid movement, so that some cells remain empty. The instantaneous
fluid configuration can be determined by finding the marker positions in subsequent time
steps. Cells which have at least one empty neighbour cell are marked as boundary cells.
In order to find the free surface position properly, a possibly uniform marker distribution
in each time step and in each fluid cell is necessary (for instance, at least four markers for
square 2D cells [65]). The marker distribution varies due to velocity gradients, inflows
and outflows, and it is necessary to redistribute the particles evenly in all fluid cells
when the they tend to spread disproportionately. In an optional step the position of the
free surface inside the boundary cells can be determined from the marker distribution in
them [115].
A great advantage of this method is the ability to handle complicated, general and arbi-
trary free surface problems: breaking surfaces [115]; dam breaking or splash of a falling
column of water [65]; fluid detachment or coalescence (droplets); or to simulate filling
complex moulding shapes [157]. This method is, however, computationally prohibitive
for large-scale, three-dimensional applications, because of the need to use a large number
of continually redistributed particles in order to find the surface shape properly from their
positions. Even with a large number of marker particles, it is difficult to obtain sufficient
information for determining the interface orientation in a cell. Because the number of
particles is finite, false regions of void cells can be generated in regions with greater
velocity shear (gradients). It is problematic to impose boundary conditions governing at
the free surface, especially for pressure. Additionally, there are also severe restrictions
relating the time step to the free surface wave speed, so that the computational grid
must have sufficiently fine resolution. Finally, no three-dimensional applications of this
method are known in the literature of the subject.

2.6.2 Volume of fluid method


Volume of fluid method (VOF) was introduced by Hirt and Nichols in 1981 [79]. It can be
applied to problems with a number of fluids with different densities and can also handle
breaking surfaces. A VOF algorithm consists generally of three parts: (1) an algorithm to
track the volume and locate the free surface, (2) a scheme to track the surface as a sharp
interface between fluids (e.g. water/air) moving through a given computational grid, and
(3) means of applying boundary conditions at the surface. The main idea is to introduce
a fractional volume (colour, filling fraction, fill state) function F (x, t) ∈ [0, 1] indicating
the fraction of a mesh cell that is filled with a fluid of a particular type, with value exactly
1 in cells occupied by fluid and exactly 0 in void cells elsewhere. A fractional value in
a cell indicates that this cell contains the free surface. For incompressible flow F can
be regarded as a normalised density. As such, it must satisfy the following conservation
equation:
dF ∂F
= + ∇ · (uF ) = 0 (2.155)
dt ∂t
52 2.6 Free surface

real configuration

r r r r r r
rr r r r r r r r r Marker-in-Cell (MAC)
rr r r r r r r r r r
rr r r r r r r r r r r
rr r r r r r r r r r r
rr r r r r r r r r r r r
rr r r r r r r r r r r r
rr r r r r r r r r r r r
rr r r r r r r r r r r r
rr r r r r r r r r r r rr
rr r r r r r r r r r r rr r
rr r r r r r r r r r r rr rr rr r

Volume of Fluid (VOF)

Figure 2.3: Volume-tracking methods: MAC and VOF


2.6 Free surface 53

i.e. F is being advected in the velocity field u but not diffused. The volume tracking
algorithm solves the advection equation for the function F in such a way that the interface
is kept sharp. The interface position is recovered from F after solution. Applying the
explicit integration scheme to equation (2.155) the following is obtained:

∆t X n
F n+1 = F n + Q (2.156)
Vi k ik

where the superscripts denote the time steps, and Qnik represents the flow rate into
element/cell i of a volume Vi through face k at the time-step n. In the computation of
Qnik geometric factors must be taken into consideration:
Z
Qnik = −αk qik
n
= −αk unj nj dΩ (2.157)
k

where αk represents the wet fraction of the cell face k, i.e. through which the fluid
n
flows. The fractional volume flux qik through the face k depends on the velocity field
and direction of the unit normal vector n.
There exists a number of different techniques based on purely geometrical considerations
n
to determine αk or qik directly according to the fill state of a given cell and its neigh-
bours. It is equivalent to reconstructing the interface position from the eventual fluid
configuration after each time step. Rudman [135] reviews and discusses advantages and
disadvantages of the existing algorithms (Hirt-Nichols’ donor-acceptor [79], SLIC [121],
FCT-VOF [170, 135] and Youngs’ method [169, 135]) in numerous test cases.
Because the equation (2.156) does not guarantee that F remains bounded, i.e. F ∈ [0, 1],
an additional flux-limiting (bookkeeping adjustment) algorithm must be applied in order
to avoid over-filled or over-emptied elements. Depending on the fill states of the neigh-
bouring cells and the values of F , the geometric factors αk are adjusted in such a way
that F has the value exactly 1 in entirely filled cells and exactly 0 in empty ones.
Applying boundary conditions with in the VOF method can be associated with difficul-
ties, especially when dealing with a number of fluids with very different densities and
viscosities and taking the surface tension into account.
Examples of interesting free surface applications of this method include dam-breaking,
Rayleigh-Taylor instability [79], wave breaking [102], gravity currents, collisions of
droplets, impacting and splashing of liquid drops at a free surface, bubbles rising and
bursting at a free surface [135], merging and fragmentation in multiphase flows [94].
In contrary to the Lagrangian MAC method, the VOF method represents an Eulerian
approach. The advantage of VOF over the MAC method is in the storage requirements:
only one word is required for one cell, compared to positions of a few marker particles.
However, this method does not conserve mass and momentum exactly, as stated in [152,
153]. The computational costs of obtaining an accurate interface shape from geometrical
considerations can be high.
54 2.6 Free surface

2.6.3 Height function methods

Other three-dimensional modelling techniques are based on treating the free surface
directly as a moving boundary. The height function method requires that the free surface
can be described in this formulation by a single-value function (height function) S(x, t)
with respect to one of the co-ordinate directions. In this method the distance between
the interface and a given reference level is calculated from a separate equation. This
approach offers a simple and robust method of simulating free surface flows. However,
the restriction to single valued functions exclude a wide class of flows, e.g. breaking
surfaces, bubbles, drops. In contrast to the MAC and VOF methods, which can be
described as volume-tracking methods, the height function methods are surface-tracking.
The position of the advancing surface can be tracked using an Eulerian approach (in a
fixed grid) or a Lagrangian one, where the mesh can be adapted to the position of the free
surface at each time step (e.g. σ-mesh). The most widely used equations for tracking the
free surface are the kinematic boundary condition and conservative free surface equation.

Compared to the MAC and VOF methods the numerical implementation of the height
function method is relatively simple. Therefore, for these flows, where the free surface
can be described by a single-valued function, this method is to be preferred. This is
the case in most fluvial and marine applications. An additional and very important
advantage is that various boundary conditions at the free surface can be implemented
without greater problems.

2.6.3.1 Kinematic boundary condition at the free surface

When the free surface can be represented by the equation F (x, y, z, t) = S(x, y, t)−z = 0,
F will always be 0 for a particle at this surface and the rate at which the function F
varies is given by its material derivative:

dF ∂S ∂S ∂S
= + us + vs − ws = 0 (2.158)
dt ∂t ∂x ∂y

The above equation is the well-known kinematic boundary condition at the free surface.
The suffix s indicates the velocity components at the free surface. Equation (2.158)
is a partial differential equation of a hyperbolic type. By requiring the continuity of
motion, it can be shown that the surface satisfying (2.158) consists always of the same
particles [96]. S(x, y, t) must be a single-valued function, so very steep surface gradients
or wave-breaking effects cannot be reproduced.
Although the evolution equation (2.158) is traditionally called a boundary condition, it
can be used to determine and advance the position of the free surface, when all three
components of the velocity at the surface are known. When the surface position is given
for some reasons, the kinematic boundary condition can be used as a usual boundary
condition, e.g., for the vertical component of the velocity.
2.6 Free surface 55

As an example of successful finite difference implementation, the models presented by


Bulgarelli, Casulli and Greenspan [18] should be mentioned. Finite element implemen-
tations using kinematic boundary condition, as [90, 165], show the main limitations of
this method, in the fact that the surface cannot break and the limited range of Reynolds
numbers, for which the numerical solution remains stable. Similar problems are encoun-
tered by Yost [168] when testing the fully implicit finite element model mentioned above.
Farmer et al. [43] use the FV method for computing the flow and the FD method for
the kinematic boundary condition.

2.6.3.2 A conservative form of the free surface equation

Integration of the continuity equation (2.7) over the depth, i.e. from the bottom z =
−B(x, y) to the free surface z = S(x, y, t), with usage of the Leibniz theorem1 in order
to exchange the differentiation and integration sequence, yields:
∂u Z S
∂v Z S
Z
∂wS
dz + dz + dz =
−B ∂x −B ∂y −B ∂z
∂ S ∂ ∂S ∂ S ∂ ∂S
Z Z
udz + ub (−B) − us + vdz + vb (−B) − vs + ws − wb = 0
∂x −B ∂x ∂x ∂y −B ∂y ∂y

The suffixes b and s indicate values on the bottom and surface, respectively. Using the
impermeability condition at the bottom (2.185) and kinematic boundary condition at
the free surface (2.158) one obtains the following conservative form of the free surface
equation:
∂S ∂ ZS ∂ ZS
+ udz + vdz = 0 (2.159)
∂t ∂x −B ∂y −B
This equation can be used to determine the free surface position instead of the kinematic
boundary condition (2.158). The main advantage of this equation is that it includes the
kinematic boundary condition and the impermeability condition at the bottom as well.
This approach brings a method of finding a free surface while automatically satisfying
the mass conservation criterion. It is applied in the pressure – free surface – continuity
step of the hydrostatic Telemac3D algorithm. Casulli and Stelling [24] apply it in their
3D-non-hydrostatic code in an implicit formulation.

2.6.4 Line segment method


The line segment method is a generalisation of the height-function method and found its
applications in 2D-models. The free surface (or any other interface) is defined as a chain
of short line segments (shorter than mesh spacing) defined by an ordered set of points
1
Leibniz formula
Z b(x,y) Z b(x,y)
∂ ∂f (x, y, z) ∂b(x, y) ∂a(x, y)
f (x, y, z)dz = dz + f (x, y, b) − f (x, y, a)
∂x a(x,y) a(x,y) ∂x ∂x ∂x
56 2.6 Free surface

whose co-ordinates are moved according to the fluid velocity. Therefore, this method is
not limited to single value surfaces. The position of the advancing surface points chain
can be tracked using Eulerian and/or Lagrangian methods [144]. The most important
difficulty appears when the chains intersect, fold over themselves or break up (merging
or break up of fluid regions) – a procedure detecting such a situation and reordering the
chain(s) is needed. Most applications of this method are found in the physicochemical
simulations of two phase flows, crystal growth, forming of ice, melting, etc.

2.6.5 Combinations of different methods

Thomas et al. in search of mass- and energy-conserving free-surface computation devel-


oped a method based on a combination of the volume of fluid method with the kinematic
boundary condition (FD) [153]. The method ensures a good mass conservation during
long simulation times and smooth surface movements (wave breaking excluded). Lim-
itations of this model are due to restrictions on the free surface movements ensuring
numerical stability.
An interesting treatment of the free surface in a fully conservative 3D FV code was
developed by Lilek [105]. After solving the momentum equations using the specified
pressure at the current free surface (the dynamic boundary conditions, see section 2.7.2),
a pressure-correction scheme with zero pressure correction at the free surface is applied
in order to enforce mass conservation. The prescribed pressure at the surface produces
a velocity correction there, resulting in a non-zero mass flux. The new position of the
free surface is determined by moving the boundaries of surface cells in such a way, that
their movement compensates the flux obtained in the previous stage. These two steps
are repeated iteratively until no compensation through movement of the free surface is
needed and the obtained free surface position fulfils the kinematic boundary condition.

2.6.6 Treating variable domain extents

The movements of the free surface create a need to update the computational domain
extents and, when needed, redefine node positions, boundary conditions, etc.
For dealing with the variable domain extents, Eulerian approaches exist, where the free
surface movement is described in a fixed mesh (mostly applied in FDM, e.g. [18], in FEM
e.g. together with the MAC method [117]). Alternatively, with the Lagrangian approach
the mesh moves with the flow (e.g. [131]). Both methods lead to different formulations
of Navier-Stokes equations, whereby with the Lagrangian method the terms describing
advection disappear.
Lagrangian methods may lead to very distorted elements negatively influencing the ac-
curacy. On the other hand, with the Eulerian approach difficulties in tracking the free
surface and applying boundary conditions at the free surface exist. Therefore, a com-
bined approach, known under a global name of Arbitrary Lagrangian-Eulerian (ALE)
methods, has found great popularity in FEM applications, e.g. [36, 130, 80]. Globally
2.7 Initial and boundary conditions for the flow 57

speaking, in order to avoid larger element distortions, the ALE-mesh moves with an ar-
bitrary velocity which is different from the local fluid velocity. When the mesh velocity
is equal to 0, an Eulerian approach is obtained, when it is equal to the fluid velocity, the
Lagrangian viewpoint is assumed. The equations are formulated and solved in a trans-
formed co-ordinate system, and at each time step (or when the mesh becomes distorted
enough) mesh re-zoning is performed in order to adapt it to the new domain geometry.
For free surface flows it is very convenient to assign the arbitrary mesh velocity to the
movements of the free surface [80, 26].
Tezduyar et al. apply a deforming space-time finite element formulation. In this case,
the finite element formulation is written in the associated time-space domain; the space-
time finite element mesh covers the space-time domain of the problem. In this way, the
mesh movement is taken into account automatically [149]. In general, this approach
also allows an arbitrary definition of mesh movement, like ALE. Tezduyar, Behr, et al.
successfully apply this method to free surface and interfacial flows [150, 151].
In the model described in this work, a well-known approach for the mesh re-zoning and
adaptation to the movement of the free surface known as σ-transformation is assumed
(described in detail in section 3.1). In principle, this method can be treated as a
simplified ALE method applied to free surface geophysical flows, where mesh nodes
can move only vertically with distances between the nodes relative to the water depth
remaining constant in time. Telemac3D also provides a method to deal with tidal flats
or flooding areas, which are not treated in this work [74].

2.6.7 Implemented free surface algorithms


Due to the σ–mesh structure of Telemac3D, requiring the free surface to be a single-
valued function, the most natural methods of finding the free surface are those based
on the height function. Eventually, two methods based on the kinematic boundary
condition and one based on the conservative free surface equation are implemented.
Implementation of MAC or VOF methods can be considered, but without changing
the mesh structure no special benefits can be expected due to the fact, that their most
important advantages concern the ability to describe arbitrary free surface configurations.

2.7 Initial and boundary conditions for the flow


In this section the initial and boundary conditions are described. In order to set a well
posed problem, initial and boundary data must be provided so that a unique solution of
the governing partial differential equations exists and depends continuously on the data.
The computational domain Ω for the geophysical flows with the free surface is bounded
vertically between the bottom and the changeable free surface, and laterally by imperme-
able (solid walls, bottom) as well as open (inflow, outflow) boundaries. They represent
physical boundaries associated with real limits of the domain as well as artificial bound-
aries due to the fact that the computational area can be truncated for practical reasons.
58 2.7 Initial and boundary conditions for the flow

The extents of these boundaries may change due to the movements of the free surface.
While at the physical boundaries the boundary conditions (BC) can be obtained from
the flow physics, the artificial boundaries require care in providing information so that
a unique solution can be obtained.
In the theory of the incompressible flows, on the domain boundary Γ one or more velocity
components w can be prescribed (imposed as a Dirichlet boundary condition), so that
[64]:
u(x, t) = w(x, t) x∈Γ t≥0 (2.160)
Z Z
∇ · wdΩ = w · ndΓ = 0 t≥0 (2.161)
Ω Γ

The equation (2.161) results from application of the divergence theorem to the continuity
equation, which must be fulfilled by w. Neumann boundary conditions for velocity on
Γ are obtained from the normal or tangential stress components:

σn = (σ · n) · n (2.162)
σt = (σ · n) · t (2.163)

As it will be shown in the next sections, most of the Neumann BC at the boundaries for
a flow variable f can be formulated in the following form:
∂f
k = n · ∇f = af + b (2.164)
∂n
where k, a, b are constants or functions independent of f , what simplifies the implemen-
tation of these BC in numerical schemes (chapter 3).
Both types of boundary condition are described in detail in following sections, preceded
by a discussion of properly set initial conditions.

2.7.1 Initial conditions


As an initial condition (IC) in the computational domain interior as well as on the
boundaries all the independent variables such as density, velocity, free surface position,
temperature and tracer concentrations must be given. Additionally, for incompressible
flows, it is of the greatest importance that the initial velocity field u0 must be solenoidal
and fulfils initial boundary conditions as well:

u(x, t) = u0 (x, 0) x∈Ω∪Γ t=0 (2.165)


∇ · u0 = 0 x∈Ω t=0 (2.166)
u0 · n = w(x, 0) · n x∈Γ t=0 (2.167)

In practice, very often a null flow field is convenient. However, it is also convenient to
initialise the flow at the velocity values equal to or interpolated from the inflow/outflow
boundary conditions. In some cases it is also important to define the pressure field at
t = 0.
2.7 Initial and boundary conditions for the flow 59

However, there are cases in which an initialisation of the model starting from a given
divergent initial velocity U0 may be dictated by practice. In this case, a projection onto
the space of divergence-free velocities (solenoidal) can be performed, treating the field
U0 given as the initial one as an intermediate velocity (see section 2.5.3). First, the
following Poisson equation is solved:

∇2 φ = ∇ · U0 (2.168)

an then, the projection is performed:

u0 = U0 − ∇φ (2.169)

obtaining a divergence-free initial field u0 . This field must fulfil all the initial boundary
conditions. As the initial pressure the hydrostatic pressure resulting from the initial free
surface configuration and density field is quite appropriate. (If the projection-2 method
is chosen, the initial hydrodynamic pressure can be set to zero in most cases.)

2.7.2 Dynamic boundary conditions

The dynamic boundary conditions are defined as the BC obtained from the stress con-
tinuity at the boundaries of the computational domain. The boundary is oriented by a
normal vector. For a boundary described by an equation F (x, y, z) = 0:

∇F
n= (2.170)
|∇F |

As a convention in fluid mechanics, the normal vector points out of the computational
domain. As the kinematic boundary condition represents the continuity of the velocity at
the free surface, the dynamic boundary conditions are the expression of the requirement
of the stress vector σ · n continuity across the free surface. However, the dynamic
boundary conditions can be applied at all kinds of boundaries. In the normal direction
to the boundary:
σ · n = (−pin + 2µD) · n = (−pout I + τ ) · n (2.171)

pin and pout are the pressures inside and outside the domain boundary oriented by the
normal vector n (and two tangential vectors t1 and t2 ). τ the outside boundary stress
tensor (e.g. wind or bottom stresses). The dynamic viscosity µj = ̺νj , where νj are
the components of the variable eddy viscosity. σ is the turbulent stress tensor and
D is the Navier-Stokes tensor, given by (2.9). For Reynolds-averaged incompressible
Navier-Stokes equations for turbulent flows (2.2), with introduction of the Boussinesq
eddy-viscosity concept [12] (note section 2.1.3) the turbulent stress tensor reads:

σij = −pin δij + µj (∂uj /∂xi ) (2.172)


60 2.7 Initial and boundary conditions for the flow

From conditions (2.171) one normal and two tangential conditions can be obtained by
projection of these three conditions in the direction of the normal or a tangential vec-
tor. This is a convenient way of treating these conditions in domains with complicated
boundary geometry. The projection is realised as follows:

σn = (σ · n) · n (2.173)
σt = (σ · n) · t (2.174)

The projections in normal and tangential directions of (2.172) yield:

∂unx ∂unz ∂unz


σn = −pin + µx + µy + µz = −pout + τnout (2.175)
∂x ∂y ∂z

∂utx ∂utz ∂utz


σt = µx + µy + µz = τtout (2.176)
∂x ∂y ∂z

where fnout and ftout are the normal and tangential outside boundary stress force (traction
force) components. Using the sloppy, but short and convenient notation ∂f /∂n = n · ∇f
the conditions (2.175 – 2.176) can be written in a compact way:

∂un ∂ut
−pin + µ = −pout + τnout µ = τtout (2.177)
∂n ∂n

For simple geometries it may be convenient to let normal and tangential directions co-
incide with the co-ordinate system axes. For the free surface given by the equation
z = S(x, y, t) from (2.171) one obtains:

∂u τxz ps ∂S ∂u ∂S ∂u ∂S
νz = − + νx + νy
∂z ̺ ̺ ∂x ∂x ∂x ∂y ∂y
∂v τyz ps ∂S ∂v ∂S ∂v ∂S
νz = − + νx + νy (2.178)
∂z ̺ ̺ ∂y ∂x ∂x ∂y ∂y
∂w τzz ps pa ∂w ∂S ∂w ∂S
νz = + − + νx + νy
∂z ̺ ̺ ̺ ∂x ∂x ∂y ∂y

where the outside stress components multiplied by surface gradients are already ne-
glected, which is a good approximation in the case of wind stresses. ps and pa are
the pressure at the surface and the atmospheric pressure, respectively. From equations
(2.178) simplified conditions can be obtained for these applications, where the gradients
of the free surface are small enough that they can be neglected and n ≈ (0, 0, 1):

∂u τxz ∂v τyz ∂w
νz = νz = ps = pa − τzz + ̺νz (2.179)
∂z ̺ ∂z ̺ ∂z

For the bottom, similar conditions can be obtained. At the lateral boundaries n =
(nx , ny , 0) and a two-dimensional form of (2.175) can be applied. For example, in a
2.7 Initial and boundary conditions for the flow 61

simple case when the lateral boundary is represented by n = (nx , 0, 0), one obtains the
following conditions:
∂u ∂v ∂w
pin = ̺νx + τxx νx = τxy νx = τxy (2.180)
∂x ∂x ∂x

The dynamic boundary conditions are usually applied in order to take the wind stress on
the surface and friction on the impermeable boundaries into consideration. They provide
a Neumann boundary condition for the velocity parallel to the boundary.
For inviscid flows, the dynamic boundary conditions reduce to the equilibrium of pres-
sures at both sides of the boundary:

pin = pout (2.181)

In the case of the free surface, (2.181) is called the inviscid free surface normal stress
boundary condition and is a very good approximation of the dynamic boundary conditions
also for small movements of the free surface for flows with low vertical viscosity νz . In
most of the simple application cases presented in this work it proves to be exact enough.
The dynamic boundary conditions also yield conditions for a pressure equation derived
from Navier-Stokes equations (section 2.7.11).

2.7.3 Impermeability: normal velocity at solid boundaries


Impermeability conditions at the bottom and (impermeable) lateral boundaries repre-
sented by a normal vector n = (nx , ny , nz ) for the velocity at the boundary ub read:

ub · n = 0 (2.182)

This condition expresses the fact that no mass flux is allowed through the impenetrable
boundary and therefore the normal velocity must vanish there. The impermeability
condition can be implemented in different ways depending on the numerical method
applied. The first method is to correct the resulting velocity field (which generally does
not fulfil the impermeability condition exactly) at the boundary ub by setting the normal
velocity at this boundary to null:

un = (ub · n) · n = 0 (2.183)

Effectively, from each resulting velocity component u, v, w the appropriate x, y and


z-components of the resulting normal velocity un are subtracted, so that the imperme-
ability condition is eventually fulfilled:

ub = u − un2x − vnx ny − wnx nz


vb = v − unx ny − vn2y − wny nz (2.184)
wb = w − unx nz − vny nz − wn2z
62 2.7 Initial and boundary conditions for the flow

Another method is to obtain a condition in the form of an equation analogical to the


kinematical boundary condition (2.158). E.g., for the bottom, described by the equation
F (x, y, z) = B(x, y) + z the direction of the normal vector is given by (∂B/∂x, ∂B/∂y, 1)
and from (2.182) one obtains:

∂B ∂B
ub + vb + wb = 0 (2.185)
∂x ∂y

This equation can be used e.g. to impose such a vertical velocity component which
fulfils the impermeability condition by given horizontal velocity components and bottom
gradients. In this way one component (w) plays a subordinate role to the two others (u,
v).

2.7.4 Tangential velocity at solid boundaries

The dynamic boundary conditions (2.177) yield Neumann boundary conditions for the
component of the velocity tangent to a solid boundary. If they cannot be applied for
some reason, they must be replaced by a Dirichlet boundary condition for the tangential
velocity (slip condition):
ut = u0 (2.186)

This boundary condition is usually connected with zero Neumann boundary condition
for the normal velocity component. u0 can be chosen to be equal 0 (no-slip condition) or
given a non-zero value. By setting the tangential velocity to zero it should be remembered
that the computational mesh must be fine enough to resolve the velocity profile between
the boundary and free stream velocity.

2.7.5 Rigid lid

The rigid lid boundary condition at the free surface is an example of a very simplified
treatment of the free surface. This boundary condition descends from the well-known
rigid lid approximation in oceanography [53]. This approximation is widely used for
investigation of flows in separated barotropic and baroclinic modes. It takes advantage
of the fact that for the baroclinic mode the displacements of the free surface are small
compared with the movements of internal interfaces. In consequence, the vertical com-
ponent of the velocity at the surface is set to zero and for the horizontal components zero
Neumann boundary conditions are applied. Obviously, at the free surface (usually as-
sumed to be flat) pressure gradients can be given representing gradients or movements of
the free surface (or atmospheric pressure gradients) for the depth-independent barotropic
mode. In practice, for purely density-driven flows with Boussinesq approximation (small
density variations), a horizontal free surface is appropriate.
2.7 Initial and boundary conditions for the flow 63

2.7.6 Bottom and free surface stresses


Bottom and free surface stress (shear) is taken into consideration by applying appropriate
Neumann boundary conditions at the bottom and/or free surface (section 2.7.2). By
application of the shear conditions defined by the outside stress tensor, in order to
obtain proper velocity profiles in the turbulent bottom or surface boundary layer, the
eddy-viscosity coefficients should be in general obtained from an appropriate turbulence
model.

2.7.6.1 Bottom shear

The bottom boundary condition for a rough bottom is given by an empirical formulation
in the form: √
τb ∂u u u2 + v 2
= νz = (2.187)
ρ ∂z C2
where τb is the bed shear stress and νz the vertical turbulent viscosity coefficient. The
roughness coefficient C is computed by assuming the existence of a logarithmic velocity
profile up to at least one tenth of the distance between the first two planes of the
computational mesh in the lowest part of the bottom boundary layer [103].

2.7.6.2 Free surface (wind) shear

The surface wind stress is in most cases computed according to the empirical formula
τw ∂uH ̺air
= νz = cf uw |uw | (2.188)
̺water ∂n ̺water

where uH is the horizontal velocity at the surface, νz vertical turbulent viscosity coeffi-
cient and uw the wind velocity. Equation (2.188) states a Neumann boundary condition
for the horizontal velocity components. The empirical coefficient cf is given e.g. by
Flather [45]:

0.565 · 10−3


 |uw | ≤ 5m/s
−3
cf =  (−0.12 + 0.137|uw | · 10 5m/s < |uw | < 19.22m/s (2.189)

2.513 · 10−3 |uw | ≥ 19.22m/s

2.7.7 Open boundary conditions


The open boundary conditions appear due to the fact that the flow domain must be
truncated for practical reasons. The mathematics forces us to select boundary conditions
there, which are physically artificial. Their form depends strongly on the type of the
mathematic character (hyperbolic, elliptic, or parabolic) of the equations to be solved.
At open boundaries the boundary conditions for the independent variables must be
given, providing appropriate information to obtain a stable and unique solution in the
64 2.7 Initial and boundary conditions for the flow

domain. Actually, the solutions computational fluid dynamics are usually driven by the
boundary conditions, and it is of extreme importance that well-posed and physically
realistic boundary conditions are applied.
The truncation of the computational domain results in two types of boundaries, namely,
the inflow and outflow boundaries. At the inflow boundaries the conditions are usually
well known and are easy to implement as imposed values. All components of the ve-
locity, density, free surface, pressure, temperature and tracer concentrations have to be
specified. However, for many applications the theory provides hints in order to reduce
the number of needed boundary conditions, e.g. in the case of the channel flow it is
enough to provide a given flux through the inflow boundary and require a zero normal
gradient of the depth.
The outflow boundary conditions require special care and are more difficult to apply
because they cannot be predicted a priori for unsteady flows. Theoretically, the stress
continuity, i.e., the dynamic boundary conditions (2.177) are to be provided. Commonly,
the outflow sections are positioned where the character of the flow is well known, e.g.,
approximately unidirectional or where the stresses are known. For example, for a fully
developed channel flow, or away from obstacles, there is no change of the velocity com-
ponents across the boundary and fn = −p as well as ft = 0. This yields the set of the
outflow boundary conditions [58, 161]:

∂T
p = pout or S = Sout =0 (2.190)
∂n

These conditions specify the pressure (or the water depth) and let the normal gradients
of other variables than the normal velocity be null. Such boundary conditions are im-
plemented in a very natural way in the finite element (and volume) methods. For the
normal velocity, there are theoretical concerns regarding outflow boundary conditions in
the form:
∂un
=0 (2.191)
∂n
are set, as stated by [58, 137]. This boundary condition is in theory under-determined,
resulting in general in an infinite number of solutions. For incompressible fluids, require-
ment ∇ · u = 0 by valid ∂un /∂n = 0 is over-restrictive for the tangential component,
causing uτ = 0. For viscous flows, the most proper boundary conditions are given by
dynamic boundary conditions, e.g. equation (2.180) set for all velocity components. In
practice however, ∂un /∂n = 0 is usually used successfully. It should be noted, however,
that the boundary conditions in the form (2.190) can be used only when all flows enter-
ing the domain are defined by imposed inflow boundary conditions. By applying outflow
boundary conditions special care must be taken when positioning the outflow sections.
Sensitivity studies should be made in order to demonstrate that the interior solution is
not affected by the outflow boundary.
When the simple forms of outflow boundary influence negatively the interior solution,
especially for unsteady flows, some kind of more complex boundary conditions must be
imposed, such as non-reflecting or absorbing boundary conditions (sections 2.7.8, 2.7.9).
2.7 Initial and boundary conditions for the flow 65

A usual, simple method to avoid applying these special and imperfect cases of boundary
conditions is to make the computational domain so large that the area of interest can
be treated as practically unbounded. Another method, e.g. for phenomena developing
in the domain interior, is to stop the simulation when the boundaries start to influence
the internal solution significantly.
In some situations, e.g., for the channel flow, the inflow and outflow can be specified
according to the type of the flow in order to provide a unique solution. E.g. one specifies
the flux (and initial depth only) at the inflow boundary and water depth at the outflow
one.
It should be noted that in some applications the inflow boundary may become an out-
flow one during the simulation (e.g. tidal flows), and vice versa (reverse flows), or even
a simultaneous inflow/outflow boundary (estuaries) – and the algorithm must react ap-
propriately. Sometimes advantage can be taken from special geometrical aspects of the
flow allowing cyclic or periodic boundary conditions where the outflow values are given
as inflow values or symmetry boundary conditions where the domain can be truncated
along lines of symmetry.
A specific type of the boundary conditions are the conditions for the global pressure or
the hydrodynamic (motion) pressure at these boundaries (section 2.7.11).

2.7.8 Non-reflecting boundary conditions

For flow problems dominated by advection or wave motion a special case of open bound-
ary conditions is needed at the open/liquid boundaries which artificially truncate the
computational domain. A boundary condition is needed which allows phenomena gener-
ated in the domain of interest to pass through the artificial boundary without undergoing
any significant distortion or without influencing the interior solution. The intrinsic diffi-
culty in imposing the boundary conditions of this kind arises due to the lack of knowledge
of the environmental behaviour outside the domain. Therefore, some extrapolation tak-
ing into account the mathematical character of the equations to be solved must be used.
Actually, most effective non-reflecting boundary conditions (NRBC) assume some spec-
ified behaviour or character of the flow (for free surface flows, mostly hyperbolic) just
outside the given boundary.
In cases where the simulation starts from an artificial or an awkwardly posed initial con-
dition (e.g. flows over obstacles starting with a uniform velocity field) the non-reflecting
boundary conditions are needed in order to allow the initially appearing wiggles or
erroneous flow structures to exit the domain without causing any instabilities. For sim-
ulations aiming at the achievement of a steady-state solution, the boundary conditions
appropriate for the final state are often applied. When the scheme is not dissipative
(physically or numerically), such boundary conditions may cause serious difficulties con-
cerning the solution convergence. A very important aspect of non-reflecting boundary
conditions is that they must be also applied at the inflow boundaries with imposed values
for flow variables.
66 2.7 Initial and boundary conditions for the flow

Most widespread non-reflecting boundary conditions for equations of hyperbolic nature


are based on the Sommerfeld radiation condition. It is valid for the wave equation in the
form:
∂2ϕ
− c2 (ω)∇2ϕ = 0 (2.192)
∂t2
The Sommerfeld radiation condition brings a simple linear relationship between the
partial time derivative of a variable in a wave motion and its spatial derivatives:
∂ϕ
+ c(ω) · ∇ϕ = 0 (2.193)
∂t

where c is a frequency-dependent wave phase velocity. It essentially states that the


(outgoing) wave propagates in the positive direction without changing its form. Or-
lanski [122] transposed the Sommerfeld radiation condition from the frequency to the
time domain replacing the frequency-dependent wave phase velocity by a time-varying,
propagation velocity, i.e. coefficient c(t):

∂ϕ
+ c(t) · ∇ϕ = 0 (2.194)
∂t

As can be easily seen, a single wave component of the form ϕ(x, t) = ϕ0 exp(k · x − ωt)
fulfils (2.194) when ci (t) = (ω/ki). In a finite difference stream function – vorticity model
applied for oceanic mesoscale flows, Orlanski [122] computed numerically the coefficient c
in the x-direction at an open boundary from the model spacing ∆x and ∆t and derivative
values at previous time-step levels:

cx = ∆x/∆t if −(∂ϕ/∂t)/(∂ϕ/∂x) > ∆x/∆t


cx = −(∂ϕ/∂t)/(∂ϕ/∂x) if 0 < −(∂ϕ/∂t)/(∂ϕ/∂x) < ∆x/∆t (2.195)
cx = 0 if −(∂ϕ/∂t)/(∂ϕ/∂x) < 0

assuring model stability and that no information (waves) from outside can enter the
computational domain. From (2.194) a Dirichlet BC for the new time level is obtained
[122]. The Orlanski condition is local both in time and space and works well when
the incident waves are monochromatic (ω = const) or for non-dispersive long waves,
where c is constant and independent of ω. For unsteady, dispersive waves and non-
monochromatic wave motions problems arise. However in many applications for e.g.
internal turbulent flows, the boundary condition in form (2.194) where c can be so
calibrated that it is independent on the outflow surface position and guarantees overall
conservation (the outflow mass flux equal to the incoming one), proved to yield good
results [44].
Enquist and Majda [42] developed a perfectly non-reflecting boundary condition for the
wave equation (2.192) that is non-local in time and space (they called it absorbing BC).
It required the complete time history of variables along the boundary in order to update
them in the next time step. This boundary condition, very awkward for numerical
computation, can be replaced by a derived local approximation to the general solution
2.7 Initial and boundary conditions for the flow 67

up to a required order of accuracy n. Higdon [75, 76] developed a more general form of
these boundary condition for a 2D finite difference wave model (including the angle of
incidence αj of the incoming waves):
 !
n
Y ∂ c ∂ 
 + ϕ=0 (2.196)
j=1 ∂t cos αj ∂x

This boundary condition is perfectly non-reflecting for a plane wave hitting the boundary
at one of the a priori given angles αj . Its first approximation n = 1 and for α1 = 0 is
the well-known Sommerfeld radiation condition.
Hedstrom [69] developed non-reflecting boundary conditions for non-linear hyperbolic
systems derived from the characteristic form of the hydrodynamic equations with an
application for gas dynamics (shock waves). Thompson [154, 155] developed this idea
further to multidimensional boundary conditions describing incoming and outgoing waves
as well. An application of so obtained boundary conditions in a wave model for shallow
and deep water waves is provided by Schröter [141, 142].
For this method, the set of n governing equations is written in the following form:
∂U ∂U
+A +C=0 (2.197)
∂t ∂x

where U is a vector of n primitive variables, A is a n × n matrix and C is a term which


does not contain any derivatives of the variables appearing in U. When the system
(2.197) is hyperbolic, the matrix A has n real eigenvalues λi . The diagonal matrix of
eigenvalues Λ, Λii = λi is obtained from similarity transformation:

SAS−1 = Λ (2.198)

where S contains as rows the left (row) eigenvectors li of A defined by li A = λi li and


the columns of S−1 are the right (column) ri eigenvectors of A defined by Ari = λi ri .
These vectors are normalised, so that li rj = δij .
Multiplying (2.197) by S yields the following characteristic form of the governing equa-
tions:
∂U ∂U
S + ΛS + SC = 0 (2.199)
∂t ∂x
or, using the eigenvectors,
∂U ∂U
li + λi li + li C = 0 (2.200)
∂t ∂x
If a function V can be defined, so that

dVi = li dU + li Cdt (2.201)

then (2.200) transforms to


∂Vi ∂Vi
+ λi =0 (2.202)
∂t ∂x
68 2.7 Initial and boundary conditions for the flow

ν(x)
6

ν(x)

-x
 - -
no damping damping zone

Figure 2.4: Defining a damping zone for absorbing boundary conditions.

The equation above is a set of wave equations stating, that the functions Vi (Riemann
invariants) do not change along a characteristic curve defined by dx/dt = λi in the plane
xt. It plays an analogous role to the Sommerfeld radiation condition (2.193).

According to the idea of Hedstrom [69], it is assumed, that the functions Vi for the
incoming waves are independent of time at the boundaries. For the waves described
by incoming characteristics it is assumed that ∂Vi /∂t = 0 (or by setting an imposed
Dirichlet value there). For the outgoing waves of known λi , the values at the boundary
nodes can be found by the method of characteristics from (2.202).

The main difficulty in applying this method of determining non-reflecting boundary


conditions at the boundaries is that Vi can be determined from (2.201) when A and C
are constant in the whole domain and (2.201) is integrable. However, arguing that the
form (2.200) can be obtained for most practical cases, Thompson [154, 155] develops a
general boundary condition formalism for hyperbolic systems.

The non-reflecting boundary conditions based on the Thompson method were also devel-
oped for the two-dimensional code Telemac2D (shallow water equations) for advection-
dominated flows [73]. The various source terms represented by C in (2.197) are treated
as known and computed explicitly in the framework of the operator splitting. For the
outgoing waves, the method of characteristics for the boundary nodes is applied, ex-
actly as for the internal nodes in the advection step (section 3.5.2). As the advection
velocity the eigenvalues λi (i.e. for shallow water equations (u, u + c, u − c) for the
direction normal to a given boundary node are assumed. The contributions from rele-
vant Riemann invariants Vi are added to obtain boundary values for u, v, h and tracer
concentrations. For Telemac3D the non-reflecting boundary conditions are currently
being developed elsewhere (J.-M. Hervouet, LNH/EDF, 1998, personal communication).
For entirely barotropic flows it is theoretically possible to use the routines developed
for Telemac2D. Consequently, in this work, no specific developments in this direction
are made and only the simplest form of non-reflecting boundary conditions is applied
(section 5.7).
2.7 Initial and boundary conditions for the flow 69

2.7.9 Absorbing boundary conditions

The absorbing boundary conditions are an effective method of truncating the computa-
tional domain. The idea of this approach is to ensure that waves generated inside the
domain are effectively absorbed at a given boundary or some buffer area before it. Typ-
ical examples of an application are numerical wave models for wave tanks or channels,
where waves generated at a wave making surface run through an area of interest and
then are absorbed at a numerical beach, sponge layer or a piston-like absorbing surface.
In the numerical beach or sponge method, an additional dissipative term is added to
the dynamic free surface condition or to the kinematic condition or to both [32]. A
coefficient playing the role of a viscosity coefficient is set to zero all over the area, except
in a zone before the artificial boundary where it gets non-zero values. According to the
particular application, it damps the wave elevation, fluid velocity, wave potential, or
a combination of them. Water waves, passing through the region, where the damping
is applied, systematically lose their energy (figure 2.4). When the damping (sponge)
zone is sufficiently long compared to the water wave length, the waves may be entirely
absorbed. This approach works effectively as a low-pass filter for the waves and the
absorbing coefficient can be tuned to the spectrum of the incoming waves. The piston-like
absorbing boundary condition is a Neumann boundary condition for the wave potential
motivated by properties of the incident waves to be absorbed and derives from a study
of the control of the physical wave absorbing devices in wave tanks [32].
For a 3D hydrodynamic model, this method can be most effectively applied for the free
surface determination with the kinematic boundary condition and computed with the
SUPG method, as described in section 3.6.2.1. It may be used together with setting an
analogical damping zone for the velocity components through viscosity coefficient values
near the boundaries.

2.7.10 Boundary conditions for the tracer

Fluxes or imposed values of the tracers must be given at the domain boundaries. For
example, in case of the temperature T :

∂T
λ = −qb or T = Tb (2.203)
∂n

where qb is the heat flux at the given boundary and λ the thermal conductivity. The
values are usually obtained from various empirical formulations or for instance, in the
case of pollutants, according to discharge characteristics. For many applications the
no-flux condition is appropriate:
∂T
=0 (2.204)
∂n
70 2.7 Initial and boundary conditions for the flow

2.7.11 Boundary conditions for the hydrodynamic pressure


equation
Boundary conditions for hydrodynamic pressure can be obtained from purely physi-
cal deliberations or from more abstract mathematical reasoning based on the particular
features of the numerical algorithm. In the case of a Poisson pressure equation from frac-
tional formulation written for the hydrodynamic pressure (section 2.5.3), the particular
type of boundary conditions depend strongly on the physical system to be modelled.
As a starting point it should be realised that setting the imposed (Dirichlet) value of
null for hydrodynamic pressure at a boundary has the physical meaning of applying
purely hydrostatic pressure there. Hydrostatic pressure means no fluid motion, or that
hydrostatic approximation is assumed to be valid. As a further consequence, on these
boundaries, where an imposed value of zero dynamic pressure appears, it is automatically
required that no velocity correction for ũ is needed in order to fulfil the incompressibility
condition.
Therefore, hydrostatic approximation π = 0 is an acceptable boundary condition for
these open inflow boundary sections, where the velocity is thoroughly defined and its
divergence is assured to be equal to zero.
For the viscous flows, at the open boundaries and at the free surface the dynamic bound-
ary conditions (2.177) can be implemented. At the free surface, the following Dirichlet
BC is valid:
∂un
πs = ̺ν (2.205)
∂n
where for simplicity the atmospheric pressure is set to zero and no outside stress (e.g.
wind) is present. When the gradients of the free surface are small enough, this boundary
condition can be simplified to:
∂w
πs = ̺νz (2.206)
∂z
It can be seen that in the inviscid case ν = 0 an imposed value of π = 0 at the free
surface is appropriate. However, when the rigid lid approximation for the free surface is
applied, a Neumann BC should be provided (section 2.7.5).
For all open vertical sections (2.177) yields for π:

π = ̺νn · ∇un (2.207)

From the equation above it can be immediately seen that for these outflow sections,
where boundary conditions (2.190) apply (i.e., ∂un /∂n = 0):

∂π
=0 (2.208)
∂n

The zero Neumann boundary condition may also be appropriate for the inflow bound-
aries.
2.8 Algorithm for the non-hydrostatic equations 71

A Neumann boundary condition for the pressure Poisson equation (2.142) at the solid
walls or at the bottom can be obtained from the equations (2.143) and the impermeability
condition for the final velocity n · un+1 = 0:
1 1
(n · un+1 − n · ũ) = − n · ∇π (2.209)
∆t ̺0

obtaining:
∂π ̺0
= n · ∇π = ũn (2.210)
∂n ∆t
Condition (2.210) yields another constraint on the hydrodynamic pressure field to be
obtained from (2.142). Namely, it should provide not only the correction of the in-
termediate velocity field in the domain interior, but also guarantee the impermeability
condition fulfilment at the solid walls. If the values of ũn have already been set to zero by
securing the impermeability boundary condition in the hydrostatic part of the algorithm
(2.139-2.141), equation (2.210) yields ∂π/∂n = 0. Boundary condition (2.209) is valid
also in all those cases, when the intermediate velocity ũ differs for some reason from a
prescribed Dirichlet value at the boundary and must be corrected.
The methods of assuring impermeability or fulfilment of a Dirichlet BC for the velocity
by obtaining specific hydrodynamic pressure gradients require a very good approxima-
tion of the derivatives at a given boundary (e.g. regular elements and fine spacing). This
can be avoided when the Dirichlet and impermeability conditions are already fulfilled
by the intermediate velocity field. In this case, instead of (2.209) a simple zero Neu-
mann boundary condition for the hydrodynamic pressure appears, which is naturally
implemented in the method of finite elements. ∂π/∂n = 0 means only that no velocity
correction is needed in the normal direction to the boundary. Nothing is given, for ex-
ample, in the tangential direction and the zero derivative is only approximately fulfilled.
In practice, the impermeability condition (2.184) can additionally be set.
In the case of projection-2 method, instead of hydrodynamic pressure a correction to
its value is found through the pressure Poisson equation. In this case in practice most
simple boundary conditions are adequate [57]: a zero Neumann BC at the solid and open
boundaries and zero Dirichlet BC at the free surface.

2.8 Algorithm for the non-hydrostatic equations

2.8.1 Equation set to be solved


For the sake of clarity, a set of three-dimensional equations (2.22–2.27) is written here
in the form applied in the non-hydrostatic algorithm, with pressure terms explicitly
introduced into them:
"Z #
∂u ∂S ∂ S ∆̺ 1 ∂π
+ u · ∇u = −g −g dz − − wfH + vfV + ∇ · (ν∇u)
∂t ∂x ∂x z ̺0 ̺0 ∂x
72 2.8 Algorithm for the non-hydrostatic equations

"Z #
∂v ∂S ∂ S ∆̺ 1 ∂π
+ u · ∇v = −g −g dz − − ufV + ∇ · (ν∇v)
∂t ∂y ∂y z ̺0 ̺0 ∂y
∂w 1 ∂π
+ u · ∇w = − + ufH + ∇ · (ν∇w)
∂t ̺0 ∂z
∂u ∂v ∂w
+ + =0 (2.211)
∂x ∂y ∂z
∂T
+ u · ∇T = ∇ · (ν T ∇T )
∂t
̺ = ̺(s, T )

Additionally, an equation for the free surface position must be solved.

2.8.2 Introducing remarks


The non-hydrostatic version is a far-reaching further development of the existing code
Telemac3D (see section 2.3). The new algorithm is coded in a parallel way to the
previous one, so that direct comparison between the new non-hydrostatic and the old
hydrostatic versions is possible by simply changing a steering variable. The new algo-
rithm is outlined in figure 2.5.
For the treatment of the governing equations (2.211) the decoupled method based on the
operator splitting, and a projection method based on splitting the global pressure into
hydrostatic and hydrodynamic part is chosen. The hydrodynamic pressure component
is found by applying the pressure equation from the fractional step formulation. Height
function methods for representing and tracking of the free surface are selected. Addi-
tionally, the appropriate boundary conditions are implemented. The developments are
optimal for the existing solution algorithm structure of Telemac3D.
Compared to the structure of the previously existing hydrostatic algorithm, briefly de-
scribed in section 2.3, the two-dimensional pressure – free surface – continuity step was
removed. In consequence, the terms representing the gradients of the free surface, the
buoyancy and Coriolis forces are moved to the diffusion step where they appear as source
terms of the diffusion equation. The modified diffusion and convection steps are per-
formed not only for the horizontal components of the velocity, but obviously also for the
vertical one. They yield the intermediate solution for the velocity field, which does not
have to fulfil the incompressibility condition (fractional step intermediate solution).
The removed pressure – free surface – continuity step is replaced by two new steps.
First, the continuity step is implemented, where the pressure Poisson equation is solved
and the projection of the velocity is performed. This step yields the solenoidal, three
dimensional velocity field. Then, in the free surface step the position of the free surface
is found.
The most important difference is in the treatment of the vertical component of the ve-
locity. In the previous algorithm the vertical component of the velocity is adjusted to
the horizontal velocity ones by means of integrating the continuity equation (2.56) so
that the incompressibility condition is satisfied. In this code, the vertical component
2.8 Algorithm for the non-hydrostatic equations 73

of the velocity is obtained from the vertical Navier-Stokes equation. Due to the hy-
drostatic approximation, the previous algorithm did not take into consideration vertical
accelerations, whereas the new algorithm does.
The characteristic feature of the non-hydrostatic algorithm is splitting the pressure into
two components: (1) the hydrostatic pressure resulting from the density in situ and
(2) the hydrodynamic pressure, as defined and described in section (2.5.5.2). For flows,
where the dynamic pressure gradients are large, the projection-2 method can be consid-
ered, where in the continuity step a pressure correction to the global pressure is found
and the hydrodynamic pressure gradients are already taken into consideration in diffusion
step.
Another typical feature is the free surface algorithm, where three numerical methods of
finding the free surface are provided for the user to choose from. The structure of the
algorithm and its software realisation allows its easy modifications in future.
In this section the outline of the implemented algorithm is provided, whereby the nu-
merical details are presented in chapter 3. For an overview, Table 2.1 is provided. The
developments described in this work do not concern the transport and turbulence mod-
elling as well as the equation solvers – they are listed in Table 2.1 for completeness.

2.8.3 Operator splitting


The time derivatives in the Navier-Stokes equations as well as transport equations for the
tracers are treated explicitly in the framework of the operator splitting (2.3.1). There-
fore, the fractional step method of solving three-dimensional hydrodynamic equations
described in subsection 2.5.4 can be applied. The equations are split into fractional
steps and treated separately by appropriate numerical methods. The explicitly formu-
lated time derivative of the variable f is:

∂f f n+1 − f d f d − f a f a − f n
= + + (2.212)
∂t ∆t ∆t ∆t

In the equation above, f stands for the components of the velocity u, v, w and super-
scripts denote the intermediate levels of advection a and diffusion d steps, as well as the
time step levels n and n + 1. The final value f n+1 is obtained in the continuity step. For
the tracers (f represents then temperature, salinity, pollutants, etc.) the continuity step
does not apply and:
∂f f n+1 − f a f a − f n
= + (2.213)
∂t ∆t ∆t

2.8.4 Advection step


As previously, the advection (also hyperbolic) step is:

fa − fn
+ u · ∇f = 0 (2.214)
∆t
74 2.8 Algorithm for the non-hydrostatic equations

Table 2.1: Basic characteristics of the solution algorithm


Computational aspect Selected approach Remarks
(algorithm step)
Mesh σ-mesh only vertically variable
Advection step Characteristics or SUPG-FEM entirely 3D
Diffusion step Standard Galerkin FEM entirely 3D
hydrostatic and Coriolis
forcing as source terms
Pressure treatment Split:
– hydrostatic: obtained explicitly
– hydrodynamic: obtained from PPE
Continuity projection method (projection-1)
Velocity projection explicitly in FEM from hydrodynamic pressure
Free surface height function only single-valued
representation no breaking waves
Free surface (a) from kinematic BC
updating – with characteristics unlimited stability
– with SUPG limited stability
(b) from conservative FS eq. conservative
(c) from St.-Venant eqs. shallow water only
Free surface BC normal:
– viscous/inviscid inviscid: p = pout
tangential:
– viscous wind stress (empiric)
Bottom BC tangential:
– viscous bottom stress (Chezy)
– inviscid impermeability, adherence
Lateral wall BC impermeability
Lateral open BC (a) free outflow Neumann
(b) inflow imposed/Neumann
(c) non-reflecting BC for simple cases only
(d) absorbing BC wave problems
Turbulence – constant eddy-viscosity
modelling – mixing length damping functions
by stratification
– k−ε
Transport – active tracers influence density
– passive tracers
– sediment deposition/erosion
– particle tracking
Equation solvers iterative, preconditioned Element-by-element
conjugate gradient methods technique
2.8 Algorithm for the non-hydrostatic equations 75

where f denotes velocity components as well as tracers. The non-linear terms of the
equation set (2.211) are treated using the method of characteristics, or SUPG method.
The only difference to the hydrostatic algorithm is that the vertical velocity component
is treated as well. When SUPG is applied, the advection step is not treated separately,
but together with the diffusion step.

2.8.5 Diffusion step


In the diffusion (also parabolic) step the diffusive terms (2.211) are treated using the
finite element method:
ud − ua
= ∇ · (ν f ∇u) + (FCor + F∇S + F∆̺ ) + qu (2.215)
∆t
Fi are different forcing terms in equations (2.211): Fcor Coriolis terms, F∇S free surface
gradients, and F∆̺ density gradients.
 
−fH w + fV v
FCor =

−fV u  (2.216)
fH u

−g ∂S
 
∂x
F∇S =  −g ∂S (2.217)
 
∂y 
0
S ∆̺
 
∂ R
−g ∂x z ̺0
dz
∂ RS ∆̺
 
F∆̺ =
 −g ∂y z ̺0
dz 
 (2.218)
0

qu represent other forcing accelerations, e.g. non-inertial ones.


For tracers T , the diffusion step takes the following form:

Td − Ta
= ∇ · (ν T ∇T ) + qT (2.219)
∆t
where qT describes the source terms needed to simulate discharge of effluents or heat
sources, for example.
As the result of the diffusion and advection steps the intermediate velocity field ũ as
described in subsection 2.5.4 is found.
In this stage, the found horizontal velocity components ũ = ud and ṽ = v d are compara-
ble with the results for un+1 and v n+1 which would result from solving three-dimensional
shallow water equations (2.37–2.41) using the previous hydrostatic Telemac3D algo-
rithm as described in section 2.3. The difference is that also the vertical component
w̃ = w d is simultaneously computed from the third, vertical momentum conservation
equation. The continuity constraint is not fulfilled by this field.
76 2.8 Algorithm for the non-hydrostatic equations

2.8.6 Continuity step

In the first part of the continuity step the 3D elliptic Poisson equation for the hydrody-
namic pressure π (2.142) is solved using the finite element method (analogically to the
diffusion step):
̺0
∇2 π = ∇ · ud (2.220)
∆t

and the resulting solenoidal velocity is found explicitly from the gradients of the hydro-
dynamic pressure, i.e. the projection of the intermediate velocity is performed (equations
2.143):
∆t
un+1 = ud − ∇π (2.221)
̺0

After this stage the resulting three-dimensional velocity field (u, v, w) satisfies the in-
compressibility condition.
When the projection-2 method is chosen, instead of hydrodynamic pressure π a more
abstract variable, namely the pressure correction to the pressure applied in the advection-
diffusion step is found (section 2.8.9).

2.8.7 Free surface step

Using the resulting, final velocity, the position of the free surface in the next step can
be found.
There exist three versions of this stage. The first two versions use the kinematic boundary
condition (2.158):
S n+1 − S n ∂S ∂S
= −us − vs + ws (2.222)
∆t ∂x ∂y

and the third one the conservative formulation of the free surface equation (2.159):

S n+1 − S n ∂ Sn ∂ Sn
Z Z
=− un+1 dz − v n+1 dz (2.223)
∆t ∂x −B ∂y −B

Both equations are partial differential equations of a hyperbolic type. Equation (2.222)
solved by the semi-implicit Streamline Upwind Petrov-Galerkin method (SUPG), or us-
ing the method of characteristics. Equation (2.223) is solved explicitly in finite elements
or semi-implicitly using SUPG, while the velocity is integrated using the free surface
elevation at the time level n.
For these flows, where only baroclinic motions are of importance, the position of the free
surface can be frozen. The appropriate boundary conditions for velocity (rigid lid) have
to be provided.
2.8 Algorithm for the non-hydrostatic equations 77

Non-hydrostatic algorithm outline

Operator-splitting

∂u un+1 − ud ud − ua ua − un
= + +
∂t ∆t ∆t ∆t

Source terms (2D/3D, in FE):


 Sn 
Z
∆̺ n
Fu = −g∇H S n − g∇H  dz  − 2Ω × un + qu
̺0
z

Advection step (3D, hyperbolic):

ua − un
+ u · ∇u = 0
∆t
Diffusion step (3D, parabolic):

ud − ua
= ∇ · (ν∇u) + Fu
∆t
Poisson equation (3D, elliptic):
̺0
∇2 π = ∇ · ud
∆t
Velocity projection (3D, in FE):

∆t
un+1 = ud − ∇π
̺0

Free surface step (2D, hyperbolic) in two versions:

S n+1 − S n ∂S ∂S
= −us − vs + ws
∆t ∂x ∂y

ZS ZS
S n+1 − S n ∂ ∂
=− udz − vdz
∆t ∂x ∂y
−B −B

Two-step tracer advection-diffusion (hyperbolic-parabolic):

Tn − Td Td − Ta
+ = ∇ · (ν T ∇T ) + qT − u · ∇T
∆t ∆t

Figure 2.5: Non-hydrostatic algorithm outline.


78 2.8 Algorithm for the non-hydrostatic equations

2.8.8 Free surface from vertically integrated momentum equa-


tions
In many geophysical flows the position of the free surface can be treated as a result of
purely barotropic motions, e.g. can be found using the vertically integrated St.-Venant
equations in the form analogical to (2.54). It means that the free surface can be
found using the shallow water approximation, but the internal motion modes are to be
resolved as well. It must be mentioned, that the restriction on the shortness of the free
surface waves that can be modelled remains. In this case the advantages of the robust
Telemac2D algorithm can be taken in finding the free surface.
In this version the advection and diffusion stages are virtually the same as described in
previous sections. In the diffusion step, however, the terms representing the free surface
gradients are set to zero and treated in the free surface step. Consequently, the baroclinic
(density gradients) part of the hydrostatic pressure is treated in the diffusion step and
the barotropic (free surface gradients) in the vertically integrated equations.
After the diffusion step the free surface step follows, in which, using the Telemac2D,
the following equations are solved in order to find the intermediate velocity field ũ and
the new free surface elevation:
∂h ∂hu ∂hv
+ + =0 (2.224)
∂t ∂x ∂y
usurf − ud ∂h ∂B
+g = −g (2.225)
∆t ∂x ∂x
vsurf − v d ∂h ∂B
+g = −g (2.226)
∆t ∂y ∂y

The solution procedure is described in section 2.3.3. After solving this set of equations,
the three dimensional velocity field is obtained by applying equations (2.55).
After this step the continuity step follows, where the resulting, solenoidal velocity field
is obtained.

2.8.9 Projection-2
For these flows, where the gradients of the hydrodynamic pressure are large, i.e. com-
parable or even greater than the gradients of the hydrostatic pressure, a version of the
non-hydrostatic algorithm based on the projection-2 method can be thought of (section
2.5.4). In this method in the diffusion step not only gradients of hydrostatic pressure,
but also the known hydrodynamic pressure gradient from the previous time step appears,
eq. (2.215). In this way too large differences between the divergent intermediate and
final divergence-free velocity are avoided.
This method requires not only that the initial velocity field must be divergence-free, but
also that the hydrodynamic pressure field must be properly initialised as well. Usually,
a zero-field is appropriate, or a known analytical solution. However, when this is not the
2.8 Algorithm for the non-hydrostatic equations 79

case, an initialisation step for velocity and hydrodynamic pressure must be performed, as
described in section 2.7.1 and gradients of the initial hydrodynamic pressure computed.
After initialisation, the hydrodynamic pressure gradients are treated in the diffusion step.
In the continuity step a pressure correction to the hydrodynamic pressure φ = π n+1 − π n
is found and the projection step for the velocity is executed. The hydrodynamic pressure
for the next step is simply π n+1 = π n + φ. However, π n+1 is not required in the next
time step – only its already computed gradients are needed and consequently updated.
80 2.8 Algorithm for the non-hydrostatic equations
Chapter 3

Numeric algorithm description

The difference between theory and practice


is smaller in theory than in practice.
Anonymous

This chapter provides a description of the numeric realisation of the non-hydrostatic


algorithm presented in chapter 2. Due to operator splitting, the algorithm has a modular
structure with distinctive parts with clear mathematical properties: advection, diffusion,
continuity, free surface. Verification and performance tests of these steps treated as
separate units are given in chapter 4. Verification of the complete algorithm is presented
in chapter 5.

3.1 Computational domain


In this section the assumptions taken in order to deal with the time-variable extents of
the computational domain are discussed.

3.1.1 σ-mesh and σ-transformation


The three-dimensional mesh structure known as σ-mesh, was presented first by Phillips
[125]. The computational domain Ω is bounded by the bottom z = B(x, y), time-variable
free surface z = S(x, y, t) (both single-valued functions) and vertical lateral boundaries
(figure 2.1). The bottom surface is discretised by a mesh of triangles, forming the two-
dimensional base mesh. The three-dimensional mesh is obtained by duplicating it in the
vertical direction in such a way that the z co-ordinates of the nodes are defined by:

z = B(x, y) + z∗ (S(x, y, t) − B(x, y)), z∗ ∈ [0, 1] (3.1)

At various mesh levels defined by a discrete set of z∗ values, the x, y node co-ordinates
remain the same as at the bottom base mesh. In consequence, the domain is divided into

81
82 3.1 Computational domain

XX
X
z  S
 z∗
6 ``` SS 
XX 6
1
X
 QQ 
```
 PP 
``` σ-
 PP 
```
 PP 
``` PP
```  PP 
  
``` 0

Ω Ω
Figure 3.1: σ-transformation. Side view of the real and transformed mesh.

a number of tiers formed by prismatic elements with 6 nodes, whose lateral, quadrilateral
faces are oriented vertically. Triangular bases of the prisms follow the mesh surfaces. The
vertical projection of this mesh on a horizontal surface builds a two-dimensional mesh
of triangles, called further the base mesh Ω0 . The vertical co-ordinates of the real mesh
nodes change with the time-variable position of the free surface, but the horizontal co-
ordinates remain constant. The z∗ values for each mesh level are also constant in time,
which allows a transformation of the time-variable real mesh into a σ-mesh, fixed in time
(figure 3.1): 
 x∗ = x


y∗ = y (3.2)
 z∗ =

 z−B(x,y)
S(x,y,t)−B(x,y)

As it will be shown later, in computation of various matrices resulting from FEM for-
mulation advantages are taken of the stationary σ-mesh geometry.

3.1.2 Comments on the σ-transformation


The σ-transformation defined by (x, y, z) → (x, y, z∗) transforms the time-dependent real
mesh into a stationary σ-mesh, where all mesh levels are parallel and horizontal, what is a
very important advantage. As it can be easily proved, the Jacobian of this transformation
is equal to the water depth: |J| = det|∂xj /∂x∗i | = h(x, y, t) = S(x, y, t) − B(x, y).
The surfaces defining the free surface and the bottom, where important boundary con-
ditions are applied, are discretised by triangles of the base mesh, following the bottom
topography or time-variable free surface configuration. In σ-transformation these bound-
aries are horizontal. The lateral boundaries consist of quadrilaterals oriented vertically.
In σ-transformation all these elements are rectangular.
The main disadvantage of such a mesh structure is the fact that with a constant number of
horizontal mesh levels deeper parts of the domain can be under-discretised and shallower
ones over-discretised. In most practical cases some compromise must be met. The lateral
boundaries have to be represented by vertical walls. For modelling of flooding and drying
areas as e.g. tidal flats special strategies are needed (masking of elements, [74]).
3.1 Computational domain 83

Another well-known disadvantage of the σ-transformation is the condition of hydrostatic


consistency, especially important for the finite difference models [110]. For very steep
bottom gradients (or free surface ones) a situation may occur where one or more of the
nodes at the bottom of a real mesh element are situated above some nodes at the element
top. For the reasons mentioned in section 3.4.4.2 this situation is not so dangerous as
in FDM, but ambiguities by computation of horizontal gradients in elements may occur,
leading to errors. This situation can be avoided already during mesh generation by finer
horizontal discretisation of greater bottom slopes or by prescribing lower discretisation
in the vertical. Steep free surface slopes cause similar problems, which cannot be avoided
without mesh adaptivity.
An advantage of the σ-transformation is the fact that the form of time derivative of a
variable f (x, y, z, t) does not change during transformation to f (x, y, z∗, t), and for the
advection equation:
! ! ! !
df (x, y, z, t) ∂f ∂f ∂f ∂f
= +u +v + w∗ =0 (3.3)
dt ∂t x,y,z∗
∂x y,z∗ ,t
∂y x,z∗ ,t
∂z∗ x,y,t

where w∗ is the transformed vertical velocity component w in σ-co-ordinates (x, y, z∗):


! ! ! !
dz∗ ∂z∗ ∂z∗ ∂z∗ ∂z∗
w∗ = = + u + v + w (3.4)
dt ∂t x,y,z
∂x y,z,t
∂y x,z,t
∂z x,y,t

The horizontal components do not change. When a σ-transformed advection velocity


u∗ = (u, v, w∗) is defined, the σ-transformation does not change the form of the advection
equation.
Applying σ-transformation to the diffusion equation leads to lengthy and complicated
expressions, which are useless in practice [110]. While in the advection step realised
with the method of characteristics in the σ-mesh full advantage can be taken of its
simple geometry (section 3.5.2), in the diffusion step (and/or advection with the SUPG-
method) another approach is chosen (sections 3.4, 3.5.3). By computation of the matrices
resulting from FEM-formulation, the appearing interpolation functions in the real mesh
elements are transformed into σ-mesh in order to simplify computations of various matrix
elements.
A practical simplification is that in the σ-mesh both the impermeability condition at the
bottom (2.185) and kinematic boundary condition at the surface (2.158) are equivalent
to setting the transformed vertical velocity component w∗ to null at the bottom and
surface.

3.1.3 Reference element and interpolation functions

The elements of the real mesh Ω are prisms with 3 quadrilateral faces oriented vertically
and two triangular faces at the top and at the bottom, which are not horizontal in general
84 3.1 Computational domain

6
@
@
@@ 5
4

3
\

 \
```
1 ``` \
``
\
2
Figure 3.2: The prismatic element of the real mesh. Local numbering of nodes is indi-
cated.

(figure 3.2). The model variables are computed at the nodes i = 1, ..., N of the mesh Ω
being the vortices of the prisms. Each function f is approximated by:
N
X
f= fi ϕi (3.5)
i

where fi are function values at the nodes and ϕi the Lagrangian basis interpolation
functions corresponding to a node i, which equal 1 at the node i and 0 at all other
nodes.
In order to simplify the computations of different matrices resulting from FEM formula-
tions, they are performed in reference elements of a simple form and with reference base
functions. Each real element is transformed into a reference one, and the elementary
contributions are computed in each element in a decoupled way. In Telemac3D, the
Element-by-Element technique is applied and a classical global matrix assembly is not
performed. The diagonal and extra-diagonal elements of the element matrices are sepa-
rately stored and vectorised procedures for performing algebraic operations are provided.
The reference element in Ωref is shown in figure 3.3. The basis functions ψi corresponding
to the 6 vortices (nodes) i of the reference elements are listed as follows:
ψ1 = (1 − α − β)(1 − γ)/2 ψ4 = (1 − α − β)(1 + γ)/2
ψ2 = α(1 − γ)/2 ψ5 = α(1 + γ)/2 (3.6)
ψ3 = β(1 − γ)/2 ψ6 = β(1 + γ)/2

The basis functions ϕi in the real mesh Ω are obtained by an iso-parametric transforma-
tion of ψi . The vortices of the prismatic elements in the real mesh Ω have co-ordinates
(xi , yi, zi ). The transformation F of the local co-ordinates of a given point P (α, β, γ)
within a given element into its real co-ordinates P (x, y, z) is given by:
6
X 6
X 6
X
x= xi ψi (α, β, γ) y= yiψi (α, β, γ) z= zi ψi (α, β, γ) (3.7)
i=1 i=1 i=1
3.1 Computational domain 85

γ
6

4 6
1 !
!!!
!
!!!
!!
5 !!

0 1 -
β

1
1
–1 !

!!! 3
α !
!!!
!
! !!
2
Figure 3.3: The prismatic reference element. Local numbering of nodes and the local
axes are shown.

The formula for (x, y, z) can be simplified because the lateral faces of prisms are vertical,
and x1 = x4 , y1 = y4 ; x2 = x5 , y2 = y5 ; x3 = x6 , y3 = y6 . In consequence, one obtains:
x = (1 − α − β)x1 + αx2 + βx3 (3.8)
y = (1 − α − β)y1 + αy2 + βy3
1 1
z = [(1 − α − β)z1 + αz2 + βz3 ](1 − γ) + [(1 − α − β)z4 + αz5 + βz6 ](1 + γ)
2 2
The Jacobian |J| of this transformation is:
1
|J| = [(x2 − x1 )(y3 − y1 ) + (x1 − x3 )(y2 − y1 )] (3.9)
2
[(1 − α − β)z4 + αz5 + βz6 − (1 − α − β)z1 − αz2 − βz3 ]

The functions ϕi are given by:


ϕi (x, y, z) = ψi (F −1 (x, y, z)) (3.10)

For computation of the diffusion matrix (3.28) in section 3.3.1 using transformation from
the real to the reference element a difficulty arises. The transformation of the expressions
typical for the diffusion matrix into the reference element yields:
Z
∂ϕi ∂ϕj
dxdydz = (3.11)
Ω ∂x ∂x ! !
∂ϕi ∂α ∂ϕi ∂β ∂ϕi ∂γ ∂ϕj ∂α ∂ϕj ∂β ∂ϕj ∂γ
Z ref
+ + + + |J| dαdβdγ
Ω ∂α ∂x ∂β ∂x ∂γ ∂x ∂α ∂x ∂β ∂x ∂γ ∂x
86 3.1 Computational domain

η η
6 6
1 1
@
@ 4 3
3@
@
@
@ -
@ ξ
@ –1 1
@
@
1 2@ 1 2
@ - ξ
0 1 –1

Figure 3.4: The two-dimensional reference elements.

where each expression in form ∂α/∂x brings a factor 1/|J| (obtained from Jacobi matrix
inversion). As a result a polynomial expression divided by |J| is obtained for integration.
With the Jacobian given by (3.9) the denominator of the expression to be integrated in
(3.11) is not constant. The analytical integration of such a rational fraction leads to very
complex logarithmic expressions, which are useless for practical purposes. Therefore,
advantage is taken of the simpler geometry of the σ-mesh Ω∗ (details in section 3.4).
As the result of the σ-transformation, the triangular faces of the real elements at the
bottom and top become horizontal, and z1∗ = z2∗ = z3∗ , z4∗ = z5∗ = z6∗ . The Jacobian of
the transformation F ∗ of the co-ordinates of a point P (α, β, γ) to P (x, y, z∗) appears to
be constant:
1
|J∗ | = [(x2 − x1 )(y3 − y1 ) + (x1 − x3 )(y2 − y1 )] (3.12)
2

and the expressions like (3.11) can be integrated analytically without difficulty by apply-
ing mathematical expert systems [72, 91]. The main advantages of this arduous procedure
(the formulas obtained tend to be very lengthy) are: (1) the obtained formulas are exact
compared to numeric integration, (2) they are vectorised in a straightforward way.

3.1.4 Two-dimensional elements

In some parts of the algorithm a two-dimensional approach is needed, as e.g. in the free
surface step or by computation of the boundary terms in the diffusion step (section 3.4).
In the case of the free surface and bottom, triangular elements with linear interpolation
are applied; for lateral boundaries quadrilaterals. The reference elements are: a triangle
with vortices in (0, 0), (0, 1), (1, 0) and a square comprised of co-ordinate points (−1, −1),
(1, −1), (1, 1), (−1, 1), figure 3.4. The linear interpolation functions are listed as follows:
3.2 Operator splitting and time discretisation 87

triangle square
ψ1 = (1 − ξ − η) ψ1 = (1 − ξ − η + ξη)/4
ψ2 = ξ ψ2 = (1 + ξ − η − ξη)/4 (3.13)
ψ3 = η ψ3 = (1 + ξ + η + ξη)/4
ψ4 = (1 − ξ + η − ξη)/4
For further details the reference [73] is recommended.

3.2 Operator splitting and time discretisation


The numerical algorithm presented here is based on the operator splitting technique. The
main idea is based on splitting the operators in the governing equations into parts of
clearly distinguishable mathematical properties. The resulting equation parts are treated
separately and subsequently in stages using the most appropriate and advantageous
methods according to the types of the obtained split operators.
Let us consider a partial differential equation for a variable f in the form analogical to
(2.22-2.26):
∂f
+ Af = 0 (3.14)
∂t
where A is a differential operator, which could be split linearly into a few parts, e.g.
A = A1 +A2 . In the particular case of equations (2.22-2.26), A1 represents the advection
and A2 the diffusion operator.
In contrast to the space discretisation, the time discretisation in Telemac3D follows
the finite difference method. The time derivatives of model variables f are approximated
by one-step forward differences. These differences are split into a sum of contributions
due to the subsequent application of the operators A1 and A2 :
∂f (x, y, z, t) f n+1 − f n f n+1 − f˜ f˜ − f n
= = + (3.15)
∂t ∆t ∆t ∆t
where f n+1 , f n represent model variable values at times tn+1 and tn , the time-step
∆t = tn+1 − tn and f˜ is an intermediate (fractional step) solution due to the operator
splitting.
In the following, the equation (3.14) is solved in two steps:
f˜ − f n
+ A1 f = 0 (3.16)
∆t
f n+1 − f˜
+ A2 f = 0 (3.17)
∆t
The time variability of the variable f is assumed to be linear and in the following the
semi-implicit formulation (Crank-Nicolson, [46]) is applied:
f (x, y, z, t) = θf n+1 + (1 − θ)f n (3.18)
88 3.3 Continuity step

which allows weighing of the approximate value of f between the time levels tn+1 and
tn with the Crank-Nicolson implicitness factor θ. By θ = 1 one obtains an implicit, by
θ = 0 an explicit formulation of the time derivative (3.15). One may choose between the
explicit and implicit formulations by changing the value of θ. The explicit formulation
is usually computationally advantageous, but with a time-step value limited by stability
criterion. The implicit formulations are computationally expensive and characterised
by a relatively greater numerical diffusion, but stable where the explicit formulation
does not work [8]. With the value of θ = 0.5 this formulation has an error O(∆t2 ), i.e.
consistency order two, what is an important advantage. For practical reasons the value
of θ is often taken slightly higher than the optimal value of 0.5, between 0.55 and 0.65,
in order to take advantage of the stabilising influence of the implicitness.
The stability of the operator splitting method is achieved, when all the fractional steps of
the resulting algorithm are stable [110]. For entirely explicit formulation the consistency
order is one, so that the accuracy depends linearly on the time step. The main advantage
of the operator splitting technique is that various methods of integration can be applied
in different algorithm parts according to the mathematical nature of the equations.

3.3 Continuity step


3.3.1 Poisson equation solver
The kernel of the non-hydrostatic algorithm consist of the Poisson equation for the
hydrodynamic pressure. The solution of this step of the algorithm gives the hydrody-
namic pressure field, allowing an immediate computation of the final, solenoidal velocity
field. The treatment of this elliptic equation follows clearly all the characteristic steps
of the finite element method (section 2.4.2) and is the best introduction to the partic-
ular techniques implemented in Telemac-3D. Therefore, this step is to be described
first, ignoring the sequence of the algorithm stages (advection, diffusion, continuity, free
surface).

3.3.2 Formulation of the pressure Poisson equation in the finite


element method
In order to simplify the notation, the Poisson equation (2.142) is written for a variable
p(x, y, z) and the right hand side is taken to be a known scalar field R(x, y, z). In the
implemented non-hydrostatic algorithm, p represents the hydrodynamic pressure π, and
R(x, y, z) = ̺0 ∆t−1 ∇ · ũ, section 2.5:
∇2 p(x, y, z) = R(x, y, z) (3.19)

The functions p and R are interpolated using Lagrangian interpolation functions ϕi for
a node i: X
p= pi ϕ i (3.20)
i
3.3 Continuity step 89

The finite element procedure follows in this clearly elliptic equation case the (standard)
Galerkin method: the equation (3.19) multiplied by the test functions (equal here to the
interpolation functions mentioned above) and then integrated in the domain Ω with the
boundary Γ(Ω):

∂2p ∂2p ∂2p


Z Z Z Z
ϕi 2 dΩ + ϕi 2 dΩ + ϕi 2 dΩ = ϕi R dΩ (3.21)
Ω ∂x Ω ∂y Ω ∂z Ω

Integrating the terms of the second order by parts, i.e. using Green’s formula1 one
obtains for the first term on the left side of (3.21):

∂2p ∂p ∂ϕi ∂p
Z Z Z
ϕi dΩ = ϕi nx dΓ − dΩ (3.22)
Ω ∂x2 Γ ∂x Ω ∂x ∂x

and analogical formulations for derivatives according to y and z. (nx , ny , nz ) is the


normal vector to the boundary. Introducing terms in the form (3.22) to (3.21) yields:
!
∂ϕi ∂p ∂ϕi ∂p ∂ϕi ∂p
Z
− + + dΩ (3.23)
Ω ∂x ∂x ∂y ∂y ∂z ∂z
!
∂p ∂p ∂p
Z Z
+ ϕi nx + ny + nz dΓ = ϕi R dΩ
Γ ∂x ∂y ∂z Ω

As a consequence of the integration, the application of linear interpolation functions ϕi in


a second-order problem (diffusion) is possible. The boundary integrals allow imposition
of the boundary conditions (Dirichlet and Neumann). Setting the interpolation formula
(3.20) to (3.23) yields:
!
∂ϕi ∂ϕj ∂ϕi ∂ϕj ∂ϕi ∂ϕj
X Z
pj + + dΩ (3.24)
j Ω ∂x ∂x ∂y ∂y ∂z ∂z
!
∂ϕk ∂ϕk ∂ϕk
X Z X Z
− pk ϕi nx + ϕi ny + ϕi nz dΓ = − Rj ϕi ϕj dΩ
k Γ ∂x ∂y ∂z j Ω

where j are internal and k boundary nodes of the computational domain. In


Telemac3D the flux through the boundary represented by the boundary integral is
approximated in a linear way. At all boundaries:
∂pk ∂pk ∂pk
nx + ny + nz = ak pk + bk (3.25)
∂x ∂y ∂z

whereby the coefficients a and b are defined per boundary node k. In the eventual
code realisation they are separated into surface, bottom and lateral boundary nodes. It
1
Green’s formula Z Z Z
αvi,i dΩ = αvi ni dΓ − α,i vi dΩ
Ω Γ Ω
90 3.3 Continuity step

is a convenient way to impose Neumann boundary conditions, although the boundary


geometry is treated in a simplified way. The normal vector (nx , ny , nz ) must also be
defined per node, which may be a source of errors for complicated boundary shapes.
Taking (3.25) into consideration yields:
!
∂ϕi ∂ϕj ∂ϕi ∂ϕj ∂ϕi ∂ϕj
X Z X Z
pj + + dΩ − pk ak ϕi ϕk dΓ = (3.26)
j Ω ∂x ∂x ∂y ∂y ∂z ∂z k Γ
X Z X Z
− Rj ϕi ϕj dΩ + pk bk ϕk dΓ
j Ω k Γ

For simplification of this equation system, the set of coefficients of the type
Z
gij = ϕi ϕj dΩ (3.27)

is interpreted as a mass matrix G and the set of coefficients


!
∂ϕi ∂ϕj ∂ϕi ∂ϕj ∂ϕi ∂ϕj
Z
cij = + + dΩ (3.28)
Ω ∂x ∂x ∂y ∂y ∂z ∂z

is interpreted as a diffusion matrix C. Following this notation, the linear equation system
to be solved for the unknown vector p = [pj ] can be written as:

Ap = B (3.29)

where A is the C matrix corrected by a known boundary term due to coefficients ak , and
B is equal to −GR, where R = [Rj ], corrected by a boundary term due to the coefficients
bk in (3.26). The imposed (Dirichlet) boundary conditions are taken into consideration
through modification of the global equation system matrix A and the right hand side
vector B, excluding the nodes with imposed, known values from consideration.
It must be mentioned that in the eventual implementation in the code the equation (3.29)
is solved in the σ-mesh, which does not change in time, and taking all advantages from
the geometrical simplifications taking place there (a constant transformation Jacobian,
section 3.1.1). The transformation of the equations and resulting terms and matrices are
thoroughly analogous to the descriptions depicting the diffusion step provided in detail
in section (3.4) and omitted here for the sake of clarity.
Integration of all matrices is carried out analytically, using mathematical expert software.
The resulting linear equation system (3.29) is solved using an effective linear equation
solver (a preconditioned conjugate gradient solver). The verification of the Poisson
equation solver using analytical solutions is given in section 4.2.

3.3.3 Resulting divergence-free velocity field


Having computed the dynamic pressure, the final, divergence-free (solenoidal) velocity
field un+1
j is computed in a straightforward way, using equations (2.221). In this way the
3.4 Diffusion step 91

projection step of the non-hydrostatic algorithm is realised:

∆t ∂π
un+1
j = udj − (3.30)
̺0 ∂xj

The 3D spatial derivatives of the dynamic pressure are computed for each node using the
weak formulation in finite elements. The derivative of a given scalar function f (x, y, z)
is computed locally in the direct surrounding Ω0 of the node i in (patch) the following
way:
! R ∂f PN R ∂ϕj
∂f Ω0 ∂x ϕi dΩ0 j fj Ω0 ∂x ϕi dΩ0
= R = PN R ∂ϕj (3.31)
∂x i Ω0 ϕi dΩ0 j xj Ω0 ∂x ϕi dΩ0

In practice, the computation of (3.31) is performed in a loop over all elements by comput-
ing contributions originating from each element to a node i. Finally, these contributions
are assembled by summing up contributions from all elements in a vectorised loop. This
method of derivative computation has the only disadvantage that the exactness of the
computation can be different at the boundaries than in the domain interior.
Finally, the impermeability boundary conditions (section 2.7.3) are imposed on the re-
sulting velocity field. In practice, it is necessary even when the dynamic pressure field
has been obtained using boundary conditions ensuring impermeability.

3.3.4 Source term of the pressure Poisson equation


As the source term denoted as R(x, y, z) in section 3.3.1 the divergence of the interme-
diate velocity must be taken, multiplied by a factor: R(x, y, z) = ̺0 ∆t−1 ∇ · ũ. The
divergence, being a sum of partial derivatives of velocity components, is computed using
the weak formulation in finite elements (3.31).

3.4 Diffusion step

3.4.1 Diffusion step: formulation in finite elements


In a sharp contrast to the σ-transformation of the advection equation, the diffusion equa-
tion cannot be transformed into an equivalent or simple form in σ-co-ordinates due to the
nonconformity of this transformation [110]. Because the Jacobian of the transformation
from the real to the reference element is not constant, the analytical integration of the
diffusion matrix formulated in the real mesh is not possible. If simplifying approxima-
tions are made, they lead to parasitic numerical diffusion [112]. In order to avoid both
these ambiguities, the diffusion step is formulated directly in the σ-mesh.
The treatment of the diffusion step follows patterns analogous to the Poisson equation
treatment (section 3.3.1). In contrary to the description of the Poisson equation solver,
the derivations in this section take the σ-transformation into consideration immediately.
92 3.4 Diffusion step

A verification of this stage of the algorithm using analytical solutions for Ekman profiles
is given in section 4.1.
The equations to be solved for a variable f (x, y, z, t) (one of the velocity components u,
v, w or a tracer concentration c) in the diffusion (parabolic) step are:

fd − fa
= ∇ · (ν f ∇f ) + qf (3.32)
∆t

where f a are results of the advection step, and f d the searched solution of the diffusion
step. The approximation of f on the right hand side of this equation follows the semi-
implicit (Crank-Nicolson) formulation:

f = θf d + (1 − θ)f n (3.33)

In order to avoid the difficulties by transforming the diffusion equation (3.32) directly,
the σ-transformation is applied after formulating the variational form of this equation
in the real mesh.
The finite element procedure follows the standard Galerkin method, applicable for
parabolic equations. The equation (3.32) is multiplied by the test functions ϕ equal
to the interpolation ones and then integrated in the real domain Ω with the boundary
Γ(Ω):

1 1
Z Z Z Z
ϕi f d dΩ − ϕi f a dΩ = ϕi ∇ · (ν f ∇f ) dΩ + qf dΩ (3.34)
∆t Ω ∆t Ω Ω Ω

It is advantageous to treat the diffusion step in the transformed, σ-mesh Ω∗ (section


3.1.1). Equation (3.34) transforms to:

1 Z ∗ d ∗ 1 Z ∗ a ∗
Z

Z
ϕi f h dΩ − ϕi f h dΩ = ϕi ∇ · (ν f ∇f )h dΩ + qf h dΩ∗ (3.35)
∆t Ω ∗ ∆t Ω ∗ Ω ∗ Ω ∗

where h(x, y, t) is the Jacobian of the σ-transformation (section 3.1.1), equal to the water
depth.
The variable f is interpolated using Lagrangian interpolation functions ϕ∗i for a node i,
in the transformed mesh, i.e.:
fj ϕ∗j
X
f= (3.36)
j

Introducing (3.36) into (3.35) yields:

1 X d 1 X a
Z Z
f ϕ∗i ϕ∗j h dΩ∗ − f ϕ∗i ϕ∗j h dΩ∗ = (3.37)
∆t j j Ω∗ ∆t j j Ω∗
Z Z
ϕ∗i ∇ · (ν f ∇ϕ∗j )h dΩ∗ + ϕ∗i ϕ∗j h dΩ∗
X X
fj qf j
j Ω∗ j Ω∗
3.4 Diffusion step 93

After taking (3.33) into consideration and some algebra we obtain the following set of
equations:
1
 Z Z 
fjd ϕ∗i ϕ∗j h ∗
ϕ∗i ∇ (ν f ∇ϕ∗j )h ∗
X
dΩ + θ · dΩ = (3.38)
j ∆t Ω∗ Ω∗

1
X  Z Z
∆tqf j + fja ϕ∗i ϕ∗j h dΩ∗ + (θ − 1) fjn ϕ∗i ∇ · (ν f ∇ϕ∗j )h dΩ∗
X

j ∆t Ω∗ j Ω∗

In the equation (3.37) the following matrices can be easily distinguished: the mass matrix
M (divided by ∆t):
1 Z
mij = ϕ∗ ϕ∗ h dΩ∗ (3.39)
∆t Ω∗ i j
and the generic diffusion matrix E:
Z
eij = ϕ∗i ∇ · (ν f ∇ϕ∗j )h dΩ∗ = (3.40)
Ω∗
∗ ∗ ∗
" ! ! !#
Z
∂ ∂ϕj ∂ ∂ϕj ∂ ∂ϕj
ϕ∗i νx + νy + νz h dΩ∗
Ω∗ ∂x ∂x ∂y ∂y ∂z ∂z

The computation of the mass matrix M is straightforward, but this is not the case for
the diffusion matrix E. For the sake of clarity, only the diffusion along the x-co-ordinate
will be considered (y analogically):

∂ϕ∗
!

Z
exij = ϕ∗i νx j h dΩ∗ (3.41)
Ω∗ ∂x ∂x

Applying the σ-transformation, the derivative according to the x-co-ordinate transforms


to:
∂f ∂f ∂z∗ ∂f
= + (3.42)
∂x ∂x ∂x ∂z∗
The diffusion term along the x-axis transforms to the following expression:

∂ϕ∗j ∂z∗ ∂ϕ∗j


( " !#
Z

exij = νx hϕ∗i + +
Ω∗ ∂x ∂x ∂x ∂z∗
∗ ∗
" !#)
∂z∗ ∂ ∂ϕ j ∂z∗ ∂ϕ j
hϕ∗i νx + dΩ∗ (3.43)
∂x ∂z∗ ∂x ∂x ∂z∗

where all derivatives are expressed in terms of the transformed co-ordinates (x, y, z∗ ).
The term exij is transformed using Green’s formula2 , i.e. integrated per partes. The first
term in equation 3.43 is integrated according to x, the second according to z∗ . The result
2
Green’s formula Z Z Z
αvi,i dΩ = αvi ni dΓ − α,i vi dΩ
Ω Γ Ω
94 3.4 Diffusion step

is a sum of a volume and a boundary integrals:


∂ϕ∗i ∂z∗ ∂ϕ∗i ∂ϕ∗j ∂z∗ ∂ϕ∗j
Z ! !
exij = dxij + sxij = − hνx + + dΩ∗ (3.44)
Ω∗ ∂x ∂x ∂z∗ ∂x ∂x ∂z∗
∂ϕ∗j ∂z∗ ∂ϕ∗j
! !
Z
∗ ∗ ∂z∗ ∗
+ hϕi νx nx + nz + dΓ∗
Γ ∗ ∂x ∂x ∂x ∂z∗

whereby the j index in the boundary integral describes only the boundary nodes at Γ∗ .
The term h∂z∗ /∂x appearing in the equation above can be evaluated in the following
way: " #
∂z∗ ∂S ∂B
Kx = h = − z∗ + (1 − z∗ ) (3.45)
∂x ∂x ∂x
and computed easily for each node. The volume integral from the equation (3.44),
denoted Dx , takes the following form:
∂ϕ∗ ∂ϕ∗
! !
νx ∂ϕi ∂ϕj
Z
dxij =− h i + Kx h j + Kx dΩ∗ (3.46)
Ω∗ h ∂x ∂z∗ ∂x ∂z∗

The volume integral D (the diffusion matrix) can be computed without greater problems
in the reference element using mathematical software. The Jacobian of the transforma-
tion from the σ-mesh-element to the reference element is constant (see section 3.1.3),
which is a great advantage when formulating the diffusion step in the σ-mesh.
For the diffusion in the y-direction terms analogous to those equation 3.44 appear. In
the z-direction, taking into consideration that ∂z∗ /∂z = 1/h:
Z
νz ∂ϕ∗i ∂ϕ∗j Z
νz ϕ∗j
ezij = − dΩ∗ + ϕ∗i n∗z dΓ∗ (3.47)
Ω∗ h ∂z∗ ∂z∗ Γ∗ h ∂z∗

which also can be integrated without problems.


The global equation set to be solved (3.38) can be written shortly in the matrix notation:

[M + θ(D + S)]f d = M(∆tqf + f a ) + (θ − 1)(D + S)f n (3.48)

whereby D is the diffusion matrix, and S the boundary integral, as stated in equations
(3.44) and (3.47). M is the mass matrix divided by the time step ∆t.

3.4.2 Boundary terms in the diffusion step


Denoting the boundary node numbers as k, the boundary integrals in all three directions
can be written as follows:
∂ϕ∗k ∂ϕ∗k
!
Z
νx
sik = ϕ∗i (hn∗x
+ h + Kx dΓ∗
Kx n∗z ) (3.49)
Γ∗ h ∂x ∂z∗

∂ϕ∗k νz ϕ∗
!
∗ νy ∂ϕk
Z Z
∗ ∗
ϕi (hny + Ky nz ) h + Ky dΓ∗ + ϕ∗i n∗z j dΓ∗
Γ∗ h ∂y ∂z∗ Γ∗ h ∂z∗
3.4 Diffusion step 95

In order to understand the physical meaning of (3.49) the flux of the variable f through
a real surface Γ oriented by a normal vector n = (nx , ny , nz ) is considered:

∂f ∂f ∂f
Φf = (ν∇f ) · n = νx nx + νy ny + νz nz (3.50)
∂x ∂y ∂z

Expressing this flux in the transformed co-ordinates one obtains:


! !
νx
f ∂f ∂f νy ∂f ∂f νz ∂f
Φ = h + Kx nx + h + Ky ny + nz (3.51)
h ∂x ∂z∗ h ∂y ∂z∗ h ∂z∗

In this stage the contributions to the boundary integral (3.49) generated at the bottom,
surface and lateral boundaries are discussed separately.
E.g. at the bottom (z∗ = 0), where n∗ = (n∗x , n∗y , n∗z ) = (0, 0, −1) (in the σ-mesh the
bottom is transformed into a horizontal surface) and Kx = −∂B/∂x = nx /nz :

∂ϕ∗k ∂ϕ∗k ∂ϕ∗ ∂ϕ∗ νz ∂ϕ∗k


" ! ! #
νx nx νy ny
Z
sB
ik = −ϕ∗i h + K x − h k + Ky k − dΓ∗B =
Γ∗B h nz ∂x ∂z∗ h nz ∂y ∂z∗ h ∂z∗
1
Z
− ∗ ϕ∗i Φf dΓ∗B (3.52)
ΓB nz

At the surface (z∗ = 1), n∗ = (0, 0, 1) a similar expression is obtained by setting Kx =


∂S/∂x = nx /nz :
1
Z
sik = + ∗ ϕ∗i Φfk dΓ∗S
S
(3.53)
ΓS nz

At the lateral boundaries (z∗ ∈ [0, 1]) n∗ = (n∗x , n∗y , 0)


Z
sLik =+ hϕ∗i Φfk dΓ∗L (3.54)
Γ∗L

For the terms at the bottom and the free surface the transformation to a 2D reference ele-
ment is very simple, the Jacobian being the given triangle surface in the σ-transformation
(or the projection of the real triangular surface s of the prismatic 3D-element at the hor-
izontal surface, s∗ = s/|nz |):

1 f s∗ f
Z Z Z
sSij = ϕ∗i Φ dΓ∗S = ϕref
i Φk dΓref
S = ϕi Φfk dΓS (3.55)
Γ∗S nz k Γref
S
nz ΓS

Actually, when computing the boundary terms there is no special advantage from doing
it in the σ-mesh. The situation is the same for the lateral boundaries (quadrilaterals).
96 3.4 Diffusion step

3.4.3 Boundary conditions for the diffusion step


The imposed boundary conditions can be taken into consideration in a very similar way
to that described in section 3.3.1. The boundary fluxes, i.e. the Neumann boundary
conditions are defined in the following way:
∂fk ∂fk k ∂fk k ∂fk k
(ν∇f ) · n = νi = νx nx + νy ny + νz n = ak fk + bk (3.56)
∂nk ∂x ∂y ∂z z

whereby the coefficients a and b are defined per boundary node k. For the determination
of the numeric values of ak and bk in the eventual application, the horizontal viscosity
(diffusivity) at the vertical lateral boundaries and the vertical one at the surface or at
the bottom are used for νi . The computation of the boundary integrals is performed
separately for the bottom, surface and lateral boundaries. The Neumann boundary
conditions are taken into consideration through the boundary integrals influencing the
linear system matrix (coefficients ak ) or the right hand side vector (coefficients bk ), in
an analogous way to section 3.3.2. If ∂fk /∂nk = 0, ak and bk are simply set to 0 and
boundary terms disappear.
The Dirichlet boundary conditions are realised in this step through modification of the
global equation system matrix and the right hand side vector, eventually excluding the
nodes with imposed, known values from the consideration.

3.4.4 Computation of the source terms


A characteristic feature of the non-hydrostatic model in comparison with its hydrostatic
complement is not only its three-dimensionality, but also the treatment of the forcing
terms (i.e. density differences, free surface gradients and Coriolis force) in the three-
dimensional diffusion step of the algorithm (see section 2.8.5). In the hydrostatic model
they are taken into consideration in the pressure – continuity – free surface step (section
2.3.3). In the non-hydrostatic one, they appear as a sum of terms forming the source
term qf in the equation (3.32).

3.4.4.1 Free surface gradient terms

The terms representing the barotropic pressure gradients given by (2.217) are two-
dimensional and independent of depth. The spatial derivatives of the free surface posi-
tion are computed for each node using the weak formulation in finite elements (3.31). In
this formulation, the spatial derivatives of the free surface position in the previous step
S(x, y, t) are computed locally in the 2D surrounding Ω0 of the node i (node patch) in
the following way:
∂S 2D PN ∂ϕ2D
ϕ2D
! R
fj j
dΩ0
R
∂S Ω0 ∂x ϕi dΩ0 j ∂x i
Ω0
= R 2D
= R ∂ϕ2D (3.57)
∂x Ω0 ϕ i dΩ0 PN j 2D
i j xj Ω0 ∂x ϕi dΩ0

where ϕ2D
i are the 2D interpolation functions.
3.4 Diffusion step 97

3.4.4.2 Buoyancy terms

The buoyancy terms, i.e. the accelerations which arise due to density differences, defined
by eq. (2.218) can be computed in two different ways. To obtain these terms, horizon-
tal derivatives of a vertically integrated relative density difference field ∆̺/̺0 must be
computed:
∂ R S ∆̺n
 
−g ∂x z ̺0
dz
∂ R S ∆̺n
 
F∆̺ =  −g ∂y z ̺0 dz 
 (3.58)
0

The first method, which seems to be natural, is to compute the relative density gradients
using the weak formulation in finite elements. The relative density field is integrated
first, and then the derivatives are computed according to the formula (3.31). However,
an analysis of this method shows that it causes spurious numerical density currents over
a non-horizontal bottom, even when the hydrostatic consistency condition is fulfilled
[110, 112] (see section 3.1.2). The clue to this problem is the succession of the integration
and differentiation of a variable interpolated by linear base functions. When the relative
density is being integrated, terms of the second order appear. Therefore, the integrals
appearing in the weak derivative formulation (3.31) must be also computed with an
exactness of the second order in the whole domain. For the formula (3.31) with linear
interpolation functions this is not the case at the boundaries. For a non-horizontal
bottom, errors appear even for stable density fields with a linear vertical profile [110].
Another formulation of the source term resulting from the density gradients must be
applied in this case. The idea is to transform the density term into an expression that
can be differentiated without problems using a first-order procedure. Naturally, this
is the case when the term is differentiated first and the result integrated next. The
integration follows the trapezium rule (NL is the number of mesh levels/planes and j is
the plane number corresponding to z value):

∂ S ∆̺n ∂ NXL −1 zi+1 ∆̺n


Z Z
Fx∆̺ = −g dz = −g dz =
∂x z ̺0 ∂x i=j zi ̺0
NX
L −1
∆̺ni+1 ∆̺ni
" !#
∂ 1
−g (zi+1 − zi ) + = (3.59)
∂x i=j 2 ̺0 ̺0
1 NX
L −1
∆̺ni+1 ∆̺ni ∆̺ni+1 ∆̺ni
" ! !#
∂ ∂
− g (zi+1 − zi ) + + (zi+1 − zi ) +
2 i=j ∂x ̺0 ̺0 ∂x ̺0 ̺0

In this way only the terms of the first order are differentiated and the weak derivative
is accurate enough in the whole domain. This formulation is exact for the linearly
changing density fields. Due to the obvious advantages of the second formulation, it is
implemented in the model. However, for non-linear density fields, errors can still appear.
In order to deal with this problem, higher order interpolation functions would have to
be introduced.
98 3.5 Advection step

3.4.4.3 Coriolis terms

The computation of the Coriolis terms defined by equation (2.216) is performed without
any difficulties using the velocity component values from the previous step:

−fH w n + fV v n
 

FCor =

−fV un   (3.60)
n
fH u

3.4.4.4 Other forcing

The term qf in equation (2.215) can also include all other forcing accelerations, for
example those appearing in non-inertial systems (e.g. sloshing of a fluid in tanks) or a
prescribed pressure gradient, etc.

3.5 Advection step

3.5.1 Introduction
The advection step of the new algorithm is identical with the previous version. It is
presented here for completeness of the description. It should be noted that virtually the
same methods (characteristics and SUPG), but in two dimensions, are applied to solve
the equations for the free surface and are tested and discussed in section 3.6.
The non-linear advection (or hyperbolic) terms in the equations (2.22-2.26):

fa − fn
+ u · ∇f = 0 (3.61)
∆t

can be treated using the method of characteristics or the streamline upwind Petrov-
Galerkin (SUPG) method. The first is realised in the σ-mesh, where a transformed
velocity field is advected in the transformed velocity field, the latter in the real mesh,
but taking advantages of the σ-transformation when computing various matrices.

3.5.2 Advection with the method of characteristics


The description of the treatment of the advection (hyperbolic) step with the method
of characteristics is given in [110]. It is presented here shortly for completeness. It
illustrates well the philosophy of the operator splitting technique – treating the operators
appearing in the hydrodynamic equations separately with most appropriate methods. In
this particular case it is the Lagrangian approach to the advection.
The σ-transformation does not change the form of the advection equation (section 3.1.2).
Therefore, in this step the transformed velocity and tracer concentration values are
3.5 Advection step 99

advected in the transformed velocity field u∗ = (u, v, w∗), according to equation (3.3) in
the time-independent σ-mesh of a simple geometry.
The non-linear advection terms in the equations (2.22-2.26):
fa − fn fa − fn ∂f ∂f ∂f
+ u∗ · ∇f = +u +v + w∗ =0 (3.62)
∆t ∆t ∂x ∂y ∂z∗
are treated using the Lagrange method. In this method the time variability of a variable
f is observed along a streamline:
df ∂f
= + u∗ · ∇f (3.63)
dt ∂t
The equation (3.62) states also that the variable f does not change along the streamline
(characteristic). In order to discretise the Lagrange time derivative (3.63), the beginning
(the base) at time tn and the end of the streamline at time tn+1 must be known. In this
method as the end points of the streamlines (characteristics) at time level tn+1 the mesh
nodes are taken. The base points xbi at time level tn are found following the streamlines
backward in time using the differential equations for the streamlines:
dxi
= ui (3.64)
dt
Integration of (3.64) between the time levels tn and tn+1 yields
xbi = xn+1
i − ui∆t (3.65)

which is equivalent to the Euler explicit method (or 1st order Runge-Kutta method).
Having found the base points, it remains to interpolate the function values fk from
the surrounding nodes of elements in which the base points have been found. This
interpolated value is equal to the value of f at the node xi at time level tn+1 , because f
is constant along the streamline.
At boundaries, a few extraordinary situations are met. When a characteristic curve,
followed backwards in time, leaves the computational domain at an open boundary, as
the base point its intersection with the boundary is assumed. When an impermeable
boundary is hit, the streamline is followed further along the boundary in the direction
of the velocity component parallel to the given boundary section. In this step of the
algorithm Dirichlet boundary conditions are also taken into consideration – the values
at given nodes are set to the imposed values.
The main advantage of this method is a natural treatment of the non-linear terms of
the advection equation (Lagrange method) and theoretically no limits on the time-step
according to the Courant number. The method of characteristics is namely the only
explicit method that is unconditionally stable according to the chosen time-step. The
main disadvantage is the fact that for the interpolation of the function values at the
position of the found base points, the linear base functions in prismatic elements are
used. Therefore serious damping (numerical diffusion) occurs due to interpolation. Using
interpolation functions of higher order may improve the exactness of this step. Another
disadvantage are problems with mass conservation using this scheme [110].
100 3.5 Advection step

3.5.3 Advection with SUPG method


The advection step in which the hyperbolic advection equation (2.214) is to be solved,
can also be treated using the Streamline Upwind Petrov-Galerkin (SUPG) method in
finite elements [16, 116]. The idea of this method is shown for the advection of the v
velocity component and, for simplicity, in the real mesh. In contrast to the method
of characteristics, the advection equation is formulated in the real co-ordinates. The
σ-transformation appears in the stage, when the resulting matrices are treated.
Using the explicit formulation for the time partial derivative in the real mesh as in
(2.214), one obtains:
va − vn ∂v ∂v ∂v
+u +v +w =0 (3.66)
∆t ∂x ∂y ∂z
v can be weighted between the levels n and a using the implicitness parameter θ:

v = θv a + (1 − θ)v n (3.67)

Therefore, for θ = 1 a fully implicit, and for θ = 0 a fully explicit advection schema is
obtained.
v(x, y, z, t) is interpolated using Lagrangian interpolation functions ϕj for a node j, i.e.:
X
v= vj ϕj (3.68)
j

The finite element procedure follows the SUPG method. The equation (3.66) is multi-
plied by the test functions ψi and then integrated in the domain Ω:
1 1 ∂v ∂v ∂v
Z Z Z Z Z
a n
ψi v dΩ− ψi v dΩ+ ψi u dΩ+ ψi v dΩ+ ψi w dΩ = 0 (3.69)
∆t Ω ∆t Ω Ω ∂x Ω ∂y Ω ∂z

Introducing (3.68) to 3.69 yields:


1 Z X Z ∂ϕj X Z ∂ϕj
vja
X
ψi ϕj dΩ + vj ψi u dΩ + vj ψi v dΩ + (3.70)
j ∆t Ω j Ω ∂x j Ω ∂y
∂ϕj 1
Z Z
vjn
X X
vj ψi w dΩ = ψi ϕj dΩ
j Ω ∂z j ∆t Ω

or shortly:
1 1
 Z  Z   Z 
vja vjn
X X X
ψi ϕj dΩ + vj ψi u · ∇ϕj dΩ = ψi ϕj dΩ (3.71)
j ∆t Ω j Ω j ∆t Ω

where u = (u, v, w). In the Upwind Streamline Petrov-Galerkin Method the test func-
tions ψi in the advection terms of (3.71) are replaced by:
u
ψi = ϕi + α ∇ϕi (3.72)
|u|
3.5 Advection step 101

where α is a real constant (upwind parameter). In this way for the advective mass
transport the upstream nodes are stronger weighted than the downstream ones. In all
other terms ψi = ϕi is set, as in the standard Galerkin method. Realising this idea in
(3.71) yields:
1 Z
  Z 
vja
X X
ϕi ϕj dΩ + vj ϕi u · ∇ϕj dΩ
j ∆t Ω j Ω
!
u 1 Z
Z  
vjn
X X
+α vj ∇ϕi u · ∇ϕj dΩ = ϕi ϕj dΩ (3.73)
j Ω |u| j ∆t Ω

Introduction of (3.67) into (3.73) yields:


" #
1 Z Z Z
u
vja
X
ϕi ϕj dΩ + θ ϕi u · ∇ϕj dΩ + θα ∇ϕi u · ∇ϕj dΩ (3.74)
j ∆t Ω Ω Ω |u|
" #
1 Z Z Z
u
vjn
X
= ϕi ϕj dΩ + (θ − 1) ϕi u · ∇ϕj dΩ + (θ − 1)α ∇H ϕi u · ∇ϕj dΩ
j ∆t Ω Ω Ω |u|

In order to solve the linear equation system (3.74) the following matrices of coefficients
must be computed:
Mass matrix M (divided by ∆t):
1
Z
mij = ϕi ϕj dΩ (3.75)
∆t Ω

Advection matrix K: Z
kij = ϕi u · ∇ϕj dΩ (3.76)

SUPG matrix U:
u
Z
uij = α ∇ϕi u · ∇ϕj dΩ (3.77)
Ω |u|
And, by implicit treatment of the advection terms the following set of linear equations
must be solved in order to find the unknown vector va :

[M + θ(K + U)]va = [M + (θ − 1)(K + U)]vn (3.78)

The computation of the matrices can be performed in the σ-mesh exactly, therefore the
integrals appearing in the mass, advection and SUPG matrix are transformed into σ-
mesh (like in the diffusion step, section 3.4) and using the definition of w∗ (equation
3.4).
1
Z
mij = ϕ∗i ϕ∗j h dΩ∗ (3.79)
∆t Ω ∗
Z
kij = ϕ∗i u∗ · ∇ϕ∗j h dΩ∗ (3.80)
Ω∗
u∗
Z
uij = α ∇ϕ∗i u∗ · ∇ϕ∗j h dΩ∗ (3.81)
Ω∗ |u∗ |
102 3.6 Free surface step

where u∗ is the transformed velocity (u, v, w∗). Eventually, in contrast to the imple-
mented method of characteristics, the nodal values are the real velocity values, but the
advection field is the transformed velocity u∗ . The integration of the matrices is carried
out analytically. By the eventual implementation in the fully three-dimensional code, a
constant value α = ∆t/2 is assumed for the upwind parameter (see also section 3.6.2).
The Dirichlet boundary conditions are taken into consideration by modification of the
global equation set (3.78).
As it can be seen after the description of the diffusion step, which leads to solution of a
linear set of equations like (3.78) as well, both steps can be treated simultaneously. A
common set of linear equations for advection and diffusion, resulting from addition of
appropriate matrices and vectors is solved for each velocity component and tracer.
Comparing (3.78) and (3.48) the equation system in the coupled advection-diffusion step
for a resulting variable vd reads:

{M + θ[(D + S) + (K + U)]} vd =
M(∆t qf ) + {[M + (θ − 1)[(D + S) + (K + U)]} vn (3.82)

3.5.4 Boundary conditions for advection step


If the advection step is realised with the method of characteristics, only Dirichlet bound-
ary conditions are imposed on the resulting field f a . If the SUPG method is applied, the
boundary conditions are imposed as described in section 3.4.

3.6 Free surface step

3.6.1 Free surface with the method of characteristics


The solution method for the kinematic boundary condition (2.158) follows exactly the
description in section 3.5.2 for the advection equation, but in two dimensions. For inter-
polating the free surface position at the bases of the characteristics the two-dimensional
linear interpolation in triangles are applied (section 3.1.4).

3.6.2 SUPG formulation for the kinematic boundary condition


The kinematic boundary condition (2.158) is a hyperbolic partial differential equation
which can be solved in the finite element method using the Streamline Upwind Petrov-
Galerkin (SUPG) method [16, 116]. The kinematic boundary condition can be treated
as a two-dimensional advection equation with a source term:

∂S ∂S ∂S
+u +v =w (3.83)
∂t ∂x ∂y
3.6 Free surface step 103

Using the explicit formulation for the time partial derivative in (3.83) one obtains:
S n+1 − S n ∂S ∂S
+u +v =w (3.84)
∆t ∂x ∂y

S can be weighted between the time level n and n + 1 using the implicitness parameter
θ:
S = θS n+1 + (1 − θ)S n (3.85)
For θ = 1 a fully implicit, and for θ = 0 a fully explicit schema is obtained. The function
S(x, y, t) is interpolated using 2D Lagrangian interpolation functions ϕi for a node i, i.e.:
X
S= Sj ϕ j (3.86)
j

As in section 3.5.3, the finite element formulation follows the SUPG method. The equa-
tion (3.84) is multiplied by the test functions ψ and then integrated in the domain Ω:
1 1 ∂S ∂S
Z Z Z Z Z
ψi S n+1 dΩ − ψi S n dΩ + ψi u dΩ + ψi v dΩ = ψi w dΩ (3.87)
∆t Ω ∆t Ω Ω ∂x Ω ∂y Ω

Introducing (3.86) to 3.87 one obtains:


1 ∂ϕj ∂ϕj
Z Z Z
Sjn+1
X X X
ψi ϕj dΩ + Sj ψi u dΩ + Sj ψi v dΩ = (3.88)
j ∆t Ω j Ω ∂x j Ω ∂y
1
Z Z
Sjn
X X
wj ψi ϕj dΩ + ψi ϕj dΩ
j Ω j ∆t Ω

or:
1
 Z  Z 
Sjn+1
X X
ψi ϕj dΩ + Sj ψi uH · ∇H ϕj dΩ = (3.89)
j ∆t Ω j Ω

1 1
 Z   Z 
Sjn
X X
∆twj ψi ϕj dΩ + ψi ϕj dΩ
j ∆t Ω j ∆t Ω

The SUPG method the test functions in the advective terms are to be replaced by:
u
ψi = ϕi + α ∇ϕi (3.90)
|u|

where α is a real constant and ϕi interpolation functions. In all other terms ψi = ϕi is


set. Introducing so defined test functions to (3.89) yields:
1 Z
 
Sjn+1
X
ϕi ϕj dΩ + (3.91)
j ∆t Ω
!
uH
X Z  X Z
Sj ϕi uH · ∇H ϕj dΩ + α Sj ∇H ϕi uH · ∇H ϕj dΩ =
j Ω j Ω |uH |
1 1
 Z   Z 
Sjn
X X
∆twj ϕi ϕj dΩ + ϕi ϕj dΩ
j ∆t Ω j ∆t Ω
104 3.6 Free surface step

Introduction of (3.85) into (3.91) yields:


" #
1 uH
Z Z Z
Sjn+1
X
ϕi ϕj dΩ + θ ϕi uH · ∇H ϕj dΩ + θα ∇H ϕi uH · ∇H ϕj dΩ
j ∆t Ω Ω Ω |uH |
1
X  Z 
= ∆t wj ϕi ϕj dΩ +
j ∆t Ω
" #
1 uH
Z Z Z
Sjn
X
ϕi ϕj dΩ + (θ − 1) ϕi uH · ∇H ϕj dΩ + (θ − 1)α ∇H ϕi uH · ∇H ϕj dΩ
j ∆t Ω Ω Ω |uH |

In order to solve the linear equation system (3.92) the following matrices of coefficients
must be computed:
mass matrix M divided by ∆t:
1
Z
mij = ϕi ϕj dΩ (3.92)
∆t Ω

advection matrix K: Z
kij = ϕi uH · ∇H ϕj dΩ (3.93)

SUPG matrix U:
uH
Z
uij = α ∇H ϕi uH · ∇H ϕj dΩ (3.94)
Ω |uH |
Finally, the following set of linear equations must be solved in order to find the unknown
vector Sn+1 = [sn+1
ij ]:
ASn+1 = B (3.95)

where A = M + θ(K + U) and the vector B = M(Sn + ∆tW) + (θ − 1)(K + U)Sn ,


where W = [wi ].
The Dirichlet boundary conditions are taken into consideration by modification of the
global equation set (3.95).
The value of upwind parameter α influences strongly the properties of the SUPG algo-
rithm. A constant value of α = ∆t/2 and the horizontal velocity vector itself uH is taken
instead of the unit vector in the direction of the horizontal velocity uH /|uH |. This leads
to the following form of the SUPG matrix (3.94):
∆t Z
uij = uH ∇H ϕi uH · ∇H ϕj dΩ (3.96)
2 Ω

A classical SUPG scheme is also available, where instead of (3.90) the following expression
is assumed
ψi = ϕi + K∇ϕi (3.97)

The vector K has the following components:


!
∆x u ∆y v
K= , (3.98)
2||u|| 2||u||
3.6 Free surface step 105

The velocity norm is computed as the mean velocity in an element, ||u|| = (u2 + v 2 )1/2
and the ∆x and ∆y are the maximum extents of an element along the x and y-axis,
respectively. The tests of the 2D-advection equation solvers are given in section 4.3.1.

3.6.2.1 Implementation of the absorbing boundary conditions

In order to implement the absorbing boundary conditions for the free surface (section
2.7.9), a dissipative term is introduced to the kinematic boundary condition for the free
surface. Equation (3.83) transforms to:
! !
∂S ∂S ∂S ∂ ∂S ∂ ∂S
+ us + vs = ws + νx + νy (3.99)
∂t ∂x ∂y ∂x ∂x ∂y ∂y

which is equivalent to a transport equation for S with a source term w. The distribution
of the values of νx and νy define the properties of damping zones. This advection-diffusion
equation is treated in 2D in the same way as in the 3D case described in section 3.5.3.
The resulting equation system can be sketched as follows:

S n+1 h i
+ θ u · ∇S n+1 − ∇ · (ν∇S n+1 ) = (3.100)
∆t
Sn
+ ws + (θ − 1) [u · ∇S n − ∇ · (ν∇S n )]
∆t

3.6.3 Explicit formulation for the conservative free surface


equation
In order to solve the conservative free surface equation (2.159) explicitly for a 2D node
i, one may write:

Sin+1 − Sin ∂(hni un+1 ) ∂(hni vn+1


!
∂ Z Sn ∂ Z Sn i i )
=− udz − vdz = − + (3.101)
∆t ∂x −B ∂y −B ∂x ∂y

where (un+1 , vn+1 ) are vertically integrated horizontal final velocity components (using
the free surface elevation from the previous step S n ). The new free surface S n+1 is
computed explicitly for each node in finite elements, i.e. without solving any equation
system. The derivatives of vertically integrated velocity components are computed for
each node using the weak formulation, equation (3.31).
Because only horizontal gradients of the integrated velocities are needed, for simplifi-
cation and in order to spare computation time two-dimensional interpolation functions
ϕ2D
i (for a 2D-mesh of triangles) are used, calling for both numerator and denominator
of (3.31) a subroutine computing the vector ai :
N Z
∂ϕ2D
j
ϕ2D dΩ0
X
ai = fj (3.102)
j Ω0 ∂x i
106 3.6 Free surface step

A version of this free surface computing method, basing on computing the derivative
of the velocity integral locally in each 3D-element can be thought of as in (3.59). The
integration in depth follows the trapezium rule (NL is the number of mesh levels/planes),
e.g. for the term containing u velocity component:

∂ ZS ∂ NXL −1 Z zi+1

udz = udz =
∂x −B ∂x i=1 zi
∂ NX
L −1
1
 
(zi+1 − zi )(uni+1 + uni ) = (3.103)
∂x i=1 2
1 NX
L −1
" #
∂ ∂
(zi+1 − zi )(uni+1 + uni ) + (zi+1 − zi ) (uni+1 + uni )
2 i=1 ∂x ∂x

3.6.4 Semi-implicit formulation for the conservative free sur-


face equation

In the semi-implicit formulation for the conservative free surface equation (2.159), the
velocity components at the time level n + 1, as well as the position of the free surface at
the time level n and n + 1 are involved. The horizontal velocity components at the time
level n + 1 can be easily integrated between the bottom B and the free surface elevation
S n yielding (un+1 , v n+1 ). The scheme can be formulated as follows (using water depth
h = S − B as the searched variable):

hn+1 − hn ∂(hun+1 ) ∂(hv n+1 )


!
=− +
∆t ∂x ∂y
n+1 n+1
!
∂(u ) ∂(v ) ∂h n+1 ∂h n+1
= −h + − u − v (3.104)
∂x ∂y ∂x ∂y

where the water depth in the terms on the right hand side of this equation can be weighted
between the time levels n and n + 1 using an implicitness factor θ: h = θhn+1 + (1 − θ)hn .
The equation (3.104) is treated using SUPG FEM in exactly the same way as presented
in section 3.6.2 (with the difference that the advection velocity is the mean one in depth),
leading to a global matrix equation in the form:

Ahn+1 = B (3.105)

where A = M + θ(K + U + Q) and the vector B = [M + (θ − 1)(K + U + Q)]hn . The


matrix Q contains the divergence of the mean horizontal velocity:

∂u ∂v
Z Z Z
qij = ∇H · uH ϕi ϕj dΩ = ϕi ϕj dΩ + ϕi ϕj dΩ (3.106)
Ω Ω ∂x Ω ∂y

The equation (3.105) is treated in exactly the same way as described in section 3.6.2.
3.6 Free surface step 107

3.6.5 Free surface stabilisation


The stability of the free surface algorithm is crucial for the stability of the global solution,
because errors in computing the free surface gradients in FEM (barotropic part of hydro-
static pressure gradients) influence negatively and immediately the whole domain. The
computation of the free surface based on the kinematic boundary condition, a hyperbolic
equation, is over-sensitive for small errors or imperfectly set boundary conditions. Great
care must be exercised in harmonising the boundary conditions for velocity, pressure and
free surface. Due to its hiperbolic nature, the kinematic boundary condition does not
include any dissipative terms which may smoothen the solution. Situations where some
instabilities of the free surface occur are common in practice. When boundary conditions
for the free surface are only approximately known, or awkwardly set, or a wave frequency
dispersion occurs for waves with a length comparable with the spatial resolution, some
stabilising techniques for the free surface are needed.
The disadvantage of stabilising techniques is an artificial, non-physical smoothing of
the free surface. The advantage is that the model can be applied in more complicated
situations regarding properly set boundary conditions.
In this work, a well-known filtering method in FEM, using the mass matrix M (equation
3.27) is applied [73]. In this technique, a vector V is modified according to the formula:

MV
V′ = (3.107)
ML

where ML is a diagonal matrix obtained by mass-lumping of M, i.e. by obtaining its


diagonal elements by totalling the terms in each row of M. The integral of the vector
V is preserved, which in the case of the free surface as the filtered variable guarantees
conservation of mass. Alternatively, the advection velocity in the free surface algorithms
can be also filtered.
108 3.6 Free surface step
Chapter 4

Verification of algorithm stages

Only a fool has no doubts.


Karel Čapek

Due to the modular structure of the developed algorithm, it is possible and appropriate to
test and verify its separate parts/stages. In this chapter the verification and performance
tests of the diffusion step, Poisson equation solver and free surface step (2D advection
equation) are presented. The theory is provided in chapter 3.

4.1 Diffusion step test: Ekman profiles


A particularly good example for testing the diffusion step is the reproduction of Ekman
spirals. According to the theory of V.W. Ekman, characteristic velocity profiles appear
in the stationary oceanic bottom and surface layers of frictional influence. In the deep
ocean, these layers are relatively thin compared to the much thicker intermediate layer in
mid-depths, where the current is not influenced by the bottom or wind shear, respectively.
In this ocean interior, according to Ekman’s theory, the current is stationary and in
geostrophic equilibrium. It means that the pressure gradient and Coriolis force cancel
each other out in this layer.
The spiral profiles appear due to the disturbance of the geostrophic equilibrium just below
the surface by the wind shear (drift current) and just above the ocean floor by the bottom
shear (slope current). Expressing it mathematically, they result from specific boundary
conditions imposed on the momentum equations in layers of a defined thickness. The
analytical solutions for these profiles were obtained for the stationary case and constant
eddy viscosities by V.W. Ekman in 1905 [41].
The solution for the drift current is determined through a Neumann wind shear boundary
condition at the surface. The solution for the slope current is determined by the driving
pressure gradient, i.e. a source term in the equation, and the Dirichlet no-slip boundary
condition at the bottom. Therefore, these examples are very well suited for verification
of the numeric diffusion step implementation, treated as a separate program unit.

109
110 4.1 Diffusion step test: Ekman profiles

Short mathematical derivations of the solutions for the drift and slope currents are given
in the following sections 4.1.1 and 4.1.2.

4.1.1 Slope current


Let us assume that a constant horizontal pressure gradient exists in the ocean. It can
result from an external atmospheric pressure field or from a constant free surface slope
prevailing for a long enough period of time. A stationary current in the interior of
the ocean appears, when the pressure gradient force is balanced by the Coriolis force
(geostrophic equilibrium).
For simplicity, the co-ordinate system is so oriented, that ∇H p = (0, ∂p/∂y). The z-
axis is oriented upwards, with z = 0 at the surface and z = −H at the bottom, H
being the water depth. Assuming viscosity νz and the vertical Coriolis coefficient f
to be constant for a stationary current in hydrostatic approximation, the momentum
conservation equations reduce to:
∂2u ∂2v 1 ∂p ∂p
f v + νz = 0, −f u + νz = , = ̺g (4.1)
∂z 2 ∂z 2 ̺ ∂y ∂z

with the following boundary conditions at the surface and the bottom:
∂u
τx = −νz z=0 (4.2)
∂z
∂v
τy = −νz z=0
∂z
u=v=0 z = −H

Defining (uE , vE ) as the velocity field describing the deviation from the geostrophic
equilibrium current (ug ) in the bottom layer, one obtains u = ug + uE and v = vE . The
velocity ug satisfies the geostrophic equation for the given constant pressure gradient:
1 ∂p
−f ug − =0 (4.3)
̺ ∂y

Introducing a complex velocity V = uE + ivE into (4.1) and (4.2) and adding equations
(4.1) one obtains the following problem:
d2 V if
2
− V =0
dz νz
dV
=0 z=0 (4.4)
dz
V = −ug z = −H

The global solution of the problem (4.4) has a form V = Aeβz + Be−βz . Setting this
form into the equation (4.4) in order to obtain β yields:
β = (1 + i)(f /2νz )1/2 = (1 + i)γ (4.5)
4.1 Diffusion step test: Ekman profiles 111

where γ ∈ R. Taking the boundary conditions in (4.4) into consideration in order to


find A, B one obtains a complex solution:

cosh[(1 + i)γz]
V = −ug (4.6)
cosh [(1 + i)γH]

which can be transformed into the solutions for real parts of the velocity components
(u, v):
!
cosh[γ(H + z)] cos[γ(H − z)] + cosh[γ(H − z)] cos[γ(H + z)]
u = ug 1 − (4.7)
cosh(2γH) + cos(2γH)
!
sinh[γ(H + z)] sin[γ(H − z)] + sinh[γ(H − z)] sin[γ(H + z)]
v = ug
cosh(2γH) + cos(2γH)

For the deep ocean, an assumption of infinite depth is often appropriate. For simplicity
of the solution, the z-axis is also oriented upwards, but this time z = 0 is taken at the
bottom, and z → ∞ towards the surface:

d2 V if
2
− V =0
dz νz
dV
=0 z→∞ (4.8)
dz
V = −ug z=0

In this case the solution has a much simpler form:

V = −ug e−γz [cos(γz) − i sin(γz)] (4.9)


u = ug [1 − e−γz cos(γz)]
v = ug e−γz sin(γz)

In order to estimate the thickness of the layer of the frictional influence Ekman introduced
a parameter called depth of frictional influence D defined as the distance above the
bottom, where the value of the velocity deviation |ve | = ug e−γz has decreased to e−π =
1/23 of the geostrophic current:
!1/2
π 2νz
D= =π (4.10)
γ f

The vertical structure of the velocity profile depends strongly on the ratio H/D (figure
4.1). For water depths H < D the geostrophic balance is never attained, and the velocity
has a component in the direction of the pressure gradient at every depth.
112 4.1 Diffusion step test: Ekman profiles

4.1.2 Drift current


The Ekman profiles for the wind-driven current in the surface layer of the ocean are
obtained in an analogical way to the profiles in the bottom layer. Introducing the
complex wind shear variable T = τx + iτy , where τx = ̺νz (∂u/∂z) and τy = ̺νz (∂v/∂z)
one obtains the following problem to be solved for finite or infinite depth:
d2 V if
2
− V =0
dz νz
dV
T = ̺νz z=0 (4.11)
dz
V =0 z = −H finite depth
V =0 z→∞ infinite depth

where it is assumed that the z-axis points upwards and z = 0 corresponds to the position
of the surface and z = −H is the flat bottom level or an infinite depth is assumed.
The solution for the infinite depth is:
π
V = V0 eγz ei(γz− 4 ) (4.12)

where
T τwx τwy
V0 = u0 + iv0 = 1/2
= 1/2
+i (4.13)
̺(νz |f |) ̺(νz |f |) ̺(νz |f |)1/2
is the complex velocity at the surface. In order to obtain a simplified form of the
analytical solution, the wind is assumed to blow in the y-direction, i.e., τx = 0, τy 6= 0.
The solution is:
π
V = v0 eγz ei(γz+ 4 )
π
u = v0 eγz cos(γz + ) (4.14)
4
π
v = v0 eγz sin(γz + )
4
From this equation it is clear, that the drift current is deflected 45o to the right of the
wind direction.
The complex solution for the finite depth case is:
sinh[(1 + i)γ(z + H)] π
V = V0 exp(− i) (4.15)
cosh[(1 + i)γH] 4
Lengthy analytical formulae for the real velocity components u and v can be obtained
from (4.15) using mathematical software. Assuming again that the wind blows in the
y-direction, i.e., τx = 0, τy 6= 0, the solution is (ζ = z + H):
√ !
2 [sinh(γζ) cos(γζ) − cosh(γζ) sin(γζ)] cosh(γH) cos(γH)
u = v0 (4.16)
2 cosh2 (γH) cos2 (γH) + sinh2 (γH) sin2 (γH)
√ !
2 [sinh(γζ) cos(γζ) + cosh(γζ) sin(γζ)] sinh(γH) sin(γH)
+v0
2 cosh2 (γH) cos2 (γH) + sinh2 (γH) sin2 (γH)
4.1 Diffusion step test: Ekman profiles 113

√ !
2 [sinh(γζ) cos(γζ) + cosh(γζ) sin(γζ)] cosh(γH) cos(γH)
v = v0 (4.17)
2 cosh2 (γH) cos2 (γH) + sinh2 (γH) sin2 (γH)
√ !
2 [sinh(γζ) cos(γζ) − cosh(γζ) sin(γζ)] sinh(γH) sin(γH)
−v0
2 cosh2 (γH) cos2 (γH) + sinh2 (γH) sin2 (γH)

4.1.3 Numerical test: drift and slope current


The Ekman spiral test for the drift current belongs to the standard tests of the
Telemac-3D software. Therefore, very similar test parameters are chosen as in the
standard test [83]. The basic mesh is shown in appendix, figure A.1. It measures 2000 m
× 2000 m, giving a horizontal resolution of approximately 100 m; in the vertical direction
21 mesh levels are applied. The bottom is flat, whereby the depth varies from case to
case in order to observe different velocity profiles. By the greatest depth applied (160 m),
the vertical resolution is 8 m.
In order to test the three-dimensional diffusion step exclusively, all other steps of the
algorithm were switched off. In the case of the drift current, the stationary solution of
a diffusion equation with a source term (the Coriolis force) for a zero initial condition
for the velocity in the domain and a given boundary condition is achieved. For the case
of the slope current, a solution of a diffusion equation with a source term (a sum of the
Coriolis force and a constant pressure gradient), with a corresponding initial condition
for the velocity and a given boundary condition is obtained.
The solutions for the Ekman spirals are two-dimensional, for (u, v), with w = 0 when
the result converges. For an additional test, if the stage is fully three-dimensional, the
velocity diffusion subroutine was called for (u, v, w), (u, w, v) and (w, v, u) in the program
code. i.e. with changing order of the velocity components, Coriolis forcing terms and
boundary conditions. The final results were compared if they are identic.

4.1.3.1 Drift current

The surface wind stress is computed according to the empirical formula given by equa-
tions (2.188) and (2.189). The wind velocity, directed North, i.e. coincidental with the
positive y–direction, is set at 32 m/s. At the assumed latitude of 50o North (Coriolis
coefficient f = 1.1 × 10−4 s−1 and with the vertical viscosity νz = 0.1 m2 /s the wind
shear yields a deflection 45o to the right from the wind direction when the stationary
state is achieved. According to (4.13), the drift current velocity is v0 ≈ 1 m/s at the
free surface. The depth of frictional influence in this case is D = 133.68 m (4.10). The
only parameter varied in the test cases is the water depth H, corresponding to 0.1D,
0.25D, 0.5D and 1.2D, (13.37 m, 33,42 m, 66.84 m and 160 m). Horizontal diffusion
coefficients are taken to equal 0. A no-slip boundary condition is set at the bottom. The
initial condition is null velocity in the entire domain and a constant water depth. The
time step is 50 s and in the integration time is 56 h, i.e. 4000 time steps. The results are
presented in the figure 4.1. For depths 0.1D–0.5D, the numerical solutions converge to
114 4.1 Diffusion step test: Ekman profiles

Ekman spiral test - slope current

1.0

H=0.25D
0.8

0.6 H=1.2D
H=0.5D
v [m/s]

H=0.1D
0.4

0.2

0.0

-0.2
-0.2 0.0 0.2 0.4 0.6 0.8 1.0
u [m/s]

Ekman spiral test - drift current

0.03
H=0.25D
H=0.5D
0.025

0.02

0.015
v [m/s]

0.01
H=1.2D
0.005

0.0

-0.005
0.0 0.01 0.02 0.03 0.04 0.05 0.06
u [m/s]

Figure 4.1: Ekman spirals in the surface layer for finite water depth and various values
of H/D Above: drift current. Below: slope current. Circles correspond to the numerical
result and solid lines are the Ekman analytical solutions.
4.2 Tests of the Poisson equation step 115

1.0

u [m/s] 0.8

0.6

0.4

0.2

0.0
0 20 40 60 80 100
computation time [h]

Figure 4.2: Ekman test case, drift current. Convergence of the u velocity component at
the surface to the analytical solution for H = 1.2D

the analytical ones within the prescribed 56 h. The convergence behaviour for the most
slowly converging case H = 1.2D is presented in figure 4.2 for an extended integration
time of 96 h.

4.1.3.2 Slope current

For the same set of parameters and mesh, the bottom Ekman profiles for the slope cur-
rent are computed for three different depths, 0.25D, 0.5D, 1.2D. The constant pressure
gradient in the y–direction (North) corresponds for the assumed latitude with the con-
stant geostrophic velocity ug = 0.05 m/s in the x–direction (East), equation (4.3). The
constant pressure gradient is given as a constant source term in the diffusion step of
the algorithm. The initial velocity is equal ug and directed East. A no-slip boundary
condition at the bottom is applied. The results are presented in figure 4.1. Again, for
the same integration time of 56 h, the numerical solutions converge to the analytical
ones.

4.2 Tests of the Poisson equation step


Two tests of the Poisson equation stage are presented. Both are comparisons of an
analytical solution with a numerical one. First, a solution of a problem thoroughly
116 4.2 Tests of the Poisson equation step

determined by boundary conditions is given in order to test the boundary conditions


implementation. The source term of the Poisson equation is set to null. In the second
case, the boundary conditions are simple, but the source term is not equal to zero in
order to observe the behaviour of the numerical scheme in the this situation.
In both cases the numerical solution was obtained using the mesh shown in fig. A.1 with
21 vertical levels, setting the absolute exactness of the linear equation solver (conjugate
gradient method with diagonal pre-conditioning) to 10−6 . Because the applied analytical
solutions are two-dimensional, for each case two runs were made: one in a vertical
and another in a horizontal plane, respectively. All other stages of the algorithm were
switched off, eventually reducing the program to a solver for the Poisson equation.

4.2.1 Testing the boundary conditions influence


The Poisson equation solver for the pressure (or the hydrodynamic pressure, or a pressure
correction) must work appropriately with complicated boundary conditions given on
various boundaries of the domain. In this section the numerical solution of the Poisson
equation is compared with an analytical one with complex boundary conditions. In order
to concentrate on the influence of the boundary conditions separately, the source term
of the right hand side of the Poisson equation is set to null (Laplace equation), so that
the problem solution depends thoroughly on the boundary conditions.
The solution of the Laplace equation ∇2 T (x, y) = 0 in a two-dimensional rectangular
area x ∈ [0, a] and y ∈ [0, b] is considered. The following boundary conditions are
applied:

1. for x = 0, y ∈ [−b, b] : T = T0 (Dirichlet);

2. for x = a, y ∈ [−b, b] : ∂T /∂x = 0 (Neumann, set to 0);

3. for y = −b and y = b, x ∈ [0, a] : ∂T /∂x = −hT (Neumann).

The solution follows the method of variable separation (Fourier method):

T (x, y) = u(x)v(y) (4.18)

u′′ v ′′
∇2 T (x, y) = 0 ⇒ u′′ (x)v(y) + u(x)v ′′ (y) = 0 ⇒ = − = p2 (4.19)
u v
Taking into consideration the boundary conditions we obtain the following equation sets
for u and v:
′′ 2

 u −p u=0


u (a) = 0 (4.20)
u(0)v(y) = T0

v ′′ + p2 u = 0



v ′ (−b) = −hv(−b) (4.21)
 ′
v (0) = −hv(b)

4.2 Tests of the Poisson equation step 117

The solution is: (


u=cosh[p(a − x)]
(4.22)
v=cos(py)

where(p = pk , k = 1..∞) are the successive solutions (eigenvalues) of:


pk 1
= (4.23)
h tan(pk b)

The searched solution has the form of a sum:



X
T (x, y) = Ak cosh[pk (a − x)] cos(pk y) (4.24)
k=1

From the Dirichlet boundary condition one finds:



X
T0 = Ak cosh(pk a) cos(pk y) (4.25)
k=1

The coefficients Ak can be obtained from the Fourier theorem:


Rb
T0 cos(pk y)dy
Ak cosh(pk a) = R0 b 2
(4.26)
0 cos (pk y)dy

which yields:
2T0 sin(bpk )
Ak = (4.27)
[bpk + sin(bpk ) cos(bpk )] cosh(apk )
The final solution is:

X 2T0 sin(bpk ) cosh[pk (a − x)] cos(pk y)
T (x, y) = (4.28)
k=1 [bpk + sin(bpk ) cos(bpk )] cosh(apk )

In order to find the values of the solution (4.28) the eigenvalues pk of (4.23) are found
numerically using the regula falsi algorithm [127]. The assumed values for the boundary
conditions are T0 = 1 and h = 1, for the geometry a = 2 and b = 1. The comparison
between the numerical and analytical solution is given in fig. 4.3, showing an excellent
agreement.

4.2.2 Testing the source term influence


In this section a solution of the Poisson equation with a non-zero source term for simple,
symmetrical Neumann boundary conditions is investigated.
The solution of the Poisson equation
∂2T ∂2T
+ =q (4.29)
∂x2 ∂y 2
118 4.2 Tests of the Poisson equation step

y numeric solution y analytic solution


1.0 1.0
0.30 0.30

0.3 0.3
5 5

0.5 0.5
0.85
0.80

0.75

0.55
0.95
0.90

0.70

0.65

0.95
0.90
0.85
0.80
0.75

0.70

0.65

0.60

0.50

0.45

0.40
0.60

0.55

0.50

0.45

0.40
0.0 0.0

-0.5 -0.5

5 5
0.3 0.3

0.30 0.30

-1.0 -1.0
0.0 0.5 1.0 1.5 2.0 0.0 0.5 1.0 1.5 2.0
x x

Figure 4.3: Poisson equation verification. Solutions determined by boundary conditions.

is considered in a two-dimensional rectangular area x ∈ [−a, a] and y ∈ [−b, b]. The


following symmetrical Neumann boundary conditions are applied:
x = −a, y ∈ [−b, b] (∂T /∂x)x=−a + hT (−a, y) = 0 (4.30)
x = a, y ∈ [−b, b] (∂T /∂x)x=a − hT (a, y) = 0 (4.31)
y = −b, x ∈ [−a, a] (∂T /∂y)y=−b + hT (x, −b) = 0
y = b, x ∈ [−a, a] (∂T /∂y)y=b − hT (x, b) = 0

Due to the symmetry of the boundary conditions it can additionally be stated that:
(∂T /∂x)x=0 = 0 (∂T /∂y)y=0 = 0 (4.32)

As in the previous section the analytic solution method follows the method of variable
separation (Fourier method), whereby the solution is searched for in the form:

X
T (x, y) = Xm (x)Yn (y) (4.33)
m,n=1

where Xm and Yn are linear combinations of the sin and cos functions:
Xm (x) = Am cos(Mm x) + Bm sin(Mm y) (4.34)
Yn (x) = Cm cos(Nm y) + Dm sin(Nm y) (4.35)

From the boundary conditions (4.30) and (4.32) one finds:


^
′ ′ µm
Xm (0) = Xm (a) + hXm (a) = 0 ⇒ Bm = 0 ∧ Mm = (4.36)
m a
γ n
Yn′ (0) = Yn′ (b) + hYn (b) = 0 ⇒ Dm = 0 ∧ Nn =
^

n b
4.3 Tests of the free surface step 119

where µm and γn are successive solutions (eigenvalues) of the following equations:


µ tan µ = ah γ tan γ = bh (4.37)

Therefore, the sought solution has the following form:



X µm γn
T (x, y) = Hmn cos( x) cos( y) (4.38)
m,n=1 a b

where Hmn = Am Cn . Substituting (4.38) into (4.29) one obtains:



µ2n γn2
!
X µm γn
Hmn + cos( x) cos( y) = −q (4.39)
m,n=1 a2 b2 a b

The coefficients Hmn are to be found from the Fourier formula:


Ra Rb µm γn
µ2n γn2
!
−q −a −b cos( a x) cos( b y)dxdy
Hmn + 2 = Ra Rb
2 µm 2 γn
(4.40)
a2 b −a −b cos ( a x) cos ( b y)dxdy

After computing integrals in (4.40) one obtains:


−4q sin µm sin γn
Hmn =  µ2 γ 2
 (4.41)
n
a2
+ bn2 (sin µm cos µm + µm )(sin γn cos γn + γn )

The final solution is:



X sin µm sin γn cos( µam x) cos( γbn y)
T (x, y) = −4q 
µ2n 2
γn
 (4.42)
m,n=1 a2
+ b2
(sin µm cos µm + µm )(sin γn cos γn + γn )

where µm and γn are obtained from the equations (4.37) numerically using the regula
falsi algorithm [127].
The assumed value for the Neumann boundary conditions is h = 1 and the source term
is set as q = 1, for the geometry a = b = 1. The comparison between the numerical and
analytical solution is given in fig. 4.4, showing a thoroughly acceptable overall agreement.
The greatest discrepancies between the solutions are in the regions where the linear
interpolation functions are not able to reproduce varying slopes of the solution function,
especially in the middle of the domain.

4.3 Tests of the free surface step


4.3.1 Kinematic boundary condition
In this section tests of the free surface algorithms based on the kinematic boundary
condition are conducted. The kinematic boundary condition is identical to the two-
dimensional advection equation and therefore standard tests for this equation type are
taken into consideration. The tests are performed for a constant, divergence-free velocity
field, and the free surface algorithm is treated entirely separately, i.e. there is no coupling
between the velocity field and free surface movements. The tests are performed, however,
using the full 3D-mesh, so that the σ-transformation takes place.
120 4.3 Tests of the free surface step

y
numeric solution y
analytic solution
1.0 1.0

0.
45

0.60
0.60

45

0.
45
0.

45
0.
0.
50

0.
50
50
0.65
0.

0.65
55

50
0.
0.
0.

0.
55
55

55
0.
0.70
0.70

0.5 0.5

0.8
0.75

0.75
0
0.80

0.0 0.0

-0.5 -0.5

0.70 0.70
55

0.
0.
0.

55

55
50
55

0.65
0.

0.65

0.
50 0.
0.

50
0. 45
50

0.
0.60 45 0.60 0.
0.

45
0.
45

-1.0 -1.0
-1.0 -0.5 0.0 0.5 1.0 -1.0 -0.5 0.0 0.5 1.0
x x

Figure 4.4: Poisson equation verification. Solution determined by a source term.

4.3.1.1 One-dimensional Gaussian profile advection

In the first test the performance of the algorithm is tested using a well-known example
of the advection of a Gaussian profile along the x-axis, i.e. a virtually 1D-example. The
numerical solution should not yield any deformation of the profile as it is advected in a
constant velocity field. In this test a square (10 × 10 m) basic mesh was used with the
resolution of ∆x = 0.2 m (figure A.1). 21 levels are used, resulting in the vertical mesh
dimension of 10 m. The mesh resolution along x-axis is 0.2 m, and a constant advection
velocity is taken: (u, v, w) = (0.2, 0, 0) m/s (which is not influenced by the profile shape).
The Courant number, characterising the problem, C = u∆t/∆x, is changed by adjusting
the time step ∆t only. As the initial condition a Gaussian profile y = exp[(x − m)2 /2σ 2 ]
centred in m = 2 m and with σ = 0.25 m is imposed. Therefore, the amplitude of the
profile is 10% of the vertical mesh dimension in the surrounding area.
The performance of the algorithm based on the method of characteristics is presented
in figure 4.5. The well-known feature of this method, yielding the best results for whole
Courant numbers C and stability independent of C, is confirmed. For C in the range
between the whole numbers, considerable numerical diffusion appears due to the spatial
interpolation in the finite elements, eventually damping the advected profile. The curves
in figure 4.5 are shown for time intervals 8 s, with total computation time of 40 s. The
first curve represents the initial condition. Because only the time-step was changed,
the total computation time is reached in different numbers of time-steps, from 400 for
C = 0.1 down to 20 for C = 2.0.
In order to check if the number of time-steps (i.e. time resolution) is additionally respon-
sible for the numerical diffusion for the lower Courant numbers, a test was conducted for
constant time-step of ∆t = 0.4 s, and different advection velocities u, equal to 0.05 m/s,
0.1 m/s, and 0.2 m/s (C = 0.4, 0.2 and 0.1). The results are shown in figure 4.6. The up-
4.3 Tests of the free surface step 121

per row of plots show the results for each 20 time-steps (i.e. 8 s) for the global advection
time of 40 s. In the lower row, the total advection times are changed accordingly to the
value of the advection velocity: 160 s, 80 s, and 40 s, so that a constant advection path
length of 10 m is obtained. The profiles are shown after travelling constant advection
paths, each 1.6 m long. It may be concluded that the number of time-steps affects the
numerical diffusion, but this effect is much less important than the errors connected with
the spatial interpolation.
For the given initial profile (10% of the water depth) and parameters the semi-implicit
SUPG (modified) algorithm for Courant numbers C > 1.6 becomes unstable. For a
Gaussian profile of 1% of the water depth, it remains stable for approximately C <
2.0. A comparison of results for different Courant numbers is given in figure 4.7. The
SUPG method shows much lower numerical diffusion for lower Courant numbers than
the characteristics method.

4.3.1.2 Rotating cone test

A well-known benchmark test for the performance of numerical solutions of the un-
steady two-dimensional advection equation is the rotating cone test, also known as the
Molenkamp test [162]. In this test a smooth, two-dimensional Gaussian shape is rotated
in a divergence-free, rotating velocity field without change of form. The parameters of
this test are standardised in the literature.
The advection equation for S(x, y, t) is solved in a domain x ∈ [−1, 1] and y ∈ [−1, 1],
where the velocity field describes a constant rigid-body rotation, (u, v) = (−ωy, ωx),
where ω = 2π. The axis of the rotation is in the middle of the computational domain.
The initial condition is given by a Gaussian distribution which extends slightly outside
the computational domain:
2
S(x, y, 0) = 0.014r , r 2 = (x + 0.5)2 + y 2 (4.43)

This means that the Gaussian hill of the amplitude equal 1 is centred in the position
between the rotation axis and the domain edge, (x0 , y0 ) = (−0.5, 0). The velocity mag-
nitude at the top of the hill is π m/s and at the edge 2π m/s.
The boundary conditions are given at the boundaries by imposing the exact analytical
solution of the problem (Dirichlet BC):
2
S(x, y, 0) = 0.014r , r 2 = (x + 0.5 cos ωt)2 + (y + 0.5 sin ωt)2 (4.44)

The numerical solution is required after one revolution, i.e. for t = 1 s.


In this particular test a √
square (2 × 2 m) basic mesh is taken with resolution of ∆x =
0.04 m (or ∆x = 0.04 · 2 m diagonally, see figure A.1). 21 levels are used, with the
bottom position −1 m (cone amplitude is also equal 1 m).
As the characteristic Courant number for this problem, the one at the top of the Gaussian
hill is assumed C = u∆t/∆x. The maximal Courant number in the domain equals 2C at
122 4.3 Tests of the free surface step
10

10
8

8
6

6
x [m]

x [m]
4

4
2

2
C=0.8

C=2.0
0

0
1.0

0.8

0.6

0.4

0.2

0.0

1.0

0.8

0.6

0.4

0.2

0.0
y [m]

y [m]
10

10
8

8
6

6
x [m]

x [m]
4

4
2

2
C=0.4

C=1.6
0

0
1.0

0.8

0.6

0.4

0.2

0.0

1.0

0.8

0.6

0.4

0.2

0.0
y [m]

y [m]
10

10
8

8
6

6
x [m]

x [m]
4

4
2

2
C=0.1

C=1.0
0

0
1.0

0.8

0.6

0.4

0.2

0.0

1.0

0.8

0.6

0.4

0.2

0.0
y [m]

y [m]
Figure 4.5: Free surface with characteristics. Solution behaviour depending on value of
the Courant number C. Initial profile centred by x = 2 m. Profiles shown for each 8 s
for the global advection time of 40 s.
4.3 Tests of the free surface step 123

10

10
8

8
6

6
x [m]

x [m]
0.2 m/s, 100 t

0.2 m/s, 100 t


4

4
2

2
0

0
1.0

0.8

0.6

0.4

0.2

0.0

1.0

0.8

0.6

0.4

0.2

0.0
y [m]

y [m]
10

10
8

8
6

6
x [m]

x [m]
0.1 m/s, 100 t

0.1 m/s, 200 t


4

4
2

2
0

0
1.0

0.8

0.6

0.4

0.2

0.0

1.0

0.8

0.6

0.4

0.2

0.0
y [m]

y [m]
10

10
8

8
6

6
0.05 m/s, 100 t

0.05 m/s, 400 t


x [m]

x [m]
4

4
2

2
0

0
1.0

0.8

0.6

0.4

0.2

0.0

1.0

0.8

0.6

0.4

0.2

0.0
y [m]

y [m]

Figure 4.6: Free surface with characteristics. Solution behaviour depending on the num-
ber of constant time-steps (∆t = 0.4 s) for various advection velocities u = 0.05 m/s,
0.1 m/s, and 0.2 m/s (C = 0.4, 0.2 and 0.1). Initial profile centred by x = 2 m. The
profiles shown: above – for each 20 time-steps (8 s), below – for each 1.6 m distance
travelled (each 80, 40 and 20 timesteps, respectively).
2 m. Profiles shown for each 8 s for the global advection time of 40 s.
behaviour depending on value of the Courant number C. Initial profile centred by x =
Figure 4.7: Comparison between the method of characteristics and SUPG. Solution

124
C=0.1, characteristics C=0.2, characteristics C=0.4, characteristics
y [m] y [m] y [m]
1.0 1.0 1.0

0.8 0.8 0.8

0.6 0.6 0.6

0.4 0.4 0.4

0.2 0.2 0.2

0.0 0.0 0.0


0 2 4 6 8 10 0 2 4 6 8 10 0 2 4 6 8 10
x [m] x [m] x [m]

C=0.1, SUPG C=0.2, SUPG C=0.4, SUPG


y [m] y [m] y [m]
1.0 1.0 1.0

4.3 Tests of the free surface step


0.8 0.8 0.8

0.6 0.6 0.6

0.4 0.4 0.4

0.2 0.2 0.2

0.0 0.0 0.0


0 2 4 6 8 10 0 2 4 6 8 10 0 2 4 6 8 10
x [m] x [m] x [m]
4.3 Tests of the free surface step 125

Characteristics Semi-implicit SUPG


C = u∆t/∆x Peak value C = u∆t/∆x Peak value
0.1 0.3070 0.1 0.9668
0.2 0.3180 0.2 0.9463
0.5 0.3648 0.5 0.8886
1.0 0.5695 0.8 0.8365
2.0 0.7072 0.9 0.8190
1.0 0.8061
1.6 0.7354

Table 4.1: Rotating profile test. Reduction of the initial peak value 0.9927 after one
revolution for methods based on kinematic BC.

the boundary, minimum 0 at the axis. The performance of both free surface algorithms
based on the kinematic boundary condition is tested for different Courant numbers,
varied between 0.1 and 2 by adjusting the time step only, resulting in time-steps varying
between 0.00125 s and 0.025 s and time-step numbers between 40 and 800, respectively.

The difficulty of this test is due to the following factors: (1) The Courant number changes
along the distance between the axis of the rotation and the domain boundary, i.e. it is
different in various parts of the advected profile; (2) the direction of advection relative to
the mesh elements changes along the way; (3) the slope of the profile is relatively large;
(4) the boundary conditions and the specific condition (u, v) = (0, 0) in the middle of
the domain interfere with the solution as the numeric diffusion takes place.

The worst and the best results for both methods are shown in figures 4.8 and 4.9 for the
method of characteristics, and in figures 4.11 and 4.10 for modified semi-implicit SUPG
(θ = 0.55), respectively. Additionally, a table with results for various C is provided
(Table 4.1). The initial peak value by the given discretisation is 0.9927 m. Both methods
for the given range of Courant numbers are free of oscillations and the general shape
of the advected profile remains. However, large numerical diffusion for lower Courant
numbers for the method of characteristics disqualifies it in comparison with SUPG. The
SUPG method brings better results for lower Courant numbers and comparable results
for higher ones, but with C > 1.6 the shape of the profile is no longer maintained and
oscillations occur disqualifying the solution for C ≈ 2. For the implicit SUPG method,
the peak value is much more damped, e.g. for C = 1.6, the peak value is ca. 50% lower,
0.3542.

The practical test conclusions are that for the lower Courant numbers (and larger num-
ber of time-steps) the SUPG method is preferred. However, for larger Courant numbers
the method of characteristics remains the only one available. Not surprisingly, the per-
formance of this method is best for whole or higher Courant numbers and lower number
of time-steps.
126 4.3 Tests of the free surface step

initial condition after rotation

Figure 4.8: Rotating profile test, kinematic BC, method of characteristics, C = 0.1. The
maximum value after one rotation is 0.307.

initial condition after rotation

Figure 4.9: Rotating profile test, kinematic BC, method of characteristics, C = 2.0. The
maximum value after one rotation is 0.707.
4.3 Tests of the free surface step 127

initial condition after rotation

Figure 4.10: Rotating profile test, kinematic BC, SUPG, C = 0.1. The maximum value
after one rotation is 0.967.

initial condition after rotation

Figure 4.11: Rotating profile test, kinematic BC, SUPG, C = 1.6. The maximum value
after one rotation is 0.736.
128 4.3 Tests of the free surface step

initial condition after rotation

Figure 4.12: Rotating profile test, entirely explicit conservative free surface equation,
C = 0.5. The maximum value after one rotation is 0.369, the minimum value is -0.122.

4.3.2 Conservative free surface equation


The algorithm for finding the free surface elevation from the conservative free surface
equation was tested using the rotating cone test for identical parameters as in the pre-
vious section.
The test disqualified the application of this algorithm in its entirely explicit version
(section 3.6.3) in comparison to the kinematic boundary condition with the method of
characteristics and SUPG. The scheme proves to be numerically very diffusive for lower
Courant numbers (for C = 0.1 the peak was damped to 0.044 from the initial value 1.0,
and for C = 0.2 to 0.107) and unstable beginning from Courant number C = 0.5 (peak
value of 0.369 by oscillations amplitude down to -0.122, figure 4.12).
For the implicit and semi-implicit formulation for the conservative free surface equation
(section 3.6.4), the results are slightly better than obtained with the methods based
on SUPG for kinematic boundary condition, table 4.2. The computations with semi-
implicit (θ = 0.55) and implicit formulations remain stable for large Courant numbers,
with growing damping of the amplitude and numerical oscillations. By θ = 0 (explicit
formulation) strong oscillations appear already by C = 0.2 leading to a program break-
up. As a consequence, only the semi-implicit SUPG version of the method based on
the conservative free surface equation is adequate for practical purposes. Additionally,
this scheme has better mass-conserving properties than the SUPG method based on
kinematic boundary condition, as investigated further in section 5.5.4.
4.3 Tests of the free surface step 129

Semi-implicit Implicit
C = u∆t/∆x Peak value Peak value
0.1 0.9602 0.8098
0.2 0.9467
0.5 0.8880
0.8 0.8369 0.4705
1.0 0.8060 0.4308
2.0 0.6993 0.3213
5.0 0.5101 0.2140

Table 4.2: Rotating profile test. Reduction of the initial peak value 0.9927 after one
revolution for the conservative free surface equation with SUPG.

initial condition after rotation

Figure 4.13: Rotating profile test, semi-implicit conservative free surface equation with
SUPG, C = 0.1. The maximum value after one rotation is 0.960.
130 4.3 Tests of the free surface step

initial condition after rotation

Figure 4.14: Rotating profile test, semi-implicit conservative free surface equation with
SUPG, C = 2.0. The maximum value after one rotation is 0.699.
Chapter 5

Model verification

Nothing shocks me, I am a scientist.


Indiana Jones

After the verification operator-splitting algorithm steps as separate units (previous chap-
ter), this chapter considers the verification of the entire algorithm.
The broad range of event-oriented test cases presented in this chapter is intended to
cover the intended extension of the initial hydrostatic model application domain. The
verification cases document the functionality of the developed model as well as provide
the evidence for the assessment of realised numerical developments.
The features of the new algorithm are illustrated in the most convincing way perform-
ing simulations of those physical systems, where a solution based on the hydrostatic
approximation is not adequate (physically infeasible), though mathematically possible.
Two examples of this type are provided without aiming for formal verification, but im-
mediately showing the new model’s capabilities: waves reflection and interaction with
topography, and a wind driven vertical circulation in a closed basin.
For internal flows, the lock exchange flow as well as the interfacial internal waves of non-
hydrostatic nature are well suited to verification of model performance. The obtained
results are compared partially with the existing theory.
The best benchmark tests allowing the strict verification using analytical solutions are
probably small-amplitude short surface waves, impossible to reproduce properly with a
hydrostatic model. The analytical solutions make it possible not only to compare the
wave periods, but velocity and pressure profiles as well. As a standard test case, a
standing wave in a closed basin is considered. This test is also a good example for the
assessment of the model’s conservative properties and its stability. For completeness,
the performance of the non-hydrostatic model is tested also for the hydrostatic (shallow-
water) case of long standing waves.
For channel flow, the subcritical and especially the supercritical flow with free surface
over a ramp, provide an excellent verification case for the free surface algorithm. For
this test case no solutions with the hydrostatic model can be obtained: in the subcritical

131
132 5.1 Waves reflection

case due to the bottom slope larger than 1:10, and for the supercritical case additionally
due to the necessity of taking hydrodynamic pressure into consideration.
The complex boundary conditions needed for the free surface computations for wave
problems or where disturbances generated inside the domain must leave it through the
boundaries (absorbing or non-reflecting BC) are tested in a channel flow example with
a bump at the bottom and in case of waves travelling over an underwater channel.
The ability of the model to describe the propagation of the non-linear waves of a finite
amplitude is tested in a few examples concerning solitary waves. In ideal conditions,
the solitary waves propagate over a flat bottom without changing their properties. The
accuracy of the model is evaluated by comparison of the wave amplitude, shape and
celerity as it travels. Wave-wave interactions and the influence of a changeable bottom
are additionally tested.
Additionally, a few examples testing the limits of applications (free surface and bottom
steepness) are presented.

5.1 Waves reflection


This section yields an immediate check whether the non-hydrostatic algorithm repro-
duces short (deep water) motion in a fully three-dimensional problem. A square basin
with a side of 100 m and a depth of 10 m is filled with a non-viscous fluid. The initial
shape of the free surface S(x, y, t = 0) is a two-dimensional cosine profile with a peak
of height η0 = 0.2 m and width of D = 40 m located exactly in the middle of the basin
(x0 , y0 ) = (50, 50) m:
S(x, y, 0) = η0 [1 + cos(2πr/D)] r = [(x − x0 )2 + (y − y0 )2 )]1/2 (5.1)

The aim of this test is to observe the wave patterns developing for t > 0 and their reflec-
tion from the boundaries. The square basic mesh has 5000 regular triangular elements
and a resolution of approx. 2m, and 21 mesh levels with equal distribution (figure A.1).
In the hydrostatic case, the convection and diffusion steps of the algorithm are switched
off, so that only the pressure-continuity step is executed. In the non-hydrostatic case,
the diffusion coefficients are set to 0 and convection step is excluded. In consequence,
Euler equations are solved.
The results are shown in figures 5.1 and 5.2 for hydrostatic and non-hydrostatic case,
respectively. The resulting wave patterns for 5, 6 and 10 s differ much due to various
wave celerity in both models. The hydrostatic model treats all occurring waves as long
waves with a celerity of c = (gh)1/2 ≈ 10.0 m/s, (gravitational acceleration g is set to
10.0 m/s2 for interpretation simplicity) whereby the non-hydrostatic one resolves the
short wave motion with lower celerity (see appendix A.3).
This simple test case immediately shows new properties of the non-hydrostatic model –
the ability to simulate short free surface waves.
One of the most interesting practical applications of this kind is the wave propagation in
harbour basins, especially in connection with sediment erosion, deposition and transport
5.1 Waves reflection
initial condition 5.0s
100 100
Figure 5.1: Wave patterns computed with hydrostatic approximation.
S [m]
S [m]
0.050
0.200 0.045
80 0.180 80 0.040
0.170 0.035
0.160 0.030
0.150 0.025
0.140 0.020
60 0.130 60 0.015
y [m]

y [m]
0.120 0.010
0.110 0.005
0.100 -0.000
0.090 -0.005
40 0.080 40 -0.010
0.070 -0.015
0.060 -0.020
0.050 -0.025
0.040 -0.030
20 0.030 20 -0.035
0.020 -0.040
0.010 -0.045
0.000 -0.050

0 0
0 20 40 60 80 100 0 20 40 60 80 100
x [m] x [m]
6.0s 10.0s
100 100
S [m] S [m]
0.050 0.050
0.045 0.045
80 0.040 80 0.040
0.035 0.035
0.030 0.030
0.025 0.025
0.020 0.020
60 0.015 60 0.015
y [m]

y [m]
0.010 0.010
0.005 0.005
-0.000 -0.000
-0.005 -0.005
40 -0.010 40 -0.010
-0.015 -0.015
-0.020 -0.020
-0.025 -0.025
-0.030 -0.030
20 -0.035 20 -0.035
-0.040 -0.040
-0.045 -0.045
-0.050 -0.050

0 0
0 20 40 60 80 100 0 20 40 60 80 100
x [m] x [m]

133
134
initial condition 5.0s

Figure 5.2: Wave patterns computed without hydrostatic approximation.


100 100
S [m]
S [m]
0.050
0.200 0.045
80 0.180 80 0.040
0.170 0.035
0.160 0.030
0.150 0.025
0.140 0.020
60 0.130 60 0.015

y [m]

y [m]
0.120 0.010
0.110 0.005
0.100 -0.000
0.090 -0.005
40 0.080 40 -0.010
0.070 -0.015
0.060 -0.020
0.050 -0.025
0.040 -0.030
20 0.030 20 -0.035
0.020 -0.040
0.010 -0.045
0.000 -0.050

0 0
0 20 40 60 80 100 0 20 40 60 80 100
x [m] x [m]
6.0s 10.0s
100 100
S [m] S [m]
0.050 0.050
0.045 0.045
80 0.040 80 0.040
0.035 0.035
0.030 0.030
0.025 0.025
0.020 0.020
60 0.015 60 0.015
y [m]

y [m]
0.010 0.010
0.005 0.005
-0.000 -0.000
-0.005 -0.005
40 -0.010 40 -0.010
-0.015 -0.015

5.1 Waves reflection


-0.020 -0.020
-0.025 -0.025
-0.030 -0.030
20 -0.035 20 -0.035
-0.040 -0.040
-0.045 -0.045
-0.050 -0.050

0 0
0 20 40 60 80 100 0 20 40 60 80 100
x [m] x [m]
5.1 Waves reflection 135

x [m]
0 10 20 30 40 50 60 70 80 90 100
100

90

80

70
[m]

60 0.45
0.40
0.35

y [m]
0.30
50 0.25
0.20
0.15
40 0.10
0.05
0.00
30 -0.05
-0.10
-0.15
-0.20
20 -0.25
-0.30
-0.35
10 -0.40
-0.45
-0.50
0

Figure 5.3: Wave pattern in a harbour basin for t = 50 s, ∆x = 1 m. Waves with


T = 4 s, λ = 24.7 m, c = 9.2 m/s.
x [m]
0 50 100 150 200 250 300 350 400 450 500
500

450

400

350
[m]

0.45
300
0.40
0.35
0.30
y [m]

250 0.25
0.20
0.15
200 0.10
0.05
0.00
-0.05
150
-0.10
-0.15
-0.20
100 -0.25
-0.30
-0.35
50 -0.40
-0.45
-0.50
0

Figure 5.4: Wave pattern in a harbour basin for t = 50 s, ∆x = 1 m. Waves with


T = 10 s, λ = 92.4 m, c = 6.2 m/s.
136 5.2 Wind-driven circulation

by currents. (In principle, the newly developed algorithm allows the simulation of all
these phenomena in one model.) A simple example of such an application is shown in
figures 5.3 and 5.4. The mesh for this test case is obtained from the square mesh shown
in figure A.1 by cutting appropriate areas out. In the first case, figure 5.3, sinusoidal
waves of a period of T = 4 s and amplitude of η0 = 0.5 m enter a small harbour basin
of a typical geometry through the southern entrance. The mesh resolution is ∆x = 1 m.
The water depth is everywhere equal to 10 m. Under this conditions, the waves have
the length λ = 24.7 m and phase speed c = 6.2 m/s (values computed with shallow
water approximation are 39.6 m and 9.9 m/s, respectively). Water viscosity is assumed
to be isotropic and equal to ν = 10−4 m2 /s. Dynamic BC are applied at the free surface.
Figure 5.3 shows the free surface configuration at t = 50 s from the begin of the wave
action. In figure 5.4 waves of a period of T = 10 s enter a five times larger basin
(mesh resolution of ∆x = 5 m). This time, the wavelength is λ = 92.4 m and phase
speed c = 9.2 m/s (values computed with shallow water approximation are 99.0 m and
9.9 m/s, respectively). The free surface configuration is shown at t = 50 s.
In both cases the model reproduces well the typical phenomena such as: interaction of
incident and reflected waves at the diagonal north-eastern wall forming a characteris-
tic chessboard pattern, or wave diffraction at the western end of the entrance channel
interfering with the incoming waves.

5.2 Wind-driven circulation


Another example illustrating the differences between the hydrostatic and non-hydrostatic
computation is the well known example of a wind-driven circulation in a closed basin.
In this case the vertical circulation in a deep water body caused by wind shear at the
free surface is treated (i.e. a Neumann BC for the velocity is given). A pool measuring
10 m × 10 m and with a depth of 5 m is considered, with a strong wind blowing over
the surface in the x-direction. The square basic mesh is shown in figure A.1; 21 evenly
distributed levels are used. The horizontal resolution is 0.2 m and vertical 0.5 m.
The wind stress is computed according to the empirical formulation (2.189). The wind
velocity is taken to equal 10 m/s. The constant horizontal and vertical eddy viscosities
are identical in value and equal to 0.001 m2 /s. The Coriolis force is neglected. As the
initial condition null velocity in the entire domain is taken. The free surface is computed
using the Telemac2D algorithm (section 2.8.8). The integration time is 1000 s (steady
state achieved) with the time step of 1 s.
The results are shown in figure 5.5. The hydrostatic computation yields an intensive
vertical circulation near the wall. The flow seems to be strongly accelerated along the
wall without presence of any vertical pressure gradient. This is an effect of obtaining the
vertical velocity component from the continuity equation only. The fluid is accelerated
along the fetch path of the wind (a Neumann BC) in the horizontal direction, but on
the wall a Dirichlet impermeability condition sets the horizontal velocity component to
zero. Due to the continuity equation ∇ · u = 0, large ∂u/∂x causes large ∂w/∂z, which
5.2 Wind-driven circulation 137

hydrostatic
0

-1

-2
z [m]

-3

-4
v [m/s]
0.1
-5
0 1 2 3 4 5 6 7 8 9 10
x [m]

non-hydrostatic
0

-1

-2
z [m]

-3

-4
v [m/s]
0.1
-5
0 1 2 3 4 5 6 7 8 9 10
x [m]

Figure 5.5: Wind-driven circulation test case.


138 5.3 Interfacial internal waves

explains the circulation pattern near the wall. This strong effect can be weakened when
larger fluid viscosities are taken, smoothing the velocity gradients.
The non-hydrostatic model results in a much more widespread circulation over the entire
area of the basin. The presence of hydrodynamic pressure gradients before the wall
decelerates the fluid in a much smoother way than in the previous case, when the only
mechanism to diminish the velocity gradients was provided only by fluid viscosity. The
hydrodynamic pressure controls also the vertical circulation pattern, which takes a much
more natural form.
This test case shows in a very illustrative way the role played by the hydrodynamic
pressure when an interaction of fluid with a wall or structure and flows in the form of a
vertical circulation is considered.

5.3 Interfacial internal waves


5.3.1 Theory
In this section waves appearing at the interface between two fluids of different density
are considered. The stratification is not continuous, so that a sharp step in fluid densities
appears at the interface. It is assumed that the stratification is stable, so that the density
of the lower layer ̺d is greater than the density of the upper layer ̺u with a free surface
(̺d > ̺u ). In the undisturbed state the thickness of the upper layer is Hu , the lower Hd
and the position of the undisturbed free surface is z = 0, the undisturbed interface z =
−Hu and of the flat bottom z = −H = −(Hu +Hd ). The problem was treated already by
Stokes in 1847 [147]. The discussion of linearised, small perturbation equations for both
layers under assumption of pressure and velocity continuity at the interface [147, 96, 92],
yields a fourth-order dispersion relationship (characteristic equation for the wave celerity
c) in the form (k = 2π/λ is the wave number):
g g2
c4 [̺u +̺d coth(kHu ) coth(kHd )]−c2 ̺d [coth(kHu )+coth(kHd )]+(̺d −̺u ) 2 = 0 (5.2)
k k
which has two solutions (eigenvalues of the problem) c1,2 [92]:
g̺d [coth(kHu ) + coth(kHd )]
c1,2 = (5.3)
2k[̺u + ̺d coth(kHu ) coth(kHd )]
#1/2
̺2d [coth(kHu ) + coth(kHd )]2
"
̺d − ̺u
± 2
− (5.4)
4[̺u + ̺d coth(kHu ) coth(kHd )] [̺u + ̺d coth(kHu ) coth(kHd )]

One of the roots (5.3) is connected with the free surface movements (c1 ), the other (c2 )
with the interface wave.
Assuming that (̺d −̺u )/̺d << 1 (as usually found in nature), the dispersion relationship
(5.2) can be transformed into the following approximate form (exact to O((̺d − ̺u )/̺d ):
" #
g g ̺d − ̺u

2 2
c − tanh[k(Hu + Hd )] c − =0 (5.5)
k k ̺u coth(kHu ) + ̺d coth(kHd )
5.3 Interfacial internal waves 139

showing two roots clearly. The first one is connected clearly with the gravity waves
in the water of total depth H = Hu + Hd , unaffected by stratification (because
(̺d − ̺u )/̺d << 1). The second root is the dispersion relationship of the waves at the
interface. ci are the celerities (phase speeds) of these wave movements. For a rigid lid
boundary condition at the surface this root is exact, but when movement of the free
surface is allowed, it is correct only to O((̺d − ̺u )/̺d ).
Discussion of (5.2) and (5.3) can be conducted in categories of thickness of the layers,
in typical configurations to be found in the nature, compared to the wave length at the
interface (figure 5.6).

1. If the thicknesses of both layers Hd and Hu are much greater than the interface
wave length λ = 2π/k, then coth(kHd ) ≈ 1 and coth(kHu ) ≈ 1 and from (5.2):
g g ̺d − ̺u
c21 = c22 = (5.6)
k k ̺u + ̺d

It should be noted that in practice coth(kHd ) → 1 and coth(kHu ) → 1 already for


kHu > π/2 and kHd > π/2. As a rule of thumb, Hd,u > 2λ. This case, with the
interface away from the free surface and the bottom, is clearly the short wave or
deep water approximation for interfacial waves. The short waves at the interface
are dispersive. By absence of the free surface mode, the velocity values vanish
exponentially with the vertical distance from the interface with an asymmetry
depending on the density difference [92, 40].
2. If the wave length λ is much greater than both layer thicknesses Hu and Hd (shallow
water), then coth(kHu ) → 1/kHu and coth(kHd ) → 1/kHd and from (5.3) the
following non-dispersive case can be obtained:
̺d − ̺u
c21 = g(Hd + Hu ) = gH c22 = gHd Hu (5.7)
̺d Hd + ̺u Hu

The celerity of the internal mode c2 can be further simplified if ̺d ≈ ̺u :


̺d − ̺u Hd Hu
c22 = g (5.8)
̺d Hd + Hu

When the deeper layer is also much thicker than the upper one, Hd >> Hu :
̺d − ̺u
c22 = g Hu (5.9)
̺d

The last case is clearly the non-dispersive long wave (or shallow water) approxi-
mation. For practical applications of this approximation, it should be noted that
coth(kHu,d ) ≈ 1/kHu,d already for λ > 20Hu,d. When the densities in both layers
are similar, (̺d − ̺u )/̺d << 1 and the deeper layer is much thicker than the upper
one (Hd >> Hu ), then:
̺d − ̺u
c22 = g Hu (5.10)
̺u
140 5.3 Interfacial internal waves

z
6
(1) 0
6
̺u λ << Hu
Hu λ << Hd
λ
?
6

̺d Hd

? −(Hu + Hd )

λ
(2a)
6 λ >> Hu
Hu λ >> Hd
?
6
Hd
?

(2b)
6 λ >> Hu
Hu λ >> Hd
? Hd >> Hu
6

Hd

(3/4)
6 λ << Hd
Hu Hu arbitrary
λ
?
6

Hd

Figure 5.6: Various configurations of interfacial internal waves. (Explanations in text.)


5.3 Interfacial internal waves 141

and the long wave movement is analogous to the non-dispersive long surface wave
in homogeneous water, but with reduced gravity acceleration g ′ = g(̺d − ̺u )/̺u .
Accordingly, by absence of the free surface mode, the profiles of the horizontal
velocity do not change with the depth.

3. If the thickness of the lower layer Hd is much greater than the interface wave length
λ (Hu arbitrary), coth(kHd ) ≈ 1 and from (5.2):

g g ̺d − ̺u
c21 = c22 = (5.11)
k k ̺u + ̺d coth(kHu )

4. If a very thin upper layer Hu < λ/20 lies over a thick, dense layer Hd > λ/2,
coth(kHu ) → 1/kHu and coth(kHu ) ≈ 1, (5.2) yields:

g ̺d − ̺u
c21 = c22 = gHu (5.12)
k ̺d + ̺u kHu

It can be shown that using as a starting point the small perturbation equation with the
hydrostatic approximation, the non-dispersive case of interfacial internal waves can be
obtained [53], which proves the equivalency of long wave and hydrostatic approximations
in this case. The celerity of the interface waves are always smaller than the celerity of
the surface ones, with the most striking example of (5.10).

5.3.2 Test realisation


For testing the performance of the non-hydrostatic algorithm the dispersive, short inter-
facial wave example is taken. This is clearly the non-hydrostatic case, when both layer
thicknesses, below and above the interface, are larger than the wave length. A standing
uni-nodal interfacial wave in a closed basin of length 1 m (i.e. wave length 2 m) and
water depth 10 m with both layer depths of 5 m has a sinusoidal initial shape with an
amplitude of 1 m. The initial velocity is assumed to be zero everywhere. The lower,
slightly denser layer has a salinity of 1 PSU and the layer above 0 PSU (see section A.4).
The densities of both fluids are computed with a linear formula (5.31) and equal 999.972
kg/m3 and 1000.722 kg/m3 , respectively. The Boussinesq reference density is 1000.347
kg/m3 .
The mesh used for computations has 2114 base elements and 1147 base nodes, resulting
in horizontal resolution of ≈ 0.01333 m. 51 mesh levels have been taken, distributed
more densely in the area of the interface movement. The viscosity of the fluid and
salinity diffusion coefficients are equal in all directions, ν = 10−2 m2 /s and time-step
0.1 s. For hydrodynamic pressure, a zero Neumann BC is applied at all boundaries.
∂w/∂z = 0 is given at the free surface (computed from kinematic BC with semi-implicit
SUPG method). The wave celerity computed with (5.6) is 0.0342 m/s and the period of
oscillation is 58.5 s. When the formula (5.7) is used, values of 0.136 m/s and 14.74 s are
obtained, respectively.
142 5.3 Interfacial internal waves

0 0

-2 -2

-4 -4
z [m]

z [m]
-6 -6

-8 -8
u [m/s]
0.5
-10 -10
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
x [m] x [m]

Figure 5.7: A short interfacial internal wave by t = 3.7 s (1/4 of the theoretical long
wave period) computed with hydrostatic approximation (physically false). Left: Velocity
iso-lines, u: dashed line, step value 0.005 m/s, w: chain-dotted line, step value 0.02 m/s.
The position of the interface shown with solid lines. Right: velocity vectors.

0 0

-2 -2

-4 -4
z [m]

z [m]

-6 -6

-8 -8
u [m/s]
0.01
-10 -10
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
x [m] x [m]

Figure 5.8: A short interfacial internal wave by t = 15 s (1/4 of the theoretical short
wave period) computed without hydrostatic approximation (physically correct). Left:
Velocity iso-lines, u: dashed line, step value 0.0005 m/s, w: chain-dotted line, step value
0.002 m/s. The position of the interface shown with solid lines. Right: velocity vectors.
5.4 Lock exchange flow 143

The results obtained with and without hydrostatic approximation for about 1/4 of the
wave oscillation period (3.7 s and 15 s, respectively), are strikingly different. While in the
(physically correct) non-hydrostatic case the velocity diminishes exponentially with the
distance from the interface, this is not the case in the hydrostatic (physically false) case,
compare figures 5.7 and 5.8. Neither model reproduces the wave period properly (the
periods are about 15% larger), due to softening of the interface profile during simulation
and viscosity influence.

As a conclusion, it can be stated that the model reproduces satisfactorily the interfacial
internal waves in the non-hydrostatic mode. However, for fully successful applications,
techniques influencing the diffusion and viscosity coefficients, which reduce strongly the
interface sharpness, should be investigated. First, it could be an appropriate turbu-
lence model (constant diffusion coefficients are applied here); second, a method to follow
the internal interface (e.g. based on separate σ-transformations below and above the
interface).

5.4 Lock exchange flow

The lock exchange flow is a well-known test case to verify the modelling of buoyancy-
driven (gravity, density) flows or currents. Two incompressible fluids of slightly different
density confined in a rectangular basin, with impermeable walls and a free surface, are
initially divided by a very thin wall, figure 5.9. The left half of the basin is filled by fluid
of higher density, and the right half by a fluid of lower density. In the initial moment of
time the impermeable thin wall in the middle of the basin is removed and a buoyancy-
driven flow develops. A mutual intrusion of both fluids takes place. The denser fluid
enters the area occupied initially by the lighter one along the bottom, and the lighter
one does the same along the surface in the opposite direction. Eventually, after the
interfacial wave has dissipated its energy, an equilibrium state will be achieved with two
layers of equal thickness, the lighter fluid over the denser one.

The name of this test case originates from the practical engineering problem concerning
intrusion of a salt water wedge under fresh water when a lock gate is opened at the
mouth of a fresh water channel leading to the sea. There are numerous other examples
of gravity currents of this type in nature e.g., on a larger scale, motion of a cold front
in a warmer atmosphere, or when a muddy stream enters a lake. Salt water wedges in
estuaries are another example.

The shape of the interface between the fluids is very different, when the flow simulation
is carried out with or without hydrostatic approximation. The hydrostatic solution
yields a characteristic rectangular pattern which is not observed in nature or in the
laboratory, where the noses of density currents advancing along a flat bottom have a
slope of approximately π/3 to the horizontal [159].
144 5.4 Lock exchange flow

z
6

̺2 ̺1 t=0

- x
z
6

̺2 ̺1 t>0

-
- x
Figure 5.9: Lock exchange problem illustration.

5.4.1 Hydrostatic interpretation


The behaviour of the hydrostatic solution of this test case problem concerning a two-
fluid system can be understood by analysing the solution of the shallow water wave
equation (i.e. hydrostatic) for an initial free surface displacement in the form of a
step. This section explains why the interface shape in the hydrostatic solution takes the
characteristic rectangular shape.
Integration of the continuity equation in the vertical direction, as performed in subsection
2.6.3.2, leads to the conservative form of the free surface equation (2.159):

∂S ∂ S ∂ S
Z Z
+ udz + vdz = 0 (5.13)
∂t ∂x −B ∂y −B

For long waves, the horizontal velocity component (A.44) is independent of depth (section
A.3) and (5.13) transforms to:

∂S ∂ ∂
+ [u(H + S)] + [v(H + S)] = 0 (5.14)
∂t ∂x ∂y

For small perturbation of the free surface compared with the water depth H + S → H
and one obtains:
∂S ∂ ∂
+ (uH) + (vH) = 0 (5.15)
∂t ∂x ∂y

The perturbation pressure in the long wave approximation is hydrostatic:

p′ = ̺gS (5.16)
5.4 Lock exchange flow 145

Introducing (5.16) to equations (A.20):

∂u/∂t = −g∂S/∂x
∂v/∂t = −g∂S/∂y (5.17)
(5.18)

Combining these equations with (5.15) leads to a wave equation:

∂2S
! !
∂ ∂S ∂ ∂S
2
− gH − gH =0 (5.19)
∂t ∂x ∂x ∂y ∂y

which in the particular case of constant water depth H transforms to a Laplace equation:

∂2S 2
∂2S
!
2 ∂ S
= c − (5.20)
∂t2 ∂x2 ∂y 2

where c = (gh)1/2 is the celerity of long waves.


The equation (5.20) has simple solutions, if there is no dependence on y. In this case for
the initial free surface displacement being a function of x only, S = F (x) and the fluid
being initially at rest, the time-dependent solution of (5.20) is:
1
S = [F (x + ct) + F (x − ct)] (5.21)
2

and the velocity from (5.17):


1
u = − c−1 g[F (x + ct) − F (x − ct)] (5.22)
2

Two important examples can illustrate the behaviour of the hydrostatic solution of the
lock exchange flow. Both describe propagation of an initial displacement of the free
surface in a form of a step.
In case (a) the initial free surface displacement is described as a step function:

F (x) = η0 sgn(x) (5.23)

where sgn is the sign function:


(
1, x>0
sgn(x) = (5.24)
−1, x < 0

In this case a pair of wavefronts of height η0 are produced.


In case (b) the initial free surface perturbation is confined to a limited area x ∈ [−L, L]:
(
η0 , |x| < L
S(x) = (5.25)
0, |x| > L
146 5.4 Lock exchange flow

z
(a)

2η0 x

c @
-c
@ x
η0

z
(b)

η0
x
@ @
-c c
c @ @ -c
η0 /2
@ @ x

Figure 5.10: Solutions of the shallow water wave equations for (a) initial disturbance in
the form of a single step function, (b) in the form of a confined displacement. Descriptions
in the text.

producing two pairs of wavefronts.

Both cases are shown schematically in fig. 5.10. These examples are especially impor-
tant for understanding qualitatively the solution of the lock exchange problem in the
hydrostatic case.

In case (a) two wavefronts of relative height η0 move out in different directions from the
position of the initial displacement with a celerity c = (gη0 )1/2 , leaving behind a region
confined between x = −ct and x = ct with zero displacement, but with a steady motion
with velocity u = c−1 gη0 directed to the right (from equation 5.22). In consequence
u = c.

In case (b) two pairs of wavefronts are produced, moving in opposite directions away
from the positions of the initial free surface displacement. The velocity is zero except in
those regions, where the surface elevation is η0 /2, and movement is directed away from
the axis of symmetry, where a wall can be placed without influencing the movement.

The characteristic rectangular shape of the interface is therefore a consequence of the


solutions shown in fig. 5.10. If the lock exchange problem is solved using hydrostatic
approximation, this shape of the interface should be observed.
5.4 Lock exchange flow 147

5.4.2 Lock exchange flow speed


The velocity with which the buoyancy-driven current advances can be estimated from
energy conservation considerations.
When the wall between two equal volumes of fluids of different density in a basin is
removed, the instability causes a movement of the volumes toward a position of equi-
librium, where a layer of lighter fluid lies above the denser one. Assuming the flow to
be conservative in terms of the mechanical energy, the velocity of the motion can be
estimated from the available potential energy being the difference between the potential
energy of the initial state and the equilibrium.
Assuming the base of the container (basin) to be a square of a width L and the water
depth H (figure 5.11), the potential energy of the initial state per unit area of the base
is:
1
Z L/2 Z H
−1
Ei = L ̺gzdxdz = (̺1 + ̺2 )gH 2 (5.26)
−L/2 0 4
where ̺1 and ̺2 are densities of both fluids and the reference level for the potential
energy is the bottom of the basin with the origin of the co-ordinate system at the
interface. Computing the energy of the equilibrium state in the same way:
1 3 1
Ee = ( ̺1 + ̺2 )gH 2 (5.27)
4 2 2
The available potential energy is:
1
∆E = Ei − Ee = (̺2 − ̺1 )gH 2 (5.28)
8
If all this energy can be converted into kinetic energy, the mean velocity of the movement
will be equal to:
1 ̺2 − ̺1
u2 = gH (5.29)
2 ̺1 + ̺2
The above equation can be used to estimate approximately the velocity of the flow.
Analogical treatment can be given a fully developed and therefore stationary lock ex-
change flow, when a perfect symmetry between the two layers of lighter and denser fluids
of equal thickness can be assumed (figure 5.9). In a time interval ∆t, the wedges ad-
vance a distance of u∆t. In this time the lighter fluid gains potential energy amounting
to ̺1 g(H/4)u∆t(H/2) per width unit of the channel; simultaneously, the heavier fluid
loses energy amounting to ̺2 g(H/4)u∆t(H/2). The kinetic energy gained by both fluids
is (1/2)(̺1 + ̺2 )u2 (H/2)u∆t. Again, when all available potential energy can be changed
to the kinetic form, formula (5.29) is obtained [167] again:
√ s s s
2 ̺2 − ̺1 ̺2 − ̺1 ̺2 − ̺1 q
u= gH = 0.71 gH ≈ 0.5 gH = 0.5 g ′H (5.30)
2 ̺1 + ̺2 ̺1 + ̺2 ̺2

where g ′ = g∆̺/̺2 is the reduced gravity acceleration. This result should be compared
with the formula for long, hydrostatic interfacial wave celerity (5.7) – it is virtually the
same. This confirms also the considerations from the section 5.4.1.
148 5.4 Lock exchange flow

 L -

6
̺1 ̺1
aa
̺2 ̺1 aa
H aa
aa
aa
̺2 ̺2

(i) (ii) (iii)


Figure 5.11: Initial and equilibrium state for the lock exchange problem.

5.4.3 Lock exchange flow – further considerations


Numerous measurements have been made for the flow generated by removal of a vertical
barrier separating two layers of equal depth and slightly different densities in completely
enclosed conduits and with channels with a free surface [4, 167]. The laboratory exper-
iments of Yih made in 1947 yield the stationary flow velocity as given by the formula
(5.30) where instead of 0.71 the √ factor 0.67 appears [167]. Barr observed in a closed
conduit the velocity of u = 0.44 g√ ′ H. When the free surface is present, the flow is

slightly asymmetrical with u = 0.47 g ′H for the underflow and u = 0.59 g ′H for the
overflow.
Von Kármán [88] and Benjamin [7] computed the interface profiles showing the charac-
teristic nose angle of π/3. Benjamin discusses the case of lock exchange flow showing
that only in the case of a flow where both wedges have identical geometries a steady non-
dissipative flow can occur; otherwise mixing, waves, and/or internal hydraulic jumps can
occur. The flow speed can be estimated from considerations concerning the mechanical
energy conservation in the case of equal depths only, as presented in section 5.11.

5.4.4 Test realisation


The comparisons are carried out in two computational settings which correspond with
both cases shown in fig. 5.10, in a channel-like basin.
The salinity is taken as the (transported) active tracer influencing fluid density. The
dependence of the density ̺ [kg/m3 ] on the salinity s [psu] is assumed to be linear
(appendix section A.4.2):

̺ = 999.972 · (1 + 0.75 × 10−3 s) (5.31)

The salinity is initially so distributed, that the left half of the basin is occupied by saline
water with s = 1 psu, and the right one by fresh water (s = 0 psu). The density of
both fluids and the mean reference density are 999.972 kg/m3 , 1000.722 kg/m3 , and
1000.347 kg/m3 respectively. In the initial condition the velocity in the whole domain
5.4 Lock exchange flow
salinity
0

Figure 5.12: Lock exchange problem. The results of the hydrostatic model. -1

s [PSU]

z [m]
-2
0.90
0.70
-3 0.50
0.30
0.10
-4
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30
x [m]

tangential velocity component


0

-1
u [m/s]
z [m]

-2 0.100
0.060
0.020
-3 -0.020
-0.060
-0.100
-4
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30
x [m]

vertical velocity component


0

-1

w [m/s]
z [m]

-2
0.200
0.100
-3 0.000
-0.100
-0.200
-4
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30

149
x [m]
150
salinity
0

Figure 5.13: Lock exchange problem. The results of the non-hydrostatic model.
-1

s [PSU]

z [m]
-2
0.90
0.70
-3 0.50
0.30
0.10
-4
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30
x [m]

tangential velocity component


0

-1
u [m/s]
z [m]

-2 0.100
0.060
0.020
-3 -0.020
-0.060
-0.100
-4
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30
x [m]

vertical velocity component


0

5.4 Lock exchange flow


-1

w [m/s]
z [m]

-2
0.040
0.020
-3 -0.000
-0.020
-0.040
-4
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30
x [m]
model.
Figure 5.14: Lock exchange problem with two interfaces. The results of the hydrostatic

5.4 Lock exchange flow


salinity
0

-1

s [PSU]
z [m] -2
0.90
0.70
-3 0.50
0.30
0.10
-4
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30
x [m]

tangential velocity component


0

-1
u [m/s]
z [m]

-2 0.100
0.060
0.020
-3 -0.020
-0.060
-0.100
-4
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30
x [m]

vertical velocity component


0

-1

w [m/s]
z [m]

-2
0.200
0.100
-3 0.000
-0.100
-0.200
-4
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30

151
x [m]
hydrostatic model.
Figure 5.15: Lock exchange problem with two interfaces.

152
salinity
0

-1

s [PSU]

z [m]
-2
0.90
0.70
-3 0.50
0.30
0.10
-4
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30
x [m]

tangential velocity component


0

-1
u [m/s]
z [m]

-2 0.100
0.060
0.020
-3 -0.020
-0.060
-0.100
-4
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30
x [m]

vertical velocity component


0
The results of the non-

5.4 Lock exchange flow


-1

w [m/s]
z [m]

-2
0.040
0.020
-3 -0.000
-0.020
-0.040
-4
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30
x [m]
5.4 Lock exchange flow 153

is set to zero, u = 0. At all lateral boundaries and at the bottom impermeability and
a free slip boundary condition are set and no flux of salinity through the boundary is
imposed. The turbulent diffusion and viscosity coefficients are set to zero (inviscid flow).
The free surface is computed from long-wave equations in the hydrostatical case. In the
non-hydrostatic case, the stabilised conservative free surface equation is applied.

The channel measures 30 m in length, 3 m in width and 4 m depth. The basic mesh
has 2416 nodes and 4500 elements, horizontal and vertical resolution is 0.2 m, 21 mesh
levels are used (figure A.2). The time-step is 1 s. The results are presented in figures
5.12 and 5.13 as vertical cuts through the domain in the x-direction for t = 100 s. The
differences in the interface shapes and velocity distribution are evident. The charac-
teristic rectangular shape is observed with the hydrostatic approximation, whereas the
non-hydrostatic solution yields a much more realistic density flow head shape.

The domain is long enough for a stationary flow to establish in the middle of the domain
when the heads of the wedges move away in opposite directions. For this setting the
theory presented in the sections above can be applied in order to check the velocity of the
flow. For the geometry and the densities of the fluids as mentioned above, the formula
(5.30) yields a velocity of ca. 0.099 m/s. The time series of the horizontal velocity in
points at the free surface and at the bottom and situated exactly in the middle of the
domain (where the initial interface is situated), are shown in figure 5.16, for the non-
hydrostatic and the hydrostatic cases. In the hydrostatic case a symmetric stationary
flow develops, but the velocity is about 20% lower than theoretically expected. No free
surface wave develops. In the non-hydrostatic case, the flows is also symmetric but with
flow speed 10% lower as expected. The buoyancy-driven flow is modulated by a free
surface standing wave motion with a period of about 9.85 s with a very small amplitude
of 3 mm.

The differences between hydrostatic and non-hydrostatic solutions are also evident in
the case where, in the initial condition, the denser fluid occupies the central part of the
channel between x = 7.5 m and x = 12.5 m. In this case one has two interfaces moving
in different directions, figures 5.14 and 5.15 (integration time 50 s, time-step 1 s). The
current will not change, if a wall is placed exactly in the middle of the channel.

A good illustration of the different flow patterns in both the hydrostatic and non-
hydrostatic cases is shown in figure 5.17. The computational domain is a cube 10 m
× 10 m × 10 m; the basis mesh is shown in figure A.1. The time step is ∆t = 0.05 s,
and the presented figure shows the flow at t = 30 s. The extreme difference between the
vertical velocity component value as well as in the interface position is evident.

As a conclusion, for proper simulation of density flows, the non-hydrostatic model should
be preferred, because it yields a realistic interface shape as well as a more acceptable
wedge intrusion speed. Additionally, the velocity distribution nearby the nose of the
intruding dense flow is described much more realistically. As in the case of interfacial
internal waves, improvements can be thought of which reduce the mixing at the interface.
154 5.4 Lock exchange flow

0.12

0.1

free surface
0.08
| u | [m/s]

0.06

0.04
bottom

0.02

0.0
0 20 40 60 80 100 120 140 160 180 200
t [s]

0.12

0.1
free surface

0.08
| u | [m/s]

0.06
bottom

0.04

0.02

0.0
0 20 40 60 80 100 120 140 160 180 200
t [s]

Figure 5.16: The lock exchange flow wedge speed time series for points at the free surface
and at the bottom and situated exactly in the middle of the domain. Non-hydrostatic
(above) and hydrostatic (below) solutions with free surface. The theoretical value for
steady flow is 0.099 m/s.
5.5 Standing wave in a closed basin 155

0 0
z [m]

z [m]
v [m/s]
0.2

-10 -10
0 10 0 10
x [m] x [m]

Figure 5.17: Flow patterns in a cubic domain 30 s after the initialisation; left: the non-
hydrostatic solution, right: the hydrostatic one. Solid line: the interface between the
fluids.

5.5 Standing wave in a closed basin

A small amplitude uni-nodal standing wave in a closed basin is a particularly good


example for testing mass and energy conservation when computing transient flows. The
test realisation is very simple, but the model properties to be tested are not trivial.
As the initial condition a sloped free surface and no motion is taken – a configuration
with maximum potential energy. A wave motion begins – and in the moment, when the
free surface is horizontal, the whole available potential energy has been exchanged into
kinetic one. When the highest water level at the opposite wall is reached, the maximum
potential and the minimum kinetic energy state is achieved again, just as at the very
beginning of the simulation. When there are no energy and mass losses (a free slip BC
for velocity at all boundaries, no shear, no viscosity), such an oscillation in a numerical
experiment should continue forever. Additionally, when the simulation is performed in a
square basin with free surface slope parallel to one of the sides (i.e. motion takes place in
a 2D vertical plane, figure 5.18), there should not be any parasitic motion perpendicular
to the wave slope.

An analytical solution of this wave problem exists. The resulting velocity patterns de-
pend on the water depth, and the role of the vertical accelerations can be immediately
seen in a situation corresponding to short (deep water) waves. Moreover, the deviations
of the numerical result from an ideal movement can be especially well observed. The
behaviour of the hydrostatic and non-hydrostatic models for short and long waves cases
and various methods of computing the free surface movements can be compared clearly.
156 5.5 Standing wave in a closed basin

z
6

η(x, y, t)
0 η0

–10 -
0 10 x

Figure 5.18: A uni-nodal standing wave in a closed basin.

5.5.1 Theory
The theory of the small amplitude progressive waves is shortly sketched in the appendix
section A.3. A standing wave is caused by the superposition of two waves of the same
period and amplitude η0 /2, but travelling in different directions:

1
η(x) = η0 [cos(kx − ωt) + cos(−kx − ωt)] = −η0 cos(kx) cos(ωt) (5.32)
2

The inviscid fluid of constant density is confined in a closed basin with a square base
L × L and with equilibrium depth H. z = 0 at the equilibrium surface level. The initial
free surface elevation is given by:

η(x) = η0 cos(kx) 0 < x < L (5.33)

where k = 2π/nL, n = 2 for a uni-nodal standing wave. The frequency ω of the


resulting wave is given by positive root of ω 2 = gktanh(kH), and the celerity of the
wave is c = ω/k. According to the small amplitude wave theory (η0 << H) describing
the resulting wave motion in the plane (x, z), the solutions for velocity potential ϕ,
free surface elevation η, velocity components (u, v, w) and perturbation pressure p′ are
[99, 100]:
ch[k(z + H)]
ϕ = cη0 cos(kx) sin(ωt) (5.34)
ch(kH)
η = η0 cos(kx) cos(ωt) (5.35)
ch[k(z + H)]
u = ωη0 sin(kx) sin(ωt) (5.36)
sh(kH)
sh[k(z + H)]
w = −ωη0 cos(kx) sin(ωt) (5.37)
sh(kH)
5.5 Standing wave in a closed basin 157

ch[k(z + H)]
p′ = ̺gη0 cos(kx) cos(ωt) (5.38)
ch(kH)

(compare with section A.3) The perturbation pressure is defined as the difference between
the global pressure and the equilibrium pressure p′ = p − p0 = ̺gη + π. Therefore, the
hydrodynamic pressure π is given by:

ch[k(z + H)]
π = −̺gη + ̺gη0 cos(kx) cos(ωt) (5.39)
ch(kH)

The proper initial conditions for t = 0 are:

η(x) = η0 cos(kx)
u=0 (5.40)
w=0
ch[k(z + H)]
π = −̺gη + ̺gη0 cos(kx)
ch(kH)

The solutions given above are valid for a wave in a basin of an arbitrary depth. For the
long wave (shallow water, H << L) the celerity c = (gH)1/2, the frequency ω = k(gH)1/2
and the following approximations are valid:
ωη0
u= sin(kx) sin(ωt) (5.41)
kH
z+H
w = −ωη0 cos(kx) sin(ωt) (5.42)
H
p′ = −̺gη0 cos(kx) cos(ωt) (5.43)
π=0 (5.44)

5.5.2 Deep water – short waves


In a closed square pool of 10 m width and 10 m length an initial slope of the free
surface in the form of (5.33) and a zero initial velocity is assumed. The initial condition
corresponds therefore to the position of maximum potential energy and minimum kinetic
energy. The equilibrium water depth is 10 m and the amplitude is set to η0 = 0.1 m, 1%
of the water depth, fulfilling the assumptions of the small amplitude waves theory. The
analytical solution of this problem is given in the previous section (5.5.1).
For simplicity, g = 10 m2 /s and ̺ = 1000 kg/m3 , so that for the equilibrium water depth
10 m, the long wave celerity (from equation A.42) is exactly c = (gh)1/2 = 10 m/s and
the period of oscillations is T = 2 s. However, the long-wave theory is not appropriate
in this case, because the water depth is comparable with the wave length. Applying the
general small amplitude wave theory, c = [(g/k) tanh(kh)]1/2 = 5.64 m/s and the period
of oscillations is 3.55 s (from equation A.34 or A.37).
158 5.5 Standing wave in a closed basin

numeric v [m/s]
analytic v [m/s]
0.05 0.05

0 0 0 0 0 0

-2 -2 0
50 0
25
50 25
-2 -2 -2 -2

-4 -4 -4 -4

z [m]
-500

-500

500
500

-6 -6 -6 -6

0
0
-750

-750
-250

750
-250

250
250

750

-8 -8 -8 -8

-10 -10 -10 -10


0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10
x [m] x [m]

Figure 5.19: Comparison between numeric and analytic solution for t = 0.1 s. Iso-lines
of dynamic pressure shown each 50 Pa.

The basic mesh is shown in figure A.1. The time step is varied between 0.01 s and
0.1 s. Inviscid flow is assumed. The free surface is computed from the kinematic BC and
the stabilised SUPG method is applied, when not stated otherwise. For velocity, at the
impermeable boundaries impermeability or free slip conditions are given. At the free sur-
face, the zero Neumann boundary condition is applied for all three velocity components.
Hydrodynamic pressure is set to zero at the surface and impermeability conditions are
imposed at all impermeable walls. The assumption of zero dynamic pressure is the only
difference between the boundary conditions applied in the numerical model and in the
analytical solution (see section 5.5.1).
A comparison between the analytical and numerical solutions is given in vertical sections,
figures 5.19, 5.20 and 5.21, for t =0.1 s, 5 s and 10 s respectively. All show a very good
agreement of the numerical results with the analytical theory of short waves. The greatest
differences occur in the pressure field, due to differences in the BC at the free surface.
In order to demonstrate the behaviour of the solution after several periods of oscillation,
time series for the free surface elevation at a boundary edges x = 0 m and x = 10 m
for 30 s simulation time (over 8 periods, ∆t = 0.05 s) are presented in figure 5.22. The
amplitude remains for a considerably long time constant and the period of oscillation
gets only slightly longer: the time obtained for 8 full periods is about two time steps
(0.1 s) longer than the theoretical one. The mass loss is about 0.03%.
A very illustrative example is to compare the results of the hydrostatic and non-
hydrostatic model for the same phase of motion, e.g. when the state of maximum kinetic
energy is achieved, for one-fourth of the oscillation period. The results are shown for
5.5 Standing wave in a closed basin 159

numeric analytic
v [m/s] v [m/s]
0.2 0.2

0 0 0 0 0 0

25 -2
0 50 50 0
-2 -2 -2 -2 25 -2

-4 -4 -4 -4

z [m]

-500

500
-500
500

-6 -6 -6 -6

0
0

-250
-250

250
250

-8 -8 -8 -8

-750

750
-10 -10 -10 -10
0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10
x [m] x [m]

Figure 5.20: Comparison between numeric and analytic solution for t = 5 s. Iso-lines of
dynamic pressure shown each 50 Pa.

numeric analytic
v [m/s] v [m/s]
0.2 0.2

0 0 0 0 0 0

-2 -2 -2 -2
0

-4 -4 -4 -4
z [m]
-250
250

-250
250

-6 -6 -6 -6
0
0

-8 -8 -8 -8

-10 -10 -10 -10


0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10
x [m] x [m]

Figure 5.21: Comparison between numeric and analytic solution for t = 10 s. Iso-lines
of dynamic pressure shown each 50 Pa.
160 5.5 Standing wave in a closed basin

0.0 3.552 7.104 10.656 14.208 17.76 21.312 24.864 28.416

0.1 0.1

0.05 0.05
z [m]

0.0 0.0

-0.05 -0.05

-0.1 -0.1

0 5 10 15 20 25 30
t [s]

Figure 5.22: Time series for the free surface elevation at a boundary edge x = 0 m and
x = 10 m (chain-dotted line) for 30 s simulation time. The oscillation period is 3.552 s,
shown along the upper x-axis.

0.9 s for the non-hydrostatic and 0.5 s for the hydrostatic one. The difference between
both solutions are striking (figure 5.23) – the hydrostatic model yields (unphysical in
this case) long-wave velocity profiles compared to the exponentially diminishing ones for
the (physical) non-hydrostatic case.
Additionally, in figure 5.24 time series for surface elevation at the basin boundary with
and without hydrostatic approximation are shown. The hydrostatic model reproduces
perfectly the (unphysical) long wave period, although it does not conserve energy – the
amplitude is almost completely damped after 30 s. On the other hand, the mass (volume)
conservation is excellent (negligible, 2.6×10−4 m3 in a tank containing 1000 m3 ). Similar
remarks can be made for the strongly damped dashed line in figure 5.24, representing
the non-hydrostatic solution with application of the vertically integrated momentum
equations for obtaining the free surface elevation and taking barotropic pressure gradients
into consideration, as described in section 2.8.8. Although the mass is conserved and the
wave period is reproduced properly, the energy is not conserved again.

5.5.3 Shallow water – long waves

In this verification case the setting is virtually the same, but the computational mesh
represents now a long channel. The mesh shown in fig. A.2 is used. The channel depth
is the same as of the pool, 10 m, width 15 m and length 150 m. This is approximately
5.5 Standing wave in a closed basin 161

hydrostatic non-hydrostatic
v [m/s] v [m/s]
0.2 0.2

0 0 0 0

-2 -2 -2 -2

-4 -4 -4 -4

z [m]
-6 -6 -6 -6

-8 -8 -8 -8

-10 -10 -10 -10


0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10
x [m] x [m]

Figure 5.23: Comparison of solutions with (left) and without (right) the hydrostatic
approximation, for the same phase of motion, T /4, equal 0.5 s and 0.9 s, respectively.

the limit case for the application of the hydrostatic long wave theory, i.e. the ratio wave
length to the depth 20:1. For the given channel it is 30:1. The initial free surface is
given by (5.33) and a zero initial velocity is assumed. The amplitude is again η0 = 0.1 m
and all other parameters are the same as before. Applying the general small amplitude
wave theory to this case, the wave phase speed is c = [(g/k) tanh(kh)]1/2 = 9.93 m/s
and the period of oscillations of 30.22 s. Long wave theory yields c = (gh)1/2 = 10 m/s
and the period of oscillations T = 30.0 s, so it can be assumed that the hydrostatic
approximation is almost entirely appropriate in this case.

Figure 5.25 compares the free surface profile development over time for about 10 peri-
ods. The non-hydrostatic model reproduces the period of 30.22 s perfectly, while the
hydrostatic one (only the pressure-continuity step switched on) yields exactly 30 s from
the long wave (hydrostatic) theory. While the hydrostatic model shows large numerical
diffusion (by the 9th period the amplitude is only 18% of the initial position), the non-
hydrostatic one for the chosen time-step of 0.1 s is slightly unstable (see next section),
yielding an increase in amplitude to 111% by the 9th period.

5.5.4 Comparison of different free surface algorithms

In this section different methods of computing the surface for the case of standing waves
are compared.
162 5.5 Standing wave in a closed basin

z [m]

-0.15

-0.05

0.05

0.15
-0.1

0.0

0.1

0.0
0
2

3.552
4
6

7.104
8
10

10.656
12

14.208
14
t [s]
16

17.76
18
20

21.312
22
24

24.864
26

28.416
28
30
-0.15

-0.1

-0.05

0.0

0.05

0.1

0.15

Figure 5.24: Time series for the free surface elevation at a boundary edge x = 0 m. The
oscillation period for the non-hydrostatic solution (chain-dotted line) is 3.552 s, for the
hydrostatic one (solid line) 2.0 s. The dashed line: the non-hydrostatic solution with
vertically integrated momentum equations for obtaining the free surface elevation.
5.5 Standing wave in a closed basin 163

non-hydrostatic hydrostatic
t [s] t [s]
300 300 300
0.00 0.00 0.00
-0.05
0.05
0.05 0.00 0.00
272.0 0.10
270 270 270
0.00 0.00
-0.05
0.05
0.05 0.00 0.00
241.7 240 240 240
0.00 0.00
-0.05
0.05
0.05 0.00 0.00
211.5 210 210 210
0.0 0.00
-0.05 0
0.05
0.05 0.00 0.00
181.3 180 180 180
0.0 0.00
-0.05 0
0.05
0.05 0.00 0.00
151.1 150 150 150
0.0 0.00
-0.05 0
0.05
0.05 0.00 0.00
120.9 120 120 120
0.0 0.00
-0.05 0
0.05 0.00
0.05 0.00
90.7 90 90 0.05 -0.05 90
0.0 0.00
-0.05 -0.05
0
0.05 0.00 0.00
0.05 0.05
60.4 60 60 60
0.00 0.00
-0.05 -0.05
0.05 0.00 0.00
30.2 0.05 0.05
-0.05 30 30 30
0.00
-0.05 -0.05
0.00 0.00 0.05
0.05 -0.05 0.05 0.00
0.0 0 0 0
0 30 60 90 120 150 0 30 60 90 120 150
x [m] x [m]

Figure 5.25: Free surface oscillations for a channel of 150 m length and 10 m depth.
Comparison between the non-hydrostatic and hydrostatic solutions. The changes of the
free surface profile in the middle of the pool are shown as a function of pool length and
time. The oscillation period for this wave is 30.2179 s, in the long wave approximation
exactly 30.0 s.
164 5.5 Standing wave in a closed basin

0.0 3.552 7.104 10.656 14.208 17.76


0.15 0.15

0.1 0.1

0.05 0.05
z [m]

0.0 0.0

-0.05 -0.05

-0.1 -0.1

-0.15 -0.15
0 2 4 6 8 10 12 14 16 18 20
t [s]

Figure 5.26: Time series for the free surface elevation at a boundary edge x = 0 m for
20 s simulation time and different time steps: 0.01 s (solid line), 0.05 s (chain-dotted
line) and 0.1 s (dashed line). The oscillation period is 3.552 s.

5.5.4.1 Dependency on the time step

First of all, not only the accuracy in reproducing the free surface oscillation period, but
also the conservation of momentum depends on time step, as shown in figure 5.26. The
computation has been performed using kinematic BC for the free surface with implicit
SUPG method. The numerical result converges to the analytical one with diminishing
time step. As expected for an operator-splitting algorithm with an overall explicit time
discretisation, the consistency order of the technique is one and the accuracy depends
approximately linearly on the time step.

5.5.4.2 Free surface stabilisation technique

The stability of the free surface algorithm (based on the kinematic boundary condition)
for larger time steps can be enhanced using the stabilisation method described in section
3.6.5. Figure 5.27 shows the effects of stabilising the resulting free surface by filtering
once and twice the results obtained by implicit SUPG method for ∆t = 0.05 s. This
smoothing technique also brings some advantages by reducing the needed overall CPU
time by 10%.

5.5.4.3 Comparison of the implemented algorithms

The properties of various free surface computation techniques are compared briefly in
figure 5.28 for ∆t = 0.05 s. By 1000.0 m3 initial volume, the stabilised implicit methods
5.6 Subcritical and supercritical flow over a ramp 165

0.0 3.552 7.104 10.656 14.208 17.76


0.15 0.15

0.1 0.1

0.05 0.05
z [m]

0.0 0.0

-0.05 -0.05

-0.1 -0.1

-0.15 -0.15
0 2 4 6 8 10 12 14 16 18 20
t [s]

Figure 5.27: Time series for the free surface elevation at a boundary edge x = 0 m for
20 s simulation time and different filtering: none (dashed line), once (chain-dotted line)
and twice (solid line). ∆t = 0.05 s.

based on kinematic BC yield comparable results by volume loss at t = 30 s of −.094 m3


for SUPG, and growth of 0.27 m3 for the method of characteristics. The conservative
equation method yields excellent volume conservation with a negligible mass increase,
0.74 × 10−11 m3 , whereby the fully explicit formulation (section 3.6.3) the momentum
is not conserved – the amplitude diminishes in 30 s to ca. 65% of its initial value. By
the implicit formulation (section 3.6.4), the results obtained with conservative equation
are entirely comparable with the ones obtained with kinematic BC, with mass losses of
0.04 m3 .
The overall computation times are comparable when applying of the method of char-
acteristics or the explicit conservative free surface equation. When applying the SUPG
method it is 5–10% greater. The implicit conservative free surface equation, based also
on SUPG, needs less equation solver iterations yielding about 10% lower computation
times then the equivalent kinematic BC method. Surprisingly in the light of test results
mentioned in section 4.3.1, for the range of applied time-steps, the results do not depend
on the implicitness of the free surface algorithms based on kinematic BC.

5.6 Subcritical and supercritical flow over a ramp


The channel flow has two characteristic regimes, namely subcritical and supercritical
regimes. The characteristic number controlling these regimes is the Froude number:
|u| |u|
Fr = √ = (5.45)
gH c0
166 5.6 Subcritical and supercritical flow over a ramp

0.0 3.552 7.104 10.656 14.208 17.76


0.15 0.15

0.1 0.1

0.05 0.05
z [m]

0.0 0.0

-0.05 -0.05

-0.1 -0.1

-0.15 -0.15
0 2 4 6 8 10 12 14 16 18 20
t [s]

Figure 5.28: Time series for the free surface elevation at a boundary edge x = 0 m for
20 s simulation time and different implicit free surface computation methods. Kinematic
BC: characteristics (dashed line), SUPG (solid line). Conservative free surface equation:
fully explicit (chain-dotted line), fully implicit (dotted line).


where H is the channel depth and c0 = gH the long free surface wave phase speed.
The supercritical regime is obtained, when F r > 1 and subcritical one for F r < 1 [171].
One of the features of the supercritical flow is the fact, that the flow velocity u is greater
than the greatest possible velocity of the free surface waves c0 and waves cannot travel
upstream. In the subcritical case, the waves can move downstream and upstream as well.
As a consequence, the inviscid supercritical channel flow requires very simple boundary
conditions: an imposed fluid velocity and depth at the inflow boundary and free outflow
boundary conditions at the outflow.
As an example of the free surface flow a supercritical flow over a ramp is chosen, with
the geometry shown in figure 5.29. The slope of the ramp is 1:5, i.e. outside the validity
domain for shallow water equations. The applied mesh is shown in figure A.2, the spacing
is ∆x = 0.1 m. The inviscid free surface flow can be described by the Bernoulli equation
valid along the streamline following the free surface from inflow and outflow sections and
continuity equations stating mass conservation between these sections (where the flow is
homogeneous):

u2
+ h + z = const
2g
uh = q = const (5.46)

For the supercritical case, as the initial condition, a free surface level of 1 m and an
initial velocity of 1 m/s is given everywhere. At the inflow boundary Dirichlet BC for
5.6 Subcritical and supercritical flow over a ramp 167

1.2 [Pa]
1.0 150
125
0.8 100
z [m]

75
0.6 50
25
0
0.4
-25
-50
0.2 -75
-100
0.0 -125
-150
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
channel length [m]

Figure 5.29: The steady state (t = 100 s) hydrodynamic pressure (π) distribution for
supercritical channel flow over a ramp with 1:5 slope (a vertical section along the chan-
nel).

all flow variables are imposed (free surface level of 1.0 m and a constant velocity of
u = 6 m/s) and zero dynamic pressure. At the outflow boundary, a zero Neumann BC
for all variables is applied. The fluid speed is almost two times greater than the wave
celerity c = 3.16 m/s (g = 10.0 m/s2 for simplicity). Solution of (5.46) for a given
constant flux q = 6 m2 /s (i.e divided by channel width) yields a value for the water
depth downstream from the ramp 1.0928 m and (for the water level 1.2928 m).
Free surface is computed with the semi-implicit SUPG algorithm with θ = 0.55. The
stabilisation of the free surface section in the middle of the channel in time for 10 s and
with time-step ∆t = 0.01 s is shown in figure 5.30. The initially occurring wave travels
out of the domain without causing any problems. The computed free surface level at the
outflow is 1.2957 m, relative error to the analytical solution ≈ 0.22%. A simulation with
∆t = 0.1 s for 100 s brings steady state water level of 1.2971 m, relative error ≈ 0.33%.
In the case of the subcritical flow, the waves generated over the ramp due to the inap-
propriate initial condition can travel in both directions. When the boundary conditions
for the stabilised channel flow are applied, e.g. Dirichlet boundary conditions for all
variables, or ∂h/∂x = 0 and q = const at the inflow section and a constant pressure
at the outflow, the generated waves are reflected at the inflow and outflow sections and
disperse over the ramp. When the algorithm does not provide a method to damp these
oscillations or provide non-reflecting or absorbing boundary conditions, these waves can
finally lead to breakup of the program.
The computation for the subcritical flow over a ramp is performed for the initial velocity
of 1 m/s everywhere and a flat free surface level of 1.0 m again. As the boundary
conditions at the outflow boundary a constant dynamic pressure of 0 and a free outflow
BC for all other variables are applied. The inflow BC are ∂h/∂x = 0 with the velocity
changing according to the water depth so that a constant flux q = 1.0 m2 /s is provided.
As previously, the inviscid flow is considered, whereby the free surface is computed with
stabilised implicit SUPG method in order to provide short wave damping. Without
damping the solution breaks up when the initial waves are reflected from boundaries.
168 5.7 Channel with a bump

10
1.00
9

8
1.00
7

6 1.00
time [s]

4 00
1.
0
1.3

1.00
3

00

1.20
2
1. 5 1.1
5
1.2
1
1.10
0 1.05
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
length [m]

Figure 5.30: Stabilisation of the free surface profile for the supercritical flow over a ramp.

The stabilisation of the free surface profile in time is shown in figure 5.31 for simulation
time 100 s and ∆t = 0.1 s. The initially occurring waves are partially reflected from the
boundaries, partially leave the domain, and finally are dissipated.
The comparison between steady state free surface profiles is given in figure 5.32. For
the subcritical flow the inflow water level stabilises at 1.12134 m and inflow velocity
0.891828 m/s, yielding q = 1.00004241 m2 /s. The outflow water level stabilises at
1.09897 m and inflow velocity 1.11229 m/s, yielding q = 0.99991534 m2 /s. This is an
excellent agreement with theory – equations (5.46) yield a free surface level of 1.09927 m.
The relative error is ≈ 0.03%.
This tests shows clearly the ability of the newly developed model to reproduce super-
critical and subcritical channel flows over a bottom topography with slopes excluding
application of codes based on shallow water equations. The limits of application are
discussed in section 5.10.2. Additionally, the stabilised method of computing the free
surface is tested.

5.7 Channel with a bump


A good test case for the properties of the free surface algorithm is a subcritical stationary
flow in a long channel with a bump in the middle of its length. For an inviscid case, the
free surface profile can be computed easily from (5.46).
As an initial condition a constant velocity and a flat free surface is usually assumed and
computation is continued until a stable state is achieved. This artificial initial condition,
5.7 Channel with a bump 169

100

90

80

70

60
time [s]

50

40

30

1.100
20

10 1.075

1.025 1.050 1.025


0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
length [m]

Figure 5.31: Stabilisation of the free surface profile for the subcritical flow over a ramp
steered with constant inflow flux.

1.4

1.3
water level [m]

1.2

1.1

1.0

0.9
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
length [m]

Figure 5.32: The steady state (t = 100 s) free surface profiles for supercritical (solid line)
and subcritical (dashed line) flows over a ramp.
170 5.7 Channel with a bump

-2

-4
z [m]

-6
u [m/s]
-8 1
0.1
-10
0 50 100 150 200 250 300 350 400 450 500 550 600
length [m]

Figure 5.33: A channel with a bump. Vertical section, steady state velocity.

for a subcritical flow, produces free surface waves travelling downstream and upstream
as well. The boundary conditions set at both open boundaries at the extremities of
the channel are crucial for the stabilisation of the solution. The usual method is to
impose boundary conditions that are appropriate for the targeted steady state. In this
case, for a model computing the free surface form shallow water equations, they are a
constant flux at the inflow open boundary and a constant water level at the outflow
boundary. As a result, the waves occurring initially directly over the bump travel in
both directions, partially leaving the domain, and partially being reflected backwards.
The numerical diffusion dampens their amplitude, as well, and after some time the final
state is achieved.
For a 3D non-hydrostatic model there are various methods of setting the boundary
conditions. For example, by imposing at both open boundaries a zero Dirichlet BC for the
pressure and free outflow boundary conditions for the free surface and velocity, reflecting
boundary conditions are obtained. This causes the waves induced by the artificial initial
condition to bounce around the domain until they are dissipated – or cause numerical
instabilities. Another method is to set absorbing or non-reflecting boundary conditions
at the open boundaries to let the incoming waves be absorbed or leave the domain.
The test case is realised in a channel of depth H = 10 m, length l = 600 m and width of
B = 28 m, where a bottom profile (a bump) given by the formula:
!
2 x − 0.5(l − l0 )
zb = A sin (5.47)
l0

with a height A = 3 m and a length of l0 = 200 m in centred in the middle (i.e. between
xa = 200 m and xb = 400 m) is localised (figure 5.33, the applied basic mesh shown
in figure A.3). For this case, when a constant flux of 10 m2 /s is given at the inflow
boundary, the Bernoulli equation (5.46) yields for a stationary flow a slight submersion
of the free surface profile with max. 0.0536 m discrepance from the initial zero level just
above the highest point of the bump.
The computations with a non-hydrostatic model are performed with a stabilised SUPG
algorithm based on the kinematic BC for the free surface. For the advection step the
method of characteristics is chosen. The flow is assumed to be inviscid. At channel
5.7 Channel with a bump 171

lateral walls, zero Neumann BC is applied. For open boundaries, two types of boundary
conditions are chosen:

• Boundary conditions corresponding to the stationary state, a zero Neumann BC


at the inflow boundary and a Dirichlet one (a constant imposed water level) at the
outflow.
• Simple non-reflecting BC at inflow and outflow.

The non-reflecting boundary conditions (section 2.7.8) are based on the properties of
long waves imposed on a channel flow with a constant flux. They are realised in the
form of Dirichlet BC for water level at the open boundaries Sb , whereby free outflow BC
for the velocity and the hydrostatic pressure (i.e. imposed hydrodynamic pressure equal
to 0) are assumed. For a channel aligned along the x-axis:
heq (un+1 − ueq )
Sb = Seq + m √ n (5.48)
gh
where un+1 is the vertical mean velocity at the time level n + 1 and hn the local water
depth at the previous time step. The variables denoted by subscript eq concern the
expected values at the open boundaries when the stable state is achieved. m = +1 at
the outflow and m = −1 at the inflow boundary [141].
Imposing Dirichlet BC for water level at open boundaries, especially those that are only
approximately valid, is connected with numerical instabilities assigned to reflections
of waves at these points and other incompatibilities. In practice some stabilisation or
filtering of the free surface is required. In this case, stabilisation by the method described
in section 3.6.5 is applied. Without it, oscillations occur at those points, where Dirichlet
boundary conditions are applied.
The computation was performed with the overall horizontal mesh resolution of 8 m,
21 mesh levels (10 m water depth before and behind the bump) and the time step
equal 1 s. The stabilisation of the free surface profile over time for reflecting (stationary
state) and simple non-reflecting boundary conditions during 1000 s and during first 100 s
(detailed) is shown in figures 5.34 and 5.35. Upon application of the non-reflecting BC
the stable state is achieved quickly, for the reflecting BC the free surface does not stabilise
in a comparable time.
The steady state free surface profile achieved by application of the non-reflecting BC is
presented in figure 5.36, the minimum value of the free surface level is 0.05451 m, with
an error of 1.6% compared to the analytical solution.
As a conclusion, for computations starting from simple homogeneous initial conditions
and aimed at obtaining a stationary free surface flow, boundary conditions are crucial
for the convergence efficiency to the stable state. Non-reflecting boundary conditions
are required. Without them, the initially occurring waves cannot leave the domain.
With boundary conditions corresponding to the stationary state, the convergence can
be enhanced only by artificially increasing the dissipative properties of the free surface
computation. However, it also shows that for practical purposes filtering of the free
surface is required in order to obtain stable and converging solutions.
172 5.7 Channel with a bump

1000 1000

0.00
900 900
0.00
0.00

800 800
0.00
-0.05

700 700

-0.05
0.00
0.0
0

600 600
0.
00
time [s]

time [s]

500 0.00 500


05
-0.
-0.05

400 0.00 400


0.00
0.05
0.0 0.00
300 5 300

0.00
-0.0
200 5 200
-0.05 -0.05
0.00
100 0.00 100 0.00
0.05
0.00 0.10 0.00
0.10 0.10
0 0.00 0 0.00 -0.05
0 100 200 300 400 500 600 0 100 200 300 400 500 600
length [m] length [m]

Figure 5.34: A channel with a bump. Stabilisation of the free surface profile for reflecting
(left) and simple non-reflecting boundary conditions (right) during 1000 s of computation
time.
5.8 Waves travelling over an underwater channel 173

100 100
0.00
0.10
90 90

0.05

-0.05
-0.05
0.0

0.
80 80
5

00
00
0.

0.0
0
0.05

70 0. 0.15
10 70
0.10

60 60

0.0
time [s]

time [s]

0.00

-0.05

0
50 50

-0.05
0.0
0.0

5
5

0.00
0.00
-0.05

40 40
0.10

0.
10
0.0

30 30
5

0.0 0.1 .15 5 0.10 0.


-0.
0.00

0
05

-0.1

5 0

.1 5
0 .1
20 20

-0 -0.0
0
0

0.00

-0.05

0
5 .1
10 10
0.1

.0 -0
0.1.05

5
0

05

0.

-0
0.
0

00 00
0 0
0 100 200 300 400 500 600 0 100 200 300 400 500 600
length [m] length [m]

Figure 5.35: A channel with a bump. Stabilisation of the free surface profile for reflecting
(left) and simple non-reflecting boundary conditions (right) for 100 s computation time
(details).

5.8 Waves travelling over an underwater channel


A simple, but demanding test for free surface models is to simulate the propagation
of waves over varying bathymetry. Additionally, the travelling waves should enter and
leave the computational domain without artificial effects. In this case a typical situation
which can be found in coastal regions is studied, with similar parameters and geometry to
those used by Schröter [142]. A wave-train with a small wave amplitude of η0 = 0.125 m
and a period of T = 4 s enters a domain represented by a long channel from the left
side (x = 0) in shallow water (equilibrium depth of 3 m) and then approaches a deep
underwater channel. The waves leave the domain after passing a final shallow water part
3 m deep.
Two geometries are chosen, shown in figure 5.38. They differ in length (600 m and
800 m), as well as in distance between the inflow and outflow domains and the channel
edges (50 m and 50 m or 100 m and 200 m, respectively), but not in the underwater
174 5.8 Waves travelling over an underwater channel

0.0

-0.01

-0.02
water level [m]

-0.03

-0.04

-0.05

-0.06
0 100 200 300 400 500 600
length [m]

Figure 5.36: A channel with a bump. Steady state free surface profile.

channel shape. The greatest depth of the 500 m long cosinusoidal channel cross-section is
43 m. The computational mesh structure is very similar in structure to the one shown in
figure A.2, but 6 elements wide and 600 elements long, with a resolution in the direction
parallel to the channel axis of 1 m or 1.33 m. In both cases the three-dimensional mesh
has 21 levels distributed more densely nearby the surface. The time-step is 0.1 s and the
simulation time 300 s. Viscous flow is assumed, with viscosity of ν = 10−4 m2 /s, but
with no shear at all impermeable channel boundaries.
Due to the small wave height non-linear effects are not to be awaited. For a wave of
this period, at the entrance (water depth of 3 m) wave length of 19 m is to be observed.
In the deep part of the channel, the deep water wavelength approaches 25 m (from the
dispersion relation A.34). In the domain, the ratio between the equilibrium water depth
and the wave length D/λ varies between 0.16 and 1.7.
At the open outflow boundary (x = 600 m or x = 800 m), a simple non-reflecting
boundary condition for the free surface elevation Sout with the incoming wave celerity
and the mean velocity in the form:

hn un+1
Sout = Seq + √ n (5.49)
gh

is imposed, with free outflow conditions for the velocity and zero Neumann boundary
condition for the hydrodynamic pressure. Seq is the equilibrium free surface elevation
(in this case it equals zero). At the inflow boundary (x = 0 m), the same boundary
conditions for the velocity and the pressure are given, but the non-reflecting boundary
condition takes the more complicated form imposing an incoming wave profile while
an underwater channel.
Figure 5.37: Water surface time development for the test case of waves travelling over

5.8 Waves travelling over an underwater channel


0.1 t=50s

z [m]
0.0

-0.1

0 100 200 300 400 500 600 700 800


x [m]
0.1 t=100s
z [m]

0.0

-0.1

0 100 200 300 400 500 600 700 800


x [m]
0.1 t=150s
z [m]

0.0

-0.1

0 100 200 300 400 500 600 700 800


x [m]
0.1 t=200s
z [m]

0.0

-0.1

0 100 200 300 400 500 600 700 800


x [m]
0.1 t=300s
z [m]

0.0

-0.1

175
0 100 200 300 400 500 600 700 800
176 5.8 Waves travelling over an underwater channel

x [m]
0 100 200 300 400 500 600
0.2 0.2

0.0 0.0
[m]

[m]
-0.2 =19m =25m -0.2
=19m

0 0
-3m -3m
z [m]

z [m]
-20 -20

-43m
-40 -40

0 100 200 300 400 500 600


x [m]

x [m]
0 100 200 300 400 500 600 700 800

0.1 0.1
[m]

[m]
0.0 0.0

=19m
-0.1 =25m -0.1
=19m

0 0
-3m -3m
z [m]

z [m]

-20 -20

-43m
-40 -40

0 100 200 300 400 500 600 700 800


x [m]

Figure 5.38: Bottom and free surface profiles (t = 300 s) for the test case of waves
travelling over an underwater channel.
5.8 Waves travelling over an underwater channel 177

simultaneously allowing waves arriving from the domain interior to leave it [142]:

hn un+1 − uimp (H + Simp )


Sin = Simp − (5.50)
ce

The imposed wave profile is:


π
Simp = η0 cos( + kx − ωt)
2
1 η0 gk (exp[k(H + Simp )])2 − 1 π
uimp (H + Simp ) = cos( + kx − ωt)
2 ω exp[k(H + Simp )] cosh(kH) 2
2
ω = gk tanh(kH) (5.51)
ce = 2π/kT

where H = 3 m is the equilibrium water depth and uimp is the vertically integrated
imposed velocity.
The development of the wave profile as it enters the domain is shown in figure 5.37 for
the longer domain. The frequency dispersion of the wave train front is the characteristic
feature of the free surface profiles after 50 s and 100 s, with longer waves travelling
quicker than the shorter ones. A reflection from the shoaling underwater channel slope
is to be observed, as well as viscous damping of the amplitude.
The observed wavelengths are in excellent agreement with the theory (appendix section
A.3), yielding approximately λ = 19 m in the shallow parts before and after passing the
deeper part, and λ = 25 m over the underwater channel (figure 5.38).
In the case of the shorter channel, the reproduction of the wave length is also correct,
but as the main wave train arrives at the steep slope between x = 300 m and x =
500 m, a strong reflection occurs causing a considerable increase of the wave amplitude.
The boundary conditions, as stated above, allow the waves to enter and to exit the
domain smoothly. However, their imperfect form (valid entirely for long waves only)
causes a reflection. When the sections truncating the domain are located too near
to the area of the main interest, the influence of the imperfect boundary conditions
may interfere strongly with the solution in the domain interior. In the case of the
shorter channel the reflection from the boundary could be misinterpreted as an influence
of the shoaling bottom. Additionally, the boundary conditions are not perfectly mass
conserving between the inflow and outflow sections. In the case of the shorter channel,
during 300 s about 0.25% of the water mass leaves the domain added to numerical mass
losses of 0.5%. In the case of the longer channel, mass losses are about 0.1%, and about
0.05% of the initial mass leaves the domain.
As a conclusion, the model is able to deal with various wave problems over a changeable
bathymetry reproducing typical phenomena properly. However, the greatest care is rec-
ommended by locating inflow and outflow sections and setting boundary conditions for
waves entering and leaving the domain so that they are mass-conserving, non-reflective,
numerically stable and do not negatively interfere with the solution in the domain inte-
rior.
178 5.9 Solitary wave in a long channel

5.9 Solitary wave in a long channel

5.9.1 Theory
A solitary wave is a single elevation of water surface above an undisturbed surrounding,
which is neither preceded nor followed by any free surface disturbances. Neglecting
dissipation, as well as bottom and lateral boundary shear, a solitary wave travels over a
horizontal bottom without changing its shape and velocity. Therefore, it is a good test
case for testing capability of free surface algorithms to describe propagation of non-linear
waves of finite amplitude. The accuracy of the model can be evaluated by comparing the
amplitude and celerity of the wave with its theoretical values, as well as the deformation
of the wave as it travels.
There are numerous analytical studies of this form of non-linear finite-amplitude wave.
The first approximation provided by Laitone [95, 132] is the most frequently used for
comparative studies. For a vertical section of an infinitely long channel of an undisturbed
depth h (z = 0 at the undisturbed surface), the following approximate formulae for
velocity components u, w, free surface elevation η, pressure p and wave celerity c of a
solitary wave with a height of H are valid:
s 
q H 3H
u= gd sech2  (x − ct) (5.52)
h 4 h3
s  s 
3/2  
H z 3H 3H
q 
w= 3gh sech2  3
(x − ct) tanh  (x − ct) (5.53)
h h 4h 4 h3
s 
3H
η = h + Hsech2  (x − ct) (5.54)
4 h3
p = ̺g(η − z) (5.55)
q
c= g(H + h) (5.56)

It is interesting that in this analytical approximation the vertical velocity component is


not treated as small, as commonly taken, but the pressure can be assumed hydrostatic
(equation 5.55) at the same level of exactness as the horizontal velocity (with O((H/h)2 )
[95]. Therefore, this initial condition is suitable for fair comparisons between models
with and without hydrostatic approximation and the initial value of zero hydrodynamic
pressure is assumed. Although the initial velocity field (5.52–5.53) is perfectly divergence-
free, larger values of the hydrodynamic pressure appear immediately after the first time
step (60% of the value of the hydrostatic pressure at the bottom).
Following the test cases provided by Ramaswamy [132], a solitary wave described by
the formulae (5.52–5.56) is applied in a long channel as an initial condition, and the
behaviour of the solution is observed thereafter. Due to the fact that the simulation is
performed in a finite domain, and Laitone’s formulae are valid for an indefinitely long
channel, care must be taken choosing the initial position of the wave crest. In order to
5.9 Solitary wave in a long channel 179

deal with it, the use of the effective wave length λ concept is made. λ is equal to the
doubled length between the wave crest and a point, where the free surface elevation is
η(x) = 0.01H. According to Laitone:
s
h3
λ = 6.9 (5.57)
H

For example, when H/h = 0.2, λ/2 ≈ 8D, and for a channel of 10 m depth, the initial
distance between the solitary wave crest and a boundary should be at least 80 m. The
steepness of the solitary wave grows with H/h. As the solitary wave travels along a
shoaling bottom, its height H increases, until wave breaking appears. As the criterion
for wave breaking the critical condition is assumed when the particle velocity at the crest
equals the wave celerity. Various approximations yield values for maximum solitary wave
height between H = 0.727h and H = 0.827h [113, 95].

5.9.2 Solitary wave propagation over a flat bottom


In order to test the free surface algorithm, the solitary wave propagation in a long
channel is studied. In an ideal case, the wave should travel without changing its form
and amplitude.
A long channel 600 m long and 6 m wide, with a constant depth of 10 m is taken.
The mesh is similar to the one shown in figure A.2, but with 6 elements wide and
600 long, with a resolution in the direction parallel to the channel axis of 1 m. The
three-dimensional mesh has 11 equidistantly distributed levels. Inviscid flow without
shear on walls and the bottom is assumed. All boundaries are impermeable. As the
initial condition the hydrostatic approximation given by formulae (5.52–5.56) is applied,
with a varying wave height of H = 2 m, 5 m, 7 m and 10 m (H = 0.2h, H = 0.5h,
H = 0.7h, H = h), and the initial crest position at x = 80 m. The time step is taken
as constant, ∆t = 0.1 s, and the simulation time 40 s (Courant number in the direction
of wave propagation varies from 0.2 to about 1.0 at the wave crest). For computation
of the free surface elevation in the non-hydrostatic case the SUPG method based on the
kinematic boundary condition or the conservative free surface equation is applied. For
hydrodynamic pressure, the zero value at the free surface is imposed.
The results for comparison of free surface computation methods are summarised in figure
5.39, where semi-implicit (implicitness factor θ = 0.55) and implicit (θ = 1.0) formu-
lations for both methods are shown for wave height H = 2.0 m. The development of
free surface profiles in time is presented. For all methods the results are very good with
amplitude error less than 5%. The influence of the implicitness factor is much larger
in case of the kinematic boundary condition, which can be better seen with a larger
wave height H = 5 m, figure 5.40. While the results with the conservative free surface
equation remains relatively independent of the implicitness, the results with kinematic
BC show large (up to 26%) dependency on it.
With larger amplitude, the shape of the wave broadens while travelling. The mass losses
also grow with the initial wave height. For kinematic BC, the mass loss for H = 2 m is
180 5.9 Solitary wave in a long channel

600

600

600

600
H = 2.05 m

H = 1.89 m

H = 1.99 m

H = 1.95 m
550

550

550

550
500

500

500

500
450

450

450

450
400

400

400

400
350

350

350

350
x [m]
300

300

300

300
250

250

250

250
200

200

200

200
150

150

150

150
100

100

100

100
H = 2.0 m

H = 2.0 m

H = 2.0 m

H = 2.0 m
50

50

50

50
Csi
Ksi

Ci
Ki
0

0
2

z [m] z [m] z [m] z [m]

Figure 5.39: Solitary wave propagation in a long channel with a flat bottom for free
surface computation methods based on SUPG: Ksi, Ki: kinematic boundary condition,
semi-implicit and implicit; Csi, Ci: conservative free surface equation, semi-implicit and
implicit. The free surface profiles shown for t = 0 s, 10 s, 20 s, 30 s, and 40 s.
5.9 Solitary wave in a long channel 181

600

600

600

600
H = 5.91 m

H = 3.73 m

H = 5.05 m

H = 4.69 m
550

550

550

550
500

500

500

500
450

450

450

450
400

400

400

400
350

350

350

350
x [m]
300

300

300

300
250

250

250

250
200

200

200

200
150

150

150

150
100

100

100

100
H = 5.0 m

H = 5.0 m

H = 5.0 m

H = 5.0 m
50

50

50

50
Csi
Ksi

Ci
Ki
0

0
6

z [m] z [m] z [m] z [m]

Figure 5.40: Solitary wave propagation in a long channel with a flat bottom for free
surface computation methods based on SUPG: Ksi, Ki: kinematic boundary condition,
semi-implicit and implicit; Csi, Ci: conservative free surface equation, semi-implicit and
implicit. The free surface profiles shown for t = 0 s, 10 s, 20 s, 30 s, and 40 s.
182 5.9 Solitary wave in a long channel

600

600

600

600
H = 1.89 m
H = 4.3 m

H = 4.2 m

H = 3.7 m
550

550

550

550
500

500

500

500
450

450

450

450
400

400

400

400
350

350

350

350
x [m]
300

300

300

300
250

250

250

250
200

200

200

200
150

150

150

150
100

100

100

100
H = 10.0 m

H = 7.0 m

H = 5.0 m

H = 2.0 m
50

50

50

50
0

0
10

z [m] z [m] z [m] z [m]

Figure 5.41: Solitary wave propagation in a long channel with a flat bottom. Computed
from implicit kinematic boundary condition with SUPG. Initial wave heights of H = 2 m,
5 m, 7 m and 10 m (H = 0.2h, H = 0.5h, H = 0.7h, H = h, h=10 m), note the changing
vertical axis range. The free surface profiles shown for t = 0 s, 10 s, 20 s, 30 s, and 40 s.
5.9 Solitary wave in a long channel 183

2.0

1.5
z [m]

1.0

0.5

0.0

0 50 100 150 200 250 300 350 400 450 500 550 600
x [m]

2.0

1.5
z [m]

1.0

0.5

0.0

0 50 100 150 200 250 300 350 400 450 500 550 600
x [m]

Figure 5.42: Solitary wave propagation in a long channel with a flat bottom. Initial
relative wave heights H = 0.2h. Above, the solution obtained with hydrostatic approx-
imation, below, without it. The free surface profiles shown for t = 0 s, 10 s, 20 s, 30 s,
and 40 s.
184 5.9 Solitary wave in a long channel

kin BC s-impl
2.0
kin BC impl
cons FSEq s-impl
cons FSEq impl
kin BC charact
1.5
z [m]

1.0

0.5

0.0

400 420 440 460 480 500 520 540 560 580 600
x [m]

Figure 5.43: Solitary wave profile for t = 40 s for various methods of free surface com-
putation.

less than 0.1%, and for H = 5 m increases to 0.2%. For conservative formulation, it is
very small for H = 2 m (about 2×10−3 %) and increases to about 0.7% for H = 5 m.
The dispersive effects behind the wave also diminish with the initial wave height. It
is necessary for the two-dimensional velocity field at the surface to be filtered once as
described in section 3.6.5 for the kinematic BC. Without filtering, the wave frequency
dispersion appearing far behind the solitary wave crest leads to very short waves with
wavelengths comparable with the mesh resolution (at t = 20 s), which causes growing
numerical instabilities. (When instead of the velocity at the free surface, the free surface
itself is filtered, then, for example, the wave height H diminishes from 2 m to 1.72 m for
the implicit case.)
For larger amplitudes, in the range of wave-breaking, the numerical oscillations grow
quickly, leading to program break-up in the semi-implicit case. For the implicit algo-
rithm, the results are not acceptable for the conservative equation, while the results
with kinematic BC for all investigated wave heights are shown in figure 5.41. For the
extreme initial wave height equal to the water depth (H = 10 m), i.e. already above the
highest possible wave according to Laitone’s formula (and unphysical), the worst results
are obtained. At t = 40 s, the amplitude diminishes to 4.3 m, 57% loss, with volume loss
during computation of 0.5%. A very strong generation of secondary waves behind the
travelling soliton and dispersive losses are observed. The results improve considerably
with diminishing initial amplitude of the wave. For H = 7 m, H = 5 m, H = 2 m, the
5.9 Solitary wave in a long channel 185

amplitude decreases to 4.2 m, 3.7 m and 1.89 m, respectively, which are equivalent to
40%, 26%, and 5.5% amplitude loss, with volume (mass) error 0.2%, 0.1%, and 0.01%.
The method of characteristics, applied for kinematic BC, yield the worst results for the
same time-step (Courant numbers 0.2-1.0). While for initial wave height H = 2 m the
mass loss is about 0.6% and the final height is H = 2.06 m, for initial wave height of 5 m
a considerable growth to 6.99 m of final amplitude with a mass loss of 3% is obtained.
The algorithm becomes quickly unstable for larger wave heights.
For the case with H = 2 m, a comparative study with a solution obtained under hydro-
static approximation is performed, figure 5.42. In both approximations the semi-implicit
formulation is applied. The hydrostatic (i.e. shallow water) solution shows a strong
amplitude dispersion, resulting in a characteristic distorted wave form, where the wave
crest moves with larger celerity than the wave slopes. The amplitude diminishes to
H = 1.417 m (29% loss), compared to the non-hydrostatic final amplitude H = 1.89 m
(5.5% amplitude loss) when kinematic BC is applied and H = 1.989 m (0.5% loss)
when use is made of the conservative free surface equation (shown in figure 5.42). The
non-hydrostatic wave crest travels with a mean celerity of 10.67 m/s or 10.77 m/s, re-
spectively (figure 5.43), which is in good agreement with a theoretical speed of 10.85 m/s
for the initial condition values and (5.56). The crest of the hydrostatic wave travels with
a mean speed of 13.6 m/s.
As a conclusion, in the newly developed algorithms the general configuration of the
solitary wave remains unchanged, no amplitude dispersion can be observed. Much more
care must be taken while calibrating models using the kinematic boundary condition
than by the ones applying the conservative free surface equation. The worst results are
obtained with the method of characteristics for kinematic BC. The frequency dispersion
behind the wave appears due to the fact that the initial condition is not the exact
solution of the solved equations (it is obtained with the hydrostatic approximation) and
because of the interaction with the wall behind the wave. This dispersion effect increases
considerably with initial height of the solitary wave. Not surprisingly, when this height
is comparable with the water depth (unphysical case), the results are not acceptable
anymore. The accuracy of the model grows for waves with smoother slopes and lower
amplitudes.
The results are thoroughly comparable, when instead of impermeable walls open bound-
aries with imposed hydrostatic pressure and simple non-reflecting boundary conditions
for the free surface elevation at the boundary Sb are given at the channel ends:
q
Sb = Seq ± (un+1 (h + η))/ g(h + η) (5.58)

5.9.3 Solitary wave in a long channel with varying depth


In order to study the ability of the model to simulate finite amplitude wave interaction
with varying bathymetry, a numerical experiment is performed in a very similar setting
to before. In the long channel from the previous case the bottom profile is changed. The
186 5.9 Solitary wave in a long channel

H=2m 0s

10s

20s

30s

40s

50s

0 100 200 300 400 500 600


x [m]

H=2m 0s

10s

20s

30s

40s

50s

0 100 200 300 400 500 600


x [m]

Figure 5.44: Profiles of solitary wave propagating in a long channel with variable depth.
Initial relative wave height H = 0.2h. The non-hydrostatic solution (above) and the
hydrostatic solution (below).
5.9 Solitary wave in a long channel 187

600

600

600
u [m/s]

u [m/s]

u [m/s]
2

2
H=1.2m
0.3 0.1

550

550

550
500

500

500
450

450

450
H=1.6m
400

400

400
0.3
0.7
350

350

350
x [m]

x [m]

x [m]
300

300

300
250

250

250
200

200

200
150

150

150
H=2m

100

100

100
0.3
1
0.

-0.5 -0.3 1
-0.
50

50

50
t=30s

t=50s
t=0s

0
-10

-10

-10
0

0
-5

-5

-5

z [m] z [m] z [m]

Figure 5.45: Solitary wave propagation in a long channel with a bottom of variable
depth. The free surface profiles shown for t = 0 s, 30 s, and 50 s. Iso-lines: vertical
velocity component. Solution obtained with hydrostatic approximation.
188 5.9 Solitary wave in a long channel

600

600

600
u [m/s]

u [m/s]

u [m/s]
2 0.3

2
H=3.3m
0.1

550

550

550
-0.1 0.7
500

500

500
H=0.8m
-0.1
450

450

450
H=2.6m
400

400

400
0.1
350

350

350
-0.1
x [m]

x [m]

x [m]
300

300

300
250

250

250
200

200

200
150

150

150
H=2m

100

100

100

0.3
1
0.

-0.5 -0.3 1
-0.
50

50

50
t=30s

t=50s
t=0s

0
0

0
-5

-5

-5
-10

-10

-10

z [m] z [m] z [m]

Figure 5.46: Solitary wave propagation in a long channel with a bottom of variable
depth. The free surface profiles shown for t = 0 s, 30 s, and 50 s. Iso-lines: vertical
velocity component. Solution obtained without hydrostatic approximation.
5.9 Solitary wave in a long channel 189

kin BC s-impl
4 kin BC impl
cons FSEq s-impl
cons FSEq impl
kin BC charact
3
z [m]

440 460 480 500 520 540 560 580 600


x [m]

Figure 5.47: Solitary waves profiles for t = 50 s for various methods of free surface
computation.

equilibrium water depth decreases now at the length between x = 160 m and x = 260 m
from 10 m to 5 m. The initial solitary wave height is H = 2 m. The obtained free
surface profiles are shown in figure 5.44. For the non-hydrostatic computation (using the
semi-implicit conservative free surface equation), when the solitary wave passes through
the area, where the water depth diminishes, the front of the wave steepens and then
the wave divides into two waves. Further on, by t = 50 s, the wave separation becomes
more evident, due to the fact that both waves move with different celerities, depending
on their height. When the hydrostatic approximation is applied, no such effects are
observed. The solutions for t = 0 s, t = 30 s and t = 50 s are additionally shown
in figures 5.46 and 5.45. Comparison of the results when applying various free surface
algorithms is shown in figure 5.47.
The laboratory works of Madsen and Mei [107] yield for similar (but scaled) geometry
the relative height of 160% for the larger wave and 71% for the smaller one, compared
to the initial wave height; Hauguel [67] provides values 163% and 56%. Similar results
(153% and 55%) were achieved in a numerical model by Schaper [139]. The results
of the model developed here are best to be seen in figure 5.47. Various methods of
computation yielded values in the range of 40-45% for the smaller wave and 110-210%
for the larger one. In consequence, for modelling the influence of bottom topography
on waves, the non-hydrostatic model yields very convincing results. However, when
applying the kinematic BC for obtaining the free surface, greater care must be taken
when calibrating the algorithm (figure 5.47).
190 5.9 Solitary wave in a long channel

0s
H=2m H=2m

10s

H=4.26m
19s

30s

H=1.96m H=1.96m
40s

0 100 200 300 400 500 600


x [m]

0s
H=2m H=2m

10s

H=3.5m
17s

30s

H=1.42m H=1.42m
40s

0 100 200 300 400 500 600


x [m]

Figure 5.48: Collision between two solitons of equal amplitude; time series of free surface
profiles obtained without (above) and without (below) hydrostatic approximation.
5.9 Solitary wave in a long channel 191

0s
H=2m H=4m

10s

H=6.21m 19s

30s

H=3.91m
40s H=1.94m

0 100 200 300 400 500 600


x [m]

Figure 5.49: Collision between two solitons of different amplitude; time series of free
surface profiles. Initial wave heights H = 2 m, H = 4 m (H = 0.2h, H = 0.4h,
h = 10 m).

5.9.4 Collision between two solitons

This test case is designed to test the capability of the model to describe wave-wave
interactions. A head-on collision of two solitary waves in a channel with a flat bottom
and water depth of h = 10 m is studied. During the collision, the vertical accelerations
of the water particles are large and approximate analytical theories are no longer valid
[102]; therefore this is a good example to test the properties of the non-hydrostatic model.
For incident waves of equal amplitude, the highly reflective properties are observed, and
after collision the wave characteristics should not change.
The results are shown in form of a time series of water surface profiles. Only the results
for the semi-implicit conservative free surface equation are shown; for other methods
no new conclusions compared to these from previous sections can be formulated. The
results are very good for the height of the waves H = 2 m (figure 5.48). At t ≈ 19 s,
the maximum amplitude is 4.26 m, at t = 40 s, the amplitude of both waves equals
1.96 m, almost the same as in the case when the wave travels without any interaction,
as in section 5.9.2. The wave profiles remain conserved. The results obtained with
hydrostatic approximation are also shown in figure 5.48.
In the case of solitons of different initial heights, the original waves after colliding are
deformed due to dispersive effects and generation of secondary waves, especially behind
the larger wave. The amplitude and celerity of the initial waves change after collision
192 5.10 Limits of application

(in the real world cases, additional dissipation occurs). A collision between two solitons
with initial wave heights H = 2 m and H = 4 m is shown in figure 5.49.
As a conclusion, the newly developed model reproduces properly the wave-wave inter-
action of non-linear waves of a finite amplitude. The accuracy of the model diminishes
with the incident wave steepness.

5.10 Limits of application

5.10.1 Free surface breaking


The algorithm developed in this work requires that the free surface and bottom must be
represented by single-valued functions. A numerical experiment is performed in order
to test the reaction of the developed algorithm situations when the physical conditions
impose infinite free surface gradients, causing the free surface breaking. For example,
such conditions are met in these flow regions, where transition between supercritical and
subcritical free surface flow occurs.
The experiment is performed in a similar setting to the supercritical inviscid channel flow,
section 5.6. A ramp is built into a 15 m long channel. The bottom level is z = −2 m
from the beginning of the channel to x = 7 m, then at a length of 1 m (between x = 7 m
and x = 8 m) the bottom ascends up to a level z = 0 m, forming a very steep slope of
2:1 (200%), resolved by 10 nodes with a spacing of ∆x = 0.1 m. The initial free surface
is flat and its level is at z = 1 m. The initial velocity is u = 6 m/s everywhere. 11 mesh
levels with equal spacing is applied. The initial flow is supercritical, but an increase of
depth over the value of 3.66 m immediately yields subcritical conditions (F r > 1).
The change of the vertical profile of the initially occurring wave over time is shown in
figure 5.50. When the subcritical conditions are achieved locally just before the ramp
(x < 7 m), the occurring free surface distortion can travel upstream and conditions for
wave breaking occur while the wave amplitude grows. (In the real world, a hydraulic
jump eventually forms.) The algorithms based on the height function free surface track-
ing method (in this case SUPG, implicit) implemented in the model cannot describe
such steep free surface slopes because they require the free surface to be a single-valued
function. The computation breaks up a few time steps after the time the last profile is
plotted, i.e. after t = 2.5 s. The break-up condition is achieved, when the free surface
oscillations reach the bottom.
A similar situation is observed, when the initial velocity is increased to u = 10 m/s
everywhere. Then supercritical conditions are ensured in the whole area, even when the
flow depth grows. The surface breaks by t = 2.0 s due to infinite free surface gradients,
figure 5.51. Similar break-up of the free surface is obtained with 1:1 step (ascend from
z = −1.0 m to 0 m by initial free surface at 1.0 m). Stable results are obtained (at
this resolution) for a slope reduced to 1:2 (ascend from z = −0.5 m to 0 m), with a free
surface slope of ca. 60%. Changing the number of the mesh levels does not improve the
results.
5.10 Limits of application 193

5
z [m]

1
0 3 6 9 12 15
x [m]

Figure 5.50: Free surface profiles for a flow over a steep bottom step by conditions
allowing appearance of a hydraulic jump. Profiles shown each 0.5 s. The free surface
breaks by ca. 2.2 s.

As a conclusion, model flow settings causing extremely steep free surface gradients must
be avoided. As the heuristic tests show, the problems with obtaining stable solutions
begin when the slopes reach and exceed π/4 (i.e. 1:1).

5.10.2 Bottom gradients for supercritical flow


In this section the stability of the numerical solution depending on the larger bottom
(and free surface) gradients is investigated. When the triangles of the basis mesh are
not far distorted from the ideal isosceles triangle for which the interpolation is most
exact, the only source of the prismatic elements deformation are the free surface and the
bottom. The computation of free surface gradients, which drive the flow, is erroneous in
very distorted elements.
In an analogous setting to the previous section, but with much larger initial water depth
of 10 m and initial velocity of 20 m/s (supercritical conditions ensured everywhere), it
is investigated at what maximum slope a steady flow can be achieved (step resolution as
in previous chapter). Supercritical conditions are chosen, because in this case the free
surface is much steeper than in the subcritical flow, which makes the test much more
severe. It was found that a stable flow after an initial oscillation over the step can still
be obtained for bottom gradients of 1:1 (it is, an ascent of the bottom from −1 m to
0 m level at a length of 1 m and resolved by 10 nodes in the flow direction). The free
surface profile is shown in figure 5.52, for t = 60 s. By bottom gradient of 2:1, free
surface gradients obtained during oscillations in the transient computation are too large,
causing the program to break up. The hydrodynamic pressure distribution is similar
194 5.10 Limits of application

5
z [m]

0
0 3 6 9 12 15
x [m]

Figure 5.51: Free surface profiles for a flow over a steep bottom step by supercritical
conditions. Profiles shown each 0.5 s. The free surface breaks by ca. 2.0 s.

11.4

11.2

11.0

10.8
z [m]

10.6

10.4

10.2

10.0

0 3 6 9 12 15
x [m]

Figure 5.52: Free surface profile for supercritical flow (velocity u = 20 m/s) over a step
of the steepnes 1:1.
5.10 Limits of application 195

20

15

10
z [m]

0 50 100 150 200


x [m]

Figure 5.53: Development of the free surface profile (shown each 1 s) for a solitary wave
travelling from the left to the right with an initial relative height of H/h = 0.7.

to the one shown in figure 5.29, but with much larger maximum values in the range of
10 kPa, i.e. comparable with the hydrostatic pressure values.
As the heuristic tests show, the problems with obtaining stable solutions begin when the
slopes exceed π/4. Additionally, the computation of the free surface gradients using the
weak formulation in finite elements (3.31) can be erroneous in very distorted elements.

5.10.3 Breaking of a solitary wave


The solitary wave breaks, when its height exceeds 70-80% of the water depth (5.9.1). In
a numerical model based on a height function free surface representation and which does
not provide any artificial damping of the free surface, this condition will cause numerical
oscillations, model break-up, mass losses, etc. For example, when as an initial condition
in the test case for solitary wave propagating over a flat bottom (5.9.2) a wave height
of H = 7 m by water depth of h = 10 m is given, wave-breaking occurs. The effects
for the semi-implicit conservative free surface equation are shown in figure 5.53, large
oscillations at the wave front cause the program to break up by t = 5.1 s. The model
reacts properly for a physically infeasible initial condition.
In the next test the limits of simulation of free surface flows are tested in a more natural
setting, independent of initial conditions. A solitary wave whose initial profile is given
by equations (5.52–5.56) propagates 200 m along a flat bottom in a depth of 10 m, and
then approaches a shore along a shoaling bottom represented by a slightly sloped plane,
with depth decreasing along a path of 400 m between 10 m to 0.1 m only. All other
parameters are similar to those in section 5.9.2. The initial wave height is 5 m, also below
196 5.10 Limits of application

10
-h
-0.8h
8

6
0s 30s
10s
20s
z [m]

4 40s

50s
2

0 100 200 300 400 500 600


x [m]

10
-h
-0.8h 30s
8
40s 48s

6 20s
0s 10s
z [m]

0 100 200 300 400 500 600


x [m]

Figure 5.54: Solitary wave (initial amplitude 5 m) approaching along a shoaling bottom.
The shoaling bottom profile is provided for comparison with the wave height. The
kinematic BC (above) or the conservative free surface equation (below) used for free
surface.
5.10 Limits of application 197

the breaking condition in the water depth of 10 m. As the wave approaches the shore,
at some point the conditions of wave breaking, when the wave height reaches about 80%
of the equilibrium depth, are fulfilled (section 5.9.1). The model is unable to simulate
the wave breaking due to the fact that the free surface is represented by a single-valued
height function, but the model behaviour in this extreme case is to be tested.
The results of this test are shown in the form of free surface profile time series in figure
5.54, for implicit kinematic BC and the implicit conservative free surface equation. As
anticipated, when the wave height exceeds about 80% of the equilibrium water depth
(wave breaking condition), the wave steepness is very large and the velocity at the
wave crest is larger than the wave celerity. However, in the case of the kinematic BC,
the model continues to compute further, but the wave peak of a height exceeding the
breaking condition is simply cut out. By then, when large differences between the slopes
of neighbour elements occur, considerable mass loss in the computational domain is
observed. For the simulation time up to 30 s the mass loss is small, below 0.1%, further
on (after 60 s) it reaches a large value of 10%. In the case of the conservative formulation,
the program breaks up – the oscillations of the free surface reach the bottom. The model
reacts as expected in a situation, when its limits of application are reached during the
computation.
198 5.10 Limits of application
Chapter 6

Summary and Conclusions

Jede Lösung eines Problems


ist ein neues Problem.
Johann Wolfgang von Goethe

6.1 Realised new developments


The newly developed algorithm, implemented in the Telemac system structure, ad-
dresses two important aspects. First of all, the vertical acceleration in free surface
incompressible flows is taken into consideration by solving the vertical momentum con-
servation equation. Second, the free surface computation method allows simulation of
its movements without limitation typical for shallow water equations – restriction to
long waves as well as gentle free surface and bottom slopes. A few limitations remain
due to the mesh structure (σ–mesh) and the linear interpolation in the finite elements,
which e.g. do not allow breaking waves and introduce spurious errors by approximating
stronger non-linearly varying functions.
The new algorithm, whose theory is presented in chapter 2 and numeric implementation
in chapter 3, is based consistently on the operator splitting scheme. This means that
the operators in the hydrodynamic equations are split into parts and the resulting equa-
tions are treated in separate steps of the algorithm using the most appropriate methods
for their mathematical character. Hydrodynamic variables are computed in subsequent
stages (fractional step method). The overall algorithm stability and exactness is pro-
vided when all stages fulfil these requirements separately. The decoupled structure of
the algorithm, which does not use the continuity equation explicitly, allows application
of the equal-order interpolation functions in finite elements for both pressure and veloc-
ity components. Applying the equal order interpolation functions is not possible with
coupled methods for incompressible flows due to the LBB condition. Additionally, a
much simpler structure of the algorithm than for finite element methods with mixed
interpolation functions results. The assumed reference element type, a prism with six
nodes and linear approximation functions, is a good compromise between the exactness

199
200 6.1 Realised new developments

of the solution, model complexity and computational cost. Compared to the widely used
mixed interpolation, with linear interpolation functions for velocity and constant ones
for pressure, much better resolution of the (hydrodynamic) pressure is obtained, which
is very important in a domain of variable geometry.
The main idea of the new non-hydrostatic model is to decompose the pressure into
two physically interpretable parts, the hydrostatic and hydrodynamic pressures, the
latter treated as a form of correction to the former. In consequence, the hydrostatic
part is computed explicitly from the free surface elevation and density field, whereas
the hydrodynamic pressure is found from a derived pressure Poisson equation (from
fractional formulation). This elegant method is associated with well-known fractional
step or projection methods of solving three-dimensional Navier-Stokes equations and
has been applied before only in finite difference methods. The main advantage of this
approach is that the final solution is much less sensitive to the well-posedness of the
boundary conditions for the hydrodynamic pressure than in the case of the global one.
In practice, very fine mesh spacing at boundaries is not required. In comparison to
other projection methods the physical interpretation of the hydrodynamic pressure yields
clearly defined and simple boundary conditions for the pressure Poisson equation. This
circumvents obstacles which other researchers confronted. Because in most free surface
flows the hydrostatic pressure gradient component usually dominates the hydrodynamic
one, the requirement that the projection procedure should only be a correction to the
intermediate solution for the velocity is fulfilled. This is a very important advantage,
e.g. compared to the application of a Poisson equation for the global pressure.
In the first step of the non-hydrostatic algorithm the hydrodynamic pressure is ex-
cluded from the consideration and the solution method follows the existing algorithm
of Telemac-3D with the important exception that the vertical Navier-Stokes equation
is treated as well. The advection and diffusion stages can be treated simultaneously
(using SUPG formulation for advective terms) or decoupled (using the method of char-
acteristics). In contrast to the previous hydrostatic algorithm, the driving forces such as
free surface gradients (barotropic pressure gradients), the density gradients (baroclinic
pressure gradients) and Coriolis force are treated in a fully three-dimensional manner in
the diffusion step.
The next two steps in the previous hydrostatic algorithm, the pressure-continuity steps,
based on shallow water equations and yielding barotropic velocity component, free sur-
face elevations and the vertical velocity component from the continuity equation, are
abandoned. They are replaced by the continuity and free surface steps of the new algo-
rithm.
The new continuity step is based on the pressure Poisson equation from the fractional for-
mulation. The provisional, intermediate solution, provided by the convection-diffusion
steps is corrected by the non-hydrostatic component. This component is computed from
the hydrodynamic pressure gradients under the assumption that the resulting final veloc-
ity must be divergence-free in the entire domain (incompressible flow) and satisfies ap-
propriate boundary conditions. In this step the formal velocity projection is performed.
The continuity step could be repeated iteratively, aiming for more exact fulfilment of the
continuity equation and boundary conditions for flows with hydrodynamic pressure gra-
6.2 Result discussion 201

dients comparable to the hydrostatic gradients, but one iteration proved to be sufficiently
exact.
Several techniques for finding the free surface position are implemented. Two of them
make use of the kinematic boundary condition, which is a hyperbolic equation for the free
surface elevation using the velocity at the surface as the advecting field: (1) a scheme
based on the method of characteristics and (2) on the semi-implicit SUPG method.
Another method is based on the vertically integrated continuity equation (the free surface
conservative equation) and solved using semi-implicit SUPG method or entirely explicit
formulation. Finally, for applications where the free surface elevation should be obtained
from the shallow water equations for some reason, but a non-hydrostatic correction
for the flow is needed, a modification of the previous pressure-free-surface step can be
applied.

6.2 Result discussion


The main attraction of the presented developments lie in the fact that they were made
in the framework of an existing, well-verified modelling system, which makes immediate
application possible in various related fields of the geophysical flows, such as water quality
or sediment transport.
The verification of the realised program developments which is documented in chapter 4,
ensure the quality of the software. Furthermore, a broad range of event-oriented verifica-
tion cases, presented in chapter 5, cover most of the intended extension of the application
domain of the initial hydrostatic model. The test cases allow critical assessment of the
properties of the new algorithm. The model is now capable of reproducing various physi-
cal phenomena, which were out of reach before. The main obstacles for flow simulations,
which must be treated without limiting approximations, are removed. The provided
evidence substantiate the claim that the developed model is an advanced, entirely three-
dimensional free surface model for incompressible flows with a broad application domain.

The main theoretical difficulties in the development of this code were encountered when
choosing the most appropriate approach for the pressure treatment and the free surface
computation, discussed in chapters 1 and 2. The ambiguities due to providing well-posed
boundary conditions excluded the application of an equation for the global pressure. Sim-
ilar difficulties were encountered when trying a global iterative solution using a pressure
correction term, as in the SIMPLE algorithm family. The finally successfully imple-
mented method of splitting the pressure into physically interpretable hydrostatic and
hydrodynamic parts (as e.g. applied by Casulli and Spalding [20]), which is a modifica-
tion of the projection method idea, suits the operator-splitting procedure of Telemac3D
well. The boundary conditions are clearly defined in this case. A projection-2 method
was implemented, but it did not show any advantages and no test cases are presented.
The method based on the pressure Poisson equation for the hydrodynamic pressure is
robust enough to yield good results even when the hydrodynamic pressure influence is
as large as the hydrostatic one.
202 6.2 Result discussion

For the free surface, the methods based on the semi-implicit SUPG for the kinematic
boundary condition and the conservative free surface equation proved to be fully satis-
factory. In contrary, the computations based on explicit formulation for the conservative
free surface show too large numerical damping or instabilities for larger Courant numbers.
The method of characteristics applied for kinematic BC, however unconditionally stable,
is the least mass conservative from all methods mentioned above, excluding it from ap-
plications with larger free surface variability. Methods based on marker-in-cell (MAC)
or volume-of-fluid (VOF) approaches were not considered. Their greatest advantage is
their ability to describe the most complex free surface geometries (e.g. breaking waves,
droplets). Their disadvantage is larger demand for machine memory and processor time,
as well as in difficulties in implementing specific boundary conditions such as surface
stresses. For the σ-mesh structure of Telemac3D, which excludes free surface shapes
which cannot be described by a single-valued function, the most attractive advantages
of these methods cannot be exploited and only the disadvantages remain.
As the verification cases show, the requirements for the boundary conditions for the free
surface at an open out- or inflow boundary are severe. Special boundary condition types
such as non-reflecting or absorbing boundary conditions are strongly recommended for
practical applications.
Other problems to be solved were connected with harmonisation of boundary conditions
applied for velocity in different parts of the operator-splitting algorithm and those for
free surface and the hydrodynamic pressure.
For successful model applications the following requirements are important:

1. universality of application for various purposes;

2. user-friendly peripheries;

3. sufficient computation precision;

4. possibly low computational effort.

From the point of view of model universality, the application domain of the developed
model is much wider than previously. The non-hydrostatic algorithm presented in this
work opens new possibilities for dealing with many free surface flow classes, which could
not be modelled properly with a hydrostatic model. As it has been shown, a larger do-
main of free surface and internal waves processes is now included (waves of various wave-
lengths, dispersion, reflexion, diffraction, interference). Compared to the wave models
(e.g. Boussinesq wave model [142]), the model yields not only the free surface configura-
tion (and fluid flux) but also the three-dimensional current structures underneath. The
density flows and various flows concerning vertical circulation, where vertical accelera-
tions play an important role, can be better simulated as well. Additionally, the model
delivers the global pressure field and therefore the interactions between free surface flows
and structures can be modelled.
Due to the fact that Telemac3D belongs to a larger, coupled family of models, us-
ing common subroutine libraries (Telemac–System, [73]), which is being continuously
6.3 Model applications and limitations 203

tested and developed in a number of institutes, the advantage of the non-hydrostatic


algorithm in many application domains can be taken immediately. Telemac sys-
tem includes libraries allowing modelling of transport (e.g. salinity, passive pollutants,
etc.) and applying various turbulence models, therefore the way is open for more ad-
vanced applications. These include simulation of sediment transport, pollution, water
quality, thermal impacts, more complex turbulence effects, particle-tracking and waves.
Being embedded in the overall Telemac system means that user-friendly post- and
pre-processing is available. However, through the new developments, the complexity of
the models and therefore the number of possible user errors have grown.
The code is written in such a way, that by switching one parameter in the input list, it
produces hydrostatic or non-hydrostatic results, so that a direct comparison of results
with identical input parameters (but, according to case, different boundary conditions)
is possible. Several test applications are presented for illustration of the effect of taking
into consideration or neglecting the hydrodynamic pressure and for model verification.
The computational effort compared to the previous, hydrostatic version is greater. Al-
though on one hand a larger algorithm step is abandoned (pressure – free surface –
continuity), on the other hand the diffusion-convection stage for the vertical velocity
component, the solution of the Poisson equation for the dynamic pressure and finding
the new free surface position are introduced. The required computational time can be
compared to the time consumed by the previous version with simultaneous transport of
two tracers (e.g. temperature, salinity).

6.3 Model applications and limitations


In practice, the developed model can be applied to those geophysical or engineering
problems, where a more proper reproduction of the physical flow properties, a high
accuracy level and an entirely three-dimensional approach is required. The price for this
precision is the higher computational cost. The applications are very important not only
for practical reasons, but also for gaining experience and defining further development
needs. The application of this model for the following fields should be straightforward:

• Wind and density driven flows in larger water bodies such as lakes with taking
complex bottom geometries (and steeper slopes) into consideration.

• Current (or waves) interaction with structures, especially in harbours or coastal


areas.

• Wave-induced currents over complex bottom topography.

• Density currents, for example the salt intrusions in estuaries, or convective circu-
lations in lakes or oceans.

• Internal waves of various kinds (e.g. for thermocline variability).


204 6.4 Recommended further developments

In this stage another, very important aspect of developing fully three-dimensional codes
must be mentioned. Many three-dimensional features of the flow can be parameterised
and introduced into two-dimensional models or even one-dimensional ones. Studies with
three-dimensional models allow a very effective parameterisation of these features in
form of coefficients or corrections introduced into models of lower complexity. In this
way three-dimensional phenomena can be taken into consideration in computationally
much more effective codes, which is of paramount importance for practical engineering
applications.
In the particular case of Telemac system, the interface between three-dimensional and
two-dimensional models is straightforward, because both use the same basic mesh. The
most important advantages of this approach can be found in the domain of sediment
transport, pollutant spreading, the interaction structures-currents and wave problems.

6.4 Recommended further developments


Further developments of the model presented here are listed in the sequence of dimin-
ishing importance. They could include:

• Mesh adaptivity. The developed algorithm is not exclusively applicable to the


σ-mesh consisting of prismatic element tiers. The algorithm should work prop-
erly for elements allowing meshing of arbitrary three-dimensional domains, as e.g.
tetrahedrons, which would open the possibility of mesh adaptation taking place in
all three directions. In the framework of the operator-splitting algorithm consist-
ing of steps which would require different mesh adaptivity from step to step, the
adaptivity criteria should probably be dictated by gradients of the hydrodynamic
variables. Development of efficient mesh generators or mesh refining/coarsening
routines would be required in order to allow continuous mesh adaptation to the
variable flow conditions and/or geometry.

• Non-reflecting boundary conditions for the open (i.e. artificial) boundaries of


the computational domain should be intensively investigated for various flow types.
In the model described in this work, the simplest forms of NRBC are applied
and absorbing boundary conditions were developed. The role of non-reflecting
boundary conditions is important in securing stable solutions for transient flows,
and for effectively converging computations of stationary flows as well. In more
advanced applications in the field of free surface and internal waves some kind of
these boundary conditions must always be imposed.

• Variable domain boundaries. The σ-transformation is a traditional solution


for the mesh adaptation to the movements of the free surface, which works well
for most typical geophysical applications. Improvements allowing application of
various vertical discretisation in different parts of a computational domain, non-
vertical lateral boundaries or introducing mesh levels defined by surfaces of equal
6.4 Recommended further developments 205

density (isopycnal co-ordinates) would allow more accurate and efficient computa-
tions for complex bottom topography and coastal structures and for the presence
of stratification. Moreover, solutions based on the Arbitrary Lagrangian-Eulerian
(ALE) method would probably allow mesh adaptation to boundary movements in
all three directions, opening the possibility of modelling interactions between fluid
and moving, deformable structures and in an arbitrary geometry. This approach
connected with specific free surface or volume tracking methods can also be ap-
plied for modelling of waves of larger amplitude, dam breaks, tidal flats, etc. The
required computational effort may, however, grow considerably.
• Stability estimates should be further investigated and built-in stability control
(e.g. adapting the time step) considered for various algorithm steps.
• Free surface algorithms. The free surface tracking algorithms are accurate, but
very sensitive to the imposed boundary conditions and less mass conservative than
volume-tracking methods. Alternative free surface algorithms, such as volume-
of-fluids or marker-in-cell, or especially mixed methods, should be investigated in
order to find equally accurate and stable solution methods comparable to or better
than the height function methods. These special algorithms can also be applied
to multiphase flows and for tracking internal interfaces between fluid/waters of
different properties.
• Overall algorithm structure. In this work, the decoupled method of solving
the hydrodynamic equations for free surface flows is applied. However, for tur-
bulence or transport modelling, the coupled approach may prove to be equally or
more effective, maybe not for the global algorithm, but for its parts. This can
be very attractive, especially if quickly converging methods for solving arbitrary
and large (non-linear) equation systems become available. On the other hand, the
projection-2 method should be further investigated. It is promising for flows where
hydrodynamic pressure gradients are comparable or greater than the hydrostatic
pressure ones.
• Turbulence modelling. The developments presented in this work have pur-
posefully been made independently of any actual turbulence modelling approach.
However, they provide a good basis for further development in this field. Although
FEM is still thought to be computationally too expensive for turbulence modelling
methods requiring very fine meshes, the advantage of the decoupled method applied
in this model is the possibility to use the simplest linear finite elements available.
For modelling turbulent flows in complex geometries the FEM (or FVM) approach
is thought to be the most appropriate one for unstructured meshes. The flexible
and open Telemac algorithm structure provides a means to investigate various
turbulence modelling methods. In Telemac3D constant eddy viscosity, mixing
length and k-ε turbulence models are already available and there are no obstacles
to developing further, more modern or application-specific ones. A decoupled finite
element method has already been applied successfully with Large Eddy Simulation
(LES), e.g. with an application for lakes [48].
206 6.4 Recommended further developments
Appendix A

Various derivations

A.1 Incompressibility
One of the most widely used approximations is to treat the water in the geophysical
flows as an incompressible fluid, i.e. a fluid whose density depends on temperature
and dissolved substances concentration, but not on the pressure. The most important
consequence of this approximation is that the velocity field in the incompressible flows
is non-divergent, i.e. solenoidal.
The mass conservation equation in its general form is:
1 d̺
+∇·u=0 (A.1)
̺ dt

Taking an equation of state ̺ = ̺(p, s, T ) into consideration, the density derivative in


(A.1) can be expanded as a function of state variables:
1 d̺ 1 ∂̺ dp 1 ∂̺ ds 1 ∂̺ dT
= + + (A.2)
̺ dt ̺ ∂p dt ̺ ∂s dt ̺ ∂T dt
which can be transformed to:
1 d̺ 1 dp α ds β dT
= 2 + + (A.3)
̺ dt ̺cs dt ̺ dt ̺ dt

where α and β are expansion coefficients with respect to the salinity and temperature
and cs is the speed of sound.
In the case of isentropic motion in the absence of viscous or diffusive effects, ds/dt = 0
and dT /dt = 0, and the mass conservation equation can be written as:
dp
(̺c2s )−1 +∇·u=0 (A.4)
dt
The simplest form of the isentropic motion is when the fluid element does not exchange
heat with the surroundings and retains fixed chemical composition.
After Batchelor [5] the motion is not isentropic when [5]:

207
208 A.2 Coriolis terms

1. Radiative energy exchange with the surroundings takes place.

2. Heat exchange through molecular conduction occurs: the flux of heat is given by
the temperature gradient −k∇T (k is the thermal conductivity).

3. Heating or cooling due to change of phase, chemical reaction or to viscous dissipa-


tion takes place.

Batchelor [5] gives the following three requirements to be fulfilled in order to neglect the
pressure derivative in (A.4), i.e. to treat the fluid as incompressible (in which changes
of the density with pressure are negligible):

1. The fluid particle velocity must be small compared with the speed of sound cs .

2. The phase speed (wavelength divided by period) of disturbances must be small


compared to the speed of sound cs .

3. The vertical scale of motion must be small compared with the scale height defined
as a mean value of ̺/|d̺/dz|.

The scale height defined above is much greater than the depth for most geophysical flows.
E.g. in the ocean this scale height is about 40 times greater than the mean ocean depth.
Speed of sound in seawater is cs ≈ 1500 m/s. The greatest speed of disturbances, the
phase speed of long gravitational waves, c = (gH)1/2 , is equal is about 200 m/s in the
ocean at 4 km depth. The assumption of the incompressibility of the water is therefore
appropriate for most hydrodynamic processes.
The usual incompressibility approximation is therefore to assume that the density is
independent of pressure and take account of the density changes due to temperature and
salinity changes through the equation of state only, without taking these changes into
consideration in the mass conservation equation. The velocity field in this approximation
is non-divergent (or solenoidal):
∇·u=0 (A.5)

So formulated, the approximation implies that volume rather than mass is conserved.
Indeed, the volume is conserved while the density changes as a function of temperature
and salinity (not pressure). Therefore this approximation should be removed from the
basic equations when the effect of the thermal expansion is to be reproduced exactly,
e.g. to cope with the water body surface position changes induced by heating or cooling
[114].

A.2 Coriolis terms


The Coriolis term arises in the momentum equation, because the mathematical descrip-
tion of the flows is formulated in a non-inertial co-ordinate system fixed to the earth,
A.2 Coriolis terms 209

which rotates around its axis. The equation of motion F = ma is valid, when a is mea-
sured in an inertial system, i.e. in a co-ordinate system, whose origin is not accelerating.
For practical reasons, as a reference inertial system, the system fixed relative to distant
stars is applicable.
The transformation from the fixed inertial system to the rotating system of the earth can
be found in most authoritative books on mechanics. The derivation is shortly repeated
here for completeness.
If i, j, k is a triad of orthogonal unit vectors in the non-inertial system fixed in the rotating
frame, then any vector s can be written in this system as s = s1 i + s2 j + s3 k. The rate
of change of the vector s with respect to the time t appears as a result of change of its
components in the moving system and due to the change of the unit vectors as the co-
ordinate frame rotates with the angular velocity Ω. It is assumed that no translational
acceleration of the non-inertial system appears. Ω is measured in the fixed, inertial
system. For the observer in the inertial frame:
! ! ! !e
ds X dsi di X dsi ds
= i + si = i + si Ω × i = +Ω×s (A.6)
dt i dt dt i dt dt

where the superscript e means relative to the rotating system. The same formula can be
applied for the velocity u of a fluid element on the surface of the earth (measured in the
inertial system): !e
dR
u= + Ω × R = ue + Ω × R (A.7)
dt
where, R the distance of the moving fluid element from the earth’s centre. Analogically,
for the acceleration a, measured in the absolute frame:
!e
du
a= + Ω × u = ae + Ω × u (A.8)
dt

It should be mentioned that Ω is the angular velocity of the earth’s rotation relative to
the inertial system fixed to the distant stars (Ω = 2π/86164 rad s−1 = 7.292×10−5 rads−1 ,
whereby 86164 s = 23 h 56 min 4 s is one sidereal or star day). (Ω measured in the
non-inertial system of the earth is obviously equal to 2π/86400 rad s−1 , i.e. 2π radians
per solar day, 24 h). Introduction of (A.7) to (A.8):
!e !e !e !
d dR dR
a= +Ω×R +Ω× +Ω×R (A.9)
dt dt dt

yields after some algebra:


!a !e
dua dΩ
a= = ae + × R + 2Ω × ue + Ω × (Ω × R) (A.10)
dt dt

The terms on the right hand side of (A.10) represent: (1) acceleration in the system
fixed to the earth, (2) apparent acceleration due to changes in the rotation velocity,
210 A.2 Coriolis terms

(3) apparent Coriolis acceleration, (4) apparent centripetal acceleration. It is usually


assumed that the earth rotates steadily and the second term is therefore ignored. The
equation of motion in an inertial system:
!
du 1
= − ∇p + gi + f (A.11)
dt i
̺

transforms into:
due 1
= − ∇p − 2Ω × ue + gi − Ω × (Ω × R) + f (A.12)
dt ̺
The transformation yields two new apparent forces per unit mass, appearing in the non-
inertial system of the earth, relating to the Coriolis and centripetal accelerations. The
true forces remain unchanged.
The centripetal acceleration Ω × (Ω × R) = −Ω2 R, required to make a body rotate
with the earth with the angular velocity Ω, and the gravitational acceleration gf have
the same direction approximately on the equator only. However, the maximum value
of the magnitude of the centripetal acceleration is only about 0.3% of the gravitational
acceleration. Therefore, the centripetal acceleration is treated in practice, in the co-
ordinate system fixed to the earth, as a small correction (in magnitude and direction)
to the gravitational acceleration. The difference gi − Ω × (Ω × R) is introduced in the
equations and is called the acceleration due to gravity g.
The local orthogonal Cartesian co-ordinate system is assumed on the rotating earth,
where x is directed eastward, y is northward, and z points vertically upward. The
earth’s angular rotation velocity vector Ω is at the geographical latitude ϕ inclined to
the y-direction (assumed here to be North). In such a system, the Coriolis acceleration
has the following components:
   
i j k 2wΩ cos ϕ − 2vΩ sin ϕ
2Ω × u =  0 2Ω cos ϕ 2Ω sin ϕ  =  2uΩ sin ϕ (A.13)
   

u v w 2uΩ cos ϕ

Introducing horizontal and vertical Coriolis coefficients fH = 2Ω cos ϕ and fV = 2Ω sin ϕ:


 
fH w − fV v
2Ω × u = 

fV u 
 (A.14)
fH u

Usually, when the horizontal velocity scale is significantly larger than the vertical one
and the hydrostatic approximation applies, just those terms, in which the horizontal
Coriolis coefficient fH = 2Ω cos ϕ appears, are neglected. Only terms with the vertical
one, fV = 2Ω sin ϕ remain. In this simplified case the Coriolis force has the following
components (f = fV ):
   
2wΩ cos ϕ − 2vΩ sin ϕ −f v
2Ω × u =  2uΩ sin ϕ  ≈  fu  (A.15)
   

2uΩ cos ϕ 0
A.3 Small amplitude free surface waves in a homogeneous fluid 211

For simulation of fully three-dimensional flows, when all motion directions and scales are
equally treated, the form of the Coriolis term (−2Ω × u) in the Navier Stokes equations
in the rotating frame (2.2) must also be fully three-dimensional. It should be noted,
that very close to the equator (ϕ = 0) the vertical Coriolis coefficient vanishes and the
horizontal prevails and care is needed in treatment of the rotational effects. Because of
this latitudinal dependence of the Coriolis parameter, the neglecting of terms with fH
are also called upper ocean approximation.
For simulation of larger areas (seas, ocean parts) using the rectangular co-ordinate sys-
tem, appropriate assumptions depicting the latitudinal variability of the Coriolis param-
eters and the curvature of the earth are made. The usual approximations are the β-plane
and the f -plane approximation. In both cases the earth’s curvature is neglected.
In the β-plane approximation the origin of the Cartesian, rectangular co-ordinate system
is centred at the latitude ϕ0 and a linear dependence of the vertical Coriolis parameter
on the y-co-ordinate is assumed in the form fV = f0 + βy, where f0 = 2Ω sin ϕ0 , β =
2Ω cos ϕ0 /R and fH = 2Ω cos ϕ0 . In mid-latitudes (|ϕ| ≈ 45o ) |f0 | ≈ 10−4 rad/s and
β ≈ 1.6 · 10−11 rad/s.
In the f -plane approximation the Coriolis parameters are assumed to be constant in the
entire computational domain. This approximation is valid as long as L cos ϕ0 /R << 1, L
being the horizontal motion scale, the condition needed in order to neglect the parameter
β.
The application limits of these approximations are as follows. When the horizontal
motion scale L is comparable with the radius R of the earth, the equations of motion
in spherical co-ordinates and Coriolis terms without any approximations must be used.
For intermediate scales up to L/R = O(10−1), the β-plane approximation is suitable,
whereas for smaller scales f -plane approximation is appropriate.

A.3 Small amplitude free surface waves in a homo-


geneous fluid

In this section the linear theory of free surface waves is reviewed. For simplification, the
non-rotating case (Ω = 0) is taken. A formal discussion of small amplitude wave theory
can be found in many books treating water waves, e.g. [99, 100]. Here only a sketch of
this theory is presented for completeness.
Let us consider a incompressible and inviscid fluid of constant density ̺ in an equilibrium
state and of uniform depth H (for instance, in a pool with a flat bottom). The position
of the free surface is z = 0 and bottom co-ordinate is z = −H, i.e., z axis points upward
and (x, y) axes form a horizontal plane. The atmospheric pressure is neglected, patm = 0.
The equilibrium pressure p0 (z) is given by:

p0 (z) = −̺gz (A.16)


212 A.3 Small amplitude free surface waves in a homogeneous fluid

Let us assume a small perturbation of the equilibrium state of the free surface, described
by a single-valued function:
z = η(x, y, t) (A.17)

The pressure is given by the sum of the hydrostatic and hydrodynamic parts π:

p = ̺(η − z) + π (A.18)

The perturbation pressure is given by:

p′ = p − p0 = ̺gη + π (A.19)

The perturbation causes a movement to appear. However, the perturbation quantities,


e.g. velocities, are enough small that their products can be neglected. The linearised
momentum conservation equations for small perturbation and continuity equation (2.7)
reduce to the following set of equations:

̺∂u/∂t = −∂p′ /∂x


̺∂v/∂t = −∂p′ /∂y
̺∂w/∂t = −∂p′ /∂z (A.20)

∂u/∂x + ∂v/∂y + ∂u/∂x = 0 (A.21)

Adding derivatives of the equations (A.20) and using (A.21) results in obtaining the
Laplace equation for p′ :
∇2 p ′ = 0 (A.22)

The boundary condition at the flat bottom is:

w = 0 for z = −H (A.23)

The kinematic boundary condition (2.6.3.1) at the surface applies (condition that a
particle at the free surface z = η remains there):

d(η − z) ∂η ∂η ∂η
= +u +v −w =0 (A.24)
dt ∂t ∂x ∂y

(A.24) reduces for small perturbations to:


∂η
w= for z = η (A.25)
∂t

When the surface tension is neglected (i.e. capillarity effects excluded), and the atmo-
spheric pressure set to zero, the pressure must vanish at the surface. The boundary
condition for the perturbation pressure at the free surface is:

p = p0 + p′ = 0 ⇒ p′ = ̺gη for z = η (A.26)


A.3 Small amplitude free surface waves in a homogeneous fluid 213

i.e. perturbation pressure near the surface is hydrostatic. Combining the third equation
of (A.20) with (A.23) one obtains the boundary condition for perturbation pressure at
the bottom:
∂p′ /∂z = 0 for z = −H (A.27)

Because the difference between z = η and z = 0 is small, the boundary conditions at the
free surface (A.25–A.26) can also be applied for z = 0.
The solution of this problem depends on the initial perturbation of the free surface.
In following it is assumed, that the basic disturbance has a sinusoidal form with an
amplitude η0 , frequency ω and wavenumber vector k = (k1 , k2 ) (κ = |k| = 2π/λ, λ being
the wavelength):
η = η0 cos(k1 x + k2 y − ωt) = η0 cos φ (A.28)

and arguing that an arbitrary disturbance can be described as a superposition of such


sinusoidal waves. Assuming that the perturbation pressure π is proportional to η in the
form of (A.28), the Laplace equation (A.22) yields

∂ 2 p′
− κ2 p′ = 0 (A.29)
∂z 2

The solution of (A.29) is to be searched in the form of a combination of exponential


functions in the imaginary domain. Taking the boundary conditions (A.26) and (A.27)
one obtains:
ch[κ(z + H)]
p′ = ̺gη0 cos(k1 x + k2 y − ωt) (A.30)
ch(κH)

and from (A.20):


k1 ch[κ(z + H)]
u = gη0 cos(k1 x + k2 y − ωt) (A.31)
ω ch(κH)
k2 ch[κ(z + H)]
v = gη0 cos(k1 x + k2 y − ωt) (A.32)
ω ch(κH)
κ sh[κ(z + H)]
w = gη0 sin(k1 x + k2 y − ωt) (A.33)
ω ch(κH)

In order to satisfy the kinematic boundary condition at the free surface (A.25) for z =
η ≈ 0, the following condition must be satisfied:

ω 2 = gκ tanh(κH) (A.34)

which is the well-known dispersion relation determining the frequency and phase speed
(celerity) c = ω/κ of gravitational waves.
As can be seen from (A.30), perturbation pressure decreases with depth, and:

π(x, −H, t) 1
= (A.35)
π(x, 0, t) cosh(κH)
214 A.3 Small amplitude free surface waves in a homogeneous fluid

This ratio decreases rapidly with increasing κH. When κH << 1 perturbation pressure
does not change with the depth.
For k2 = 0, the movement takes place in the (x, z)-plane. The tip of the velocity vector
describes an ellipse; the ratio between the vertical and horizontal axes of this ellipse is:
|w|
= tanh[κ(z + H)] (A.36)
|u|

The ellipse flattens with the depth and the ratio (A.36) vanishes at the bottom for
z = −H as w = 0. For κH >> 1, i.e. in deep water |w|/|u| ≈ 1 and the ellipse reduces
to a circle.
Two very important approximations for gravitational waves apply. They are formulated
with the length scale being the fluid equilibrium depth H.

1. Short waves for κ−1 << H, called also deep-water waves. The dispersion relation
(A.34) can be approximated to:
ω 2 = gκ (A.37)

and the perturbation pressure is from (A.30):


p′ = ̺gη0 exp(κz) cos(kx + ly − ωt) (A.38)

The pressure perturbation, having a maximal value at the surface, vanishes expo-
nentially with the depth. The phase speed c is a function of κ, i.e., short waves
are dispersive. The velocity components are:
k1
u = gη0 exp(κz) cos(k1 x + k2 y − ωt) (A.39)
ω
k2
v = gη0 exp(κz) cos(k1 x + k2 y − ωt) (A.40)
ω
κ
w = gη0 exp(κz) sin(k1 x + k2 y − ωt) (A.41)
ω
2. Long waves for κ−1 >> H, also called shallow-water waves. The dispersion relation
(A.34) is again approximated to:

ω 2 = κ2 gH (A.42)

The perturbation pressure is:


p′ = ̺gη0 cos(kx + ly − ωt) (A.43)

The pressure perturbation is independent of depth. The long waves are non-
dispersive, because their phase speed c = (gH)1/2 is independent of the wavenum-
ber. The velocity components are:
k1
u = gη0 cos(k1 x + k2 y − ωt) (A.44)
ω
A.4 Equations of state for seawater 215

k2
v = gη0 cos(k1 x + k2 y − ωt) (A.45)
ω
κ2
w = gη0 (H + z) sin(k1 x + k2 y − ωt) (A.46)
ω
The horizontal velocity component is independent of the water depth and the
vertical component changes linearly with the depth.

Because the considered fluid has a constant density ̺ (perturbation of density is zero),
the perturbation pressure for the long wave case (A.43) has the same value that would be
obtained if the pressure were calculated from the hydrostatic equation. The horizontal
velocity component does not change with the water depth and the influence of the free
surface movement reaches down to the bottom. Therefore, for the fluid of constant den-
sity, the hydrostatic approximation and the long-wave (or shallow-water) approximations
are equivalent.

A.4 Equations of state for seawater

A.4.1 EOS–80: UNESCO equation of state


The UNESCO EOS–80 (International Equation of State for Seawater, 1980) equation of
state for seawater is an internationally accepted standard for geophysical flows, defined
in 1980 by the Joint Panel on Oceanographic Tables and Standards (UNESCO, 1981)
[160]. The density of seawater ̺ (kg/m3 ) is treated as a function of the practical salinity
s [PSU], temperature T (o C) and pressure p [bar] (e.g. for deep-sea applications). The
equation is formulated in an algorithmic form; the relevant parts of the computations
can be omitted when e.g. pressure can be assumed constant. The salinity is given
as a dimensionless variable in practical salinity units, PSU 1 . This unit is for practical
purposes almost identical with ppt (parts per thousand).
This equation is valid for the following ranges of parameter variability:

T = −2 to 40o C,
s = 0 to 42,
p = 0 to 1000 bar,
1
’Practical salinity, 1978, definition: The practical salinity, symbol S, of a sample of seawater,
is defined in terms of the ratio K15 of the electrical conductivity of the seawater sample at the
temperature of 15o C and the pressure of one standard atmosphere, to that of a potassium chloride
(KCl) solution, in which the mass fraction of KCl is 32.4356 × 10−3 , at the same temperature and
pressure. The K15 value exactly equal 1 corresponds, by definition, to a practical salinity exactly
equal to 35. The practical salinity is defined in terms of the ratio K15 by the following equation:
1/2 3/2 2 5/2
S = 0.0080 − 0.1692K15 + 25.3851K15 + 14.0941K15 − 7.0261K15 + 2.7081K15 , formulated and
adopted by the Unesco/ICES/SCOR/IAPSO Joint Panel on Oceanographic Tables and Standards...etc.
...This equation is valid for a practical salinity S from 2 to 42.’ [160]
216 A.4 Equations of state for seawater

with an exactness of 3.5 · 10−6 (standard error).

In the first step the reference density of pure water is computed as a function of temper-
ature T :
̺w = a0 + a1 T + a2 T 2 + a3 T 3 + a4 T 4 + a5 T 5 . (A.47)
a0 = + 999.842594,
a1 = + 6.793952e-2,
a2 = – 9.095290e-3,
a3 = + 1.001685e-4,
a4 = – 1.120083e-6,
a5 = + 6.536332e-9.
Then the density by one standard atmosphere (i.e. by normal atmospheric conditions at
the free surface, p = 0) is obtained as a function of salinity s:

̺(T, s, 0) = ̺w + (b0 + b1 T + b2 T 2 + b3 T 3 + b4 T 4 )s + (c0 + c1 T + c2 T 2 )s3/2 + d0 s2 . (A.48)

b0 = + 8.24493e-1, c0 = – 5.72466e-3,
b1 = – 4.0899e-3, c1 = + 1.0227e-4,
b2 = + 7.6438e-5, c2 = – 1.6546e-6,
b3 = – 8.2467e-7,
b4 = + 5.3875e-9, d0 = + 4.8314e-4.
Finally, the influence of pressure p can be taken into consideration by:

̺(T, s, p) = ̺(s, T, 0)/(1 − p/K(s, T, p)) (A.49)

where K is the compressibility of seawater: (K = −1/V dV /dp = 1/̺d̺/dp ) which is


obtained from a procedure containing polynomials with empirical constants. It should be
noted that the reference densities used in the equation of state should not be identified
with the reference average density ̺0 appearing due to the Boussinesq approximation in
the momentum conservation equation (section 2.1.4).

A.4.2 Simple forms of the equation of state


The density variations can be treated as a function of active tracer concentrations:
∆̺
= f (T1 , ..., Tn ) (A.50)
̺0

The tracer concentrations are transported according to the transport equation (2.3).
For small variations of density and the tracer concentrations, a linearised form of the
equation of state can be applied:
∆̺ X
=− βi (T − T0 )i (A.51)
̺0 i
A.4 Equations of state for seawater 217

Figure A.1: Mesh used in various verification cases for a square domain (Ekman profiles,
free surface tests, waves reflection, wind-driven circulation, standing wave in a closed
basin, etc.) with 2601 nodes and 5000 elements. The mesh size is scaled from case to
case.

where βi are dilatation coefficients (in the simplest case constant, for example positive
for temperature or negative for salinity) and (T0 )i are reference values for tracers i.

For particular applications, some optimised equations of state can be obtained. For
example, for thermal transport in North-European stratified estuaries of lower salinity,
the following equation of state in form ̺ = ̺(T, s) proved to be accurate enough [101]:

̺ = ̺0 [1 − αT (T − T0 )2 + αs s] (A.52)

where ̺0 = 999.972 kgm−3 , T0 = 4o C, αT = 7.0 · 10−6 K−2 , αs = 0.75 · 10−3 PSU−1 .


218 A.4 Equations of state for seawater

Figure A.2: The channel basic mesh with 1060 nodes and 1806 elements applied in the
lock exchange flow test, as well as standing wave in a channel and the flow over a ramp
tests. Dimensions scaled from case to case.

Figure A.3: The channel mesh with 1147 nodes and 2114 elements used in the channel
with a bump test.
A.5 Meshes 219

A.4.3 Suspended sediment influence on the water density

The suspended sediment concentration can also be taken as a parameter affecting water
density and stratification. The dependence of the fluid density on the dry mass sediment
concentration c [kg/m3 ] is taken as:

̺sed − ̺(p, S, T )
̺(p, S, T, c) = ̺(p, S, T ) + c (A.53)
̺sed

where ̺(p, S, T ) is the water density obtained from the EOS 80 equation of state and
̺sed is the sediment constituent density [kg/m3 ].

A.5 Meshes
The basic (2D) meshes applied in various test and verification cases presented in chapters
4 and 5 are shown in figures A.1, A.2 and A.3. For all test cases, the horizontal mesh
spacing is uniform all over the computational domain. Only in three verification cases
(Ekman profiles, section 4.1, interfacial internal waves, section 5.3 and waves over an
underwater channel, section 5.8) the variable spacing in the vertical direction is used.
In all other cases, the per cent distribution of the mesh levels is uniform, without larger
discrepancies in the vertical and horizontal resolution.

A.6 Systematic approach to model development


A model, understood as an abstraction of reality, a mental, physical or purely mathe-
matical representation or description of a system, can be implemented on a computer
only if the model can be converted into an algorithm. The process of developing and
implementation of a computer model consists of physical system definition, definition of
the mental or conceptual model, its algorithmic implementation and finally, expressing
it in terms of usable computer program, i.e. software.
The definition of the physical system to be described is the fundamental step in the de-
velopment process determining further capabilities of the model. It consists of collection
and description of physical processes or phenomena relevant to the purpose of the model.
The modelling process requires that these physical mechanisms and relevant effects of
the model environment are formulated in terms of model variables. Usually in this stage,
the formulation of the conceptual model being a mathematical (or verbal, logical) rep-
resentation of the system, usually several approximations and model assumptions are
introduced.
The computer model requires that the conceptual model must be expressed in terms of a
set of prescriptions or rules suitable for computation: an algorithm. The last development
stage is the implementation of the algorithm in the form of a computer program, involving
220 A.6 Systematic approach to model development

not only coding of the algorithm using a programming language, but also designing data
structures, and pre- and post processing of input and output data.
The results of the development process determine the functionality of a computer model.
The functionality is described usually in terms of how realistic the results in various
applications are, and in terms of the response produced by the physical phenomena or
events included in the model. Additionally, it concerns its ability to be adjusted for
various applications.
A computer or numeric model can be understood as a function transforming input data
from a certain domain into output data. Input data can be categorised as: system param-
eters, describing the modelled physical system, algorithmic parameters which configure
the algorithmic implementation (e.g. time-step) and software parameters (e.g. machine
constants). Special attention must be paid to the system parameters. These include, for
example, data which describe the configuration or properties of the modelled system. In
a numeric hydrodynamic model these are: the geometry of the domain, fluid properties
assumed to be fixed (e.g. fluid incompressibility), and parameters describing the way
the boundary conditions are imposed.
Data is incorporated at different development stages of the model, influencing the as-
sumptions and approximations introduced in the conceptual model. These assumptions
result in various constraints set upon the model parameters. For example, certain as-
sumptions depicting the system geometry may lead, when the water depth is small
compared to the horizontal scale of the domain, to the formulation of the hydrodynamic
model in terms of vertically integrated shallow water equations. These require, in turn,
that the data describing the boundary conditions must also be vertically integrated or
that the bottom gradients must be small.
A very important set of parameters includes those describing the initial condition or state
of the system, as: initial velocities, free surface levels or concentrations of transported
substances. Last, but not least is the data describing the variable boundary conditions,
or environment of the model, such as wind influence or turbulence model parameters,
applying on boundaries.
Due to the high development costs most advanced hydrodynamic models are developed
with the aim to cover an application domain as broadly as possible within the limits
of their assumptions. They are usually designed to operate flexibly on a wide variety
of system parameters. It is usually the case to obtain models for specific applications
by narrowing the scope of a broader, complex, generic model with situation-specific
datasets. However, for simple physical systems, models dedicated only to a specific
situation are also developed and applied.
The model validation process can be defined as a process of testing and documenting
the quality and functionality of a developed computer model in relation to its intended
applications and the physical system it represents [35]. The overall aim of the model
validation process is therefore to determine the model domain of application, its reliabil-
ity in describing the physical system it represents, and eventually to improve the model.
The validation process begins during model development and continues until the model
is operational. Broadly speaking, every activity involving a given model, every appli-
A.6 Systematic approach to model development 221

cation, yields information about the model’s quality. Usually, a documented validation
process also describes the model’s functionality.
So, broadly defined, the validation process begins with a series of verifications of model
properties, assumptions, criteria, and results, together covering as much as possible of
the intended application domain. Verification is a test of the model’s performance and
accuracy in a well-defined situation or case, when the true, real results are known [172].
It shows whether the mathematically formulated problem is properly posed and solved.
As mentioned above, the numeric model consists of three basic elements, a conceptual
model and its algorithmic and software implementations. Leaving aside the software
verification, concerned with testing if the software actually computes exactly what is
prescribed and algorithm verification, being a series of tests whether the algorithm rep-
resents the conceptual model or not, let us concentrate on the third component: the
conceptual model itself.
Formally, the verification of the conceptual model can be divided into the event-oriented
verification, where the physical phenomena included in the model are tested, and the
application-oriented verification, where the applicability of the model in different oper-
ational situations or even real conditions is tested.
The event-oriented verification tests are concerned with systematic modelling of separate
physical processes, often in idealised conditions and with complicating factors eliminated,
and with confirmation of whether or not the model behaves according to expectations.
For example, the model is adapted to reproduce an analytical solution, or an experimen-
tal measurement series. A strict model verification must provide a comparison of the
result of the numerical solution of the equations describing the physical system with an
analytical solution of the given test problem or with a dataset obtained by an appropriate
laboratory experiment or measurements in nature.
The application-oriented verification is based on the testing of accuracy and quality of
different model applications, where e.g. the entire scope of the model’s capabilities is
used, and the results can be compared with measurements. It is also possible to verify a
model comparing its results with the output of another, more general, complex and well
validated model, representing the same system.
One of the most typical parts of verification is sensitivity analysis, helping to assess the
response of the model to the parameter variations. It helps to determine which param-
eters are crucial to the accuracy and functionality of the model, as well as to formulate
practical guidelines for the applicability of the model. (This is also an important step of
model calibration, understood as adjustment of model parameters to a given application
in order to obtain improved results.)
A true validation in the case of a hydrodynamic numeric model for geophysical flows,
i.e. for natural flows, is based on tests which are designed to find out how closely the
output of a verified and possibly calibrated model represents the real environment. It
requires, that the model output is compared with field measurements in situ which have
not been used in the calibration process. Therefore, validation is possible only if there
is knowledge about the natural geophysical system, which has not already been built
into the model. The validation activities are concerned in this case with the question
222 A.6 Systematic approach to model development

whether or not, for a given application, the conceptual model meets the requirements.
The results of the overall validation process must provide the user with the ability to
assess the model’s reliability and accuracy with respect to given applications as well as
the model ability to be adjusted to specific conditions.
Bibliography

[1] Abbot, M., 1979. Computational hydraulics. Elements of the theory of free surface flow.
Pitman, London.

[2] Arnold, H., 1996. Simulation dammbruchinduzierter Flutwellen. PhD thesis, Universität
Hannover. Institut für Strömungsmechanik und ERiB.

[3] Aziz, K., and Hellums, J., 1967. Numerical solution of the three-dimensional equations
of motion for laminar natural convection. Physics of Fluids, 10(2), 314–324.

[4] Barr, D., 1967. Densimetric exchange flow in rectangular channels. III. Large scale
experiments. La Houille Blanche, 22, 619–632.

[5] Batchelor, G., 1967. An introduction to fluid dynamics. Cambridge Univ. Press, London
and New York. 615 pp.

[6] Behr, M., and Tezduyar, T., 1994. Finite element strategies for large-scale flow simula-
tions. Computer Methods in Applied Mechanics and Engineering, 112, 3–24.

[7] Benjamin, T., 1968. Gravity currents and related phenomena. Journal of Fluid Mechan-
ics, 31(2), 209–248.

[8] Benqué, J.-P., Hauguel, A., and Viollet, P.-L., 1982. Engineering applications of computa-
tional hydraulics, Vol. II. Numerical models in environmental fluid mechanics, M.B. Ab-
bot and J.A. Counge, Eds. Pitman, London. 160 pp.

[9] Bleck, R., Hanson, H., Hu, D., and Kraus, R., 1989. Mixed layer – thermocline interaction
in a three–dimensional isopycnic coordinate model. Journal of Physical Oceanography,
19, 1417–1439.

[10] Bleck, R., Rooth, C., Hu, D., and Smith, L., 1992. Salinity-driven thermohaline tran-
sients in wind- and thermohaline-forced isopycnic coordinate model of the North Atlantic.
Journal of Physical Oceanography, 22, 1486–1515.

[11] Blumberg, A., and Mellor, G., 1987. A description of a three-dimensional coastal ocean
circulation model. In Three-dimensional coastal ocean models, N. Heaps, Ed., vol. 4 of
Coastal and estuarine sciences. American Geophysical Union, Washington, D.C., pp. 1–
16.

[12] Boussinesq, J., 1823. Essai sur la théorie des eaux courants. Mém. prés. par div. savants
à l’Acad. Sci. Paris, 23, 380–398.

[13] Boussinesq, J., 1903. Théorie analytique de la chaleur, vol. 2. Gauthier–Villars, Paris.

223
224 Bibliography

[14] Braess, D., 1997. Finite Elemente: Theorie, schnelle Löser und Anwendungen in der
Elastizitätstheorie. Springer Verlag, Berlin Heidelberg New York Tokyo. 2nd Ed., 317pp.

[15] Brezzi, F., and Fortin, M., 1991. Mixed and hybrid finite element methods. Springer–
Verlag, Berlin Heidelberg New York Tokyo. 350pp.

[16] Brooks, A., and Hughes, T., 1982. Streamline Upwind / Petrov-Galerkin formulations for
convection dominated flows with particular emphasis on the incompressible Navier-Stokes
equations. Computer Methods in Applied Mechanics and Engineering, 32, 199–259.

[17] Buddha Śakyamuni, recorded in 1st century AC in Ceylon. Suttapitaka. In Pali Canon,
vol. AN 3,65 (Tika Nipata, Mahavagga, Sutta 65). Translated from Pali to English by
Soma Thera.

[18] Bulgarelli, U., Casulli, V., and Greenspan, D., 1984. Pressure methods for the numerical
solution of free surface fluid flows. Pineridge Press, Swansea, U.K.

[19] Casulli, V., 1990. Semi-implicit finite difference methods for the two-dimensional shallow
water equations. Journal of Computational Physics, 86, 56–74.

[20] Casulli, V., 1995. Recent developments in semi-implicit numerical methods for free sur-
face hydrodynamics. In Advances in hydro-science and -engineering, vol. 2, Tsinghua
University Press, Beijing, pp. 2174–2181.

[21] Casulli, V., 1997. Numerical simulation of three-dimensional free surface flow in isopycnal
co-ordinates. International Journal for Numerical Methods in Fluids, 25, 645–658.

[22] Casulli, V., and Cattani, E., 1994. Stability, accuracy and efficiency of a semi-implicit
method for three-dimensional shallow water flow. Computers Math. Applic., 27(4), 99–
112.

[23] Casulli, V., and Cheng, R., 1992. Semi-implicit finite difference methods for three-
dimensional shallow water flow. International Journal for Numerical Methods in Fluids,
15, 629–648.

[24] Casulli, V., and Stelling, G., 1995. Simulation of three-dimensional, non-hydrostatic
free-surface flows for estuaries and coastal seas. In Estuarine and Coastal Modeling.
Proceedings of the 4th International Conference, San Diego, M. Spaulding and R. Cheng,
Eds., ASCE, New York, pp. 1–12.

[25] Casulli, V., and Stelling, G., 1998. Numerical simulation of 3D quasi-hydrostatic, free
surface flows. J. Hydraulic Engineering, 124(7), 678–686.

[26] Chen, D., 1997. Numerische Simulation von Strömungsvorgängen mit der Arbitrary
Lagrangian-Eulerian Method (ALE-Methode). PhD thesis, Rheinisch-Westfälische Tech-
nische Hochschule Aachen. Mitteilungen Band 110, Lehrstuhl und Institut für Wasserbau
und Wasserwirtschaft.

[27] Cheng, R., Casulli, V., and Gartner, J., 1993. Tidal, Residual, Intertidal Mudflat (TRIM)
model and its applications to San Francisco Bay, California. Estuarine, Coastal and Shelf
Science, 36, 235–280.
Bibliography 225

[28] Chorin, A., 1967. A numerical method for solving incompressible viscous flow problem.
Journal of Computational Physics, 2, 12–26.

[29] Chorin, A., 1968. Numerical solution of the Navier-Stokes equations. Math. Comp., 22,
745–762.

[30] Chorin, A., and Marsden, J., 1993. A mathematical introduction to fluid mechanics.
Springer-Verlag, New York. 3rd ed. (1st: 1979).

[31] Chung, T., 1978. Finite element analysis in fluid mechanics. McGraw–Hill, New York.

[32] Clément, A., 1996. Coupling of two absorbing boundary conditions for 2D time-domain
simulations of free surface gravity waves. Journal of Computational Physics, 126, 139–
151.

[33] Davies, A., and Lawrence, J., 1994. A three-dimensional model of the M4 tide in the
Irish Sea: The importance of open boundary conditions and influence of wind. Journal
of Geophysical Research, 12(C8), 16197–16227.

[34] Deardoff, J., 1970. A numerical study of three-dimensional turbulent channel flow at
large Reynolds numbers. Journal of Fluid Mechanics, 41(part 2), 453–480.

[35] Dee, D., and van der Marel, M., 1991. Validation of computer models. Research report,,
Delft Hydraulics. 31 pp.

[36] Donea, J., 1983. Arbitrary Lagrangian-Eulerian Finite Element Methods. Computational
Methods for Transient Analysis, 1, 101–119.

[37] Donea, J., 1984. A Taylor-Galerkin method for convective transport problems. Interna-
tional Journal for Numerical Methods in Engineering, 20, 101–119.

[38] Donea, J., Quartapelle, L., and Selmin, V., 1987. An analysis of time discretization in
the finite element solution of hyperbolic problems. Journal of Computational Physics,
70, 463–499.

[39] Donea, J., S., G., Laval, H., and Quartapelle, L., 1982. Finite element solution of the
unsteady Navier-Stokes equations by a fractional step method. Computer Methods in
Applied Mechanics and Engineering, 30, 53–73.

[40] Druet, C., 1995. Dynamika stratyfikowanego oceanu. Wydawnictwo Naukowe PWN,
Warszawa. (in Polish).

[41] Ekman, V., 1905. On the influence of the Earth’s rotation on ocean-currents. Ark. Math.
Astron. Fys., 2(11), 1–52.

[42] Enquist, B., and Majda, A., 1977. Absorbing boundary conditions for the numerical
simulation of waves. Mathematics of Computation, 31(139), 629–651.

[43] Farmer, J., Martinelli, L., and Jameson, A., 1994. Fast multigrid method for solving
incompressible hydrodynamic problems with free surfaces. AIAA Journal, 32, 1175–
1182.

[44] Ferziger, J., and Perić, M., 1996. Computational methods for fluid dynamics. Springer–
Verlag, Berlin Heidelberg New York.
226 Bibliography

[45] Flather, R., 1976. Results from surge prediction model of the North–West European
continental shelf for April, November and December 1973. Report, 24–1976, Institute of
Oceanography, U.K.

[46] Fletcher, C., 1988. Computational techniques for fluid dynamics 1. Fundamental and
general techniques. Springer–Verlag, Berlin Heidelberg New York.

[47] Fletcher, C., 1988. Computational techniques for fluid dynamics 2. Specific techniques
for different flow categories. Springer–Verlag, Berlin Heidelberg New York.

[48] Forkel, C., 1995. Die Grobstruktursimulation turbulenter Strömungs- und Stoffaus-
breitungsprozesse in komplexen Geometrien. PhD thesis, Rheinisch-Westfälische Tech-
nische Hochschule Aachen. Mitteilungen Band 102, Lehrstuhl und Institut für Wasserbau
und Wasserwirtschaft.

[49] Fortin, M., 1981. Old and new finite elements for incompressible flow. International
Journal for Numerical Methods in Fluids, 1, 347–364.

[50] Frederiksen, C., and Watts, A., 1981. Finite-element method for time-dependent incom-
pressible free surface flow. Journal of Computational Physics, 39, 282–304.

[51] Gaarthuis, J., 1995. A non-hydrostatic pressure model for shallow water flow. Tech. rep.,
Delft Hydraulics Research Report.

[52] Galland, J.-C., Goutal, N., and Hervouet, J.-M., 1991. TELEMAC: A new numerical
model for solving shallow water equations. Adv. Water Resources, 14(3), 138–148.

[53] Gill, A. E., 1982. Atmosphere-ocean dynamics. Academic Press, San Diego, Ca. 662 pp.

[54] Girault, V., and Raviart, P.-A., 1986. Finite elemant methods for Navier–Stokes Equa-
tions. Theory and algorithms. Springer–Verlag, Berlin Heidelberg New York Tokyo.
374pp.

[55] Glowinski, R., and Pironneau, O., 1992. Finite element methods for Navier-Stokes equa-
tions. Annu. Rev. Fluid Mech., 24, 167–204.

[56] Goutal, N., 1989. Finite element solution for the transcritcal shallow-water equation.
Mathematical Methods in the Applied Sciences, 11, 503–524.

[57] Gresho, P., 1990. On the theory of semi-implicit projection methods for viscous incom-
pressible flow and its implementation via a finite element method that also introduces
a nearly consistent mass matrix. Part 1: Theory. International Journal for Numerical
Methods in Fluids, 11, 587–620.

[58] Gresho, P., 1991. Incompressible fluid dynamics: some fundamental formulation issues.
Annu. Rev. Fluid Mech., 23, 413–453.

[59] Gresho, P., 1991. Some current CFD issues relevant to the incompressible Navier-Stokes
equations. Computer Methods in Applied Mechanics and Engineering, 87, 201–252.

[60] Gresho, P., and Chan, S., 1990. On the theory of semi-implicit projection methods
for viscous incompressible flow and its implementation via a finite element method that
also introduces a nearly consistent mass matrix. Part 2: Implementation. International
Journal for Numerical Methods in Fluids, 11, 621–659.
Bibliography 227

[61] Gresho, P., Chan, S., and Christon, M., 1994. An equal-order approximate projection
FEM. Report, UCRL-JC-115897, Lawrence Livermore National Laboratory.

[62] Gresho, P., Chan, S., Christon, M., and Hindmarsh, A., 1995. A little more on stabilized
Q1 Q1 for transient viscous incompressible flow . International Journal for Numerical
Methods in Fluids, 21, 837–856.

[63] Gresho, P., Chan, S., Lee, R., and Upson, C., 1984. A modified finite element method
for the time-dependent incompressible Navier–Stokes equations Part 1: Theory. Inter-
national Journal for Numerical Methods in Fluids, 4, 557–598.

[64] Gresho, P., and Sani, R., 1987. On pressure boundary conditions for the incompressible
Navier–Stokes equations. International Journal for Numerical Methods in Fluids, 7,
1111–1145.

[65] Harlow, F., and Welch, J., 1965. Numerical calculation of time-dependent viscous incom-
pressible flow of fluid with free surface. The Physics of Fluids, 8(12), 2182–2189.

[66] Hassanzadeh, S., Sonnad, V., and Foresti, S., 1994. Finite element implementation of
boundary conditions for the pressure Poisson equation of incompressible flow. Interna-
tional Journal for Numerical Methods in Fluids, 18, 1009–1019.

[67] Hauguel, A., 1980. Adaptation of tidal numerical models to shallow water wave problems.
In Proceedings 17th Coastal Engineering Conference, Sydney, Australia.

[68] Heaps, N., Ed., 1987. Three-dimensional coastal ocean models, vol. 4 of Coastal and
estuarine sciences. American Geophysical Union, Washington, D.C.

[69] Hedstrom, G., 1979. Nonreflecting boundary conditions for nonlinear hyperbolic systems.
Journal of Computational Physics, 30, 222–237.

[70] Helmig, R., 1996. Gekoppelte Strömungs- und Transportprozesse im Untergrund. Ein
Beitrag zur Hydrosystemmodellierung. Habilitationsschrift, Fakultät für Bauingenieur-
und Vermessungswesen, Universität Stuttgart. 327 pp.

[71] Helmig, R., 1997. Multiphase flow and transport processes in the subsurface: a contri-
bution to the modeling of hydrosystems. Springer Verlag, Heidelberg, Berlin, New York.
367 pp.

[72] Hervouet, J.-M., 1997. Calcul symbolique de matrices en éléments finis avec MAPLE V.
Tech. Rep. HE-43/97/031/A, Electricité de France (EDF-DER), Département Labora-
toire National d’Hydraulique.

[73] Hervouet, J.-M., and Janin, J.-M., 1997. Guide for programming in the TELEMAC
system version 3.2. Tech. Rep. HE-43/97/015/B, Electricité de France (EDF-DER),
Département Laboratoire National d’Hydraulique.

[74] Hervouet, J.-M., and van Haren, L., 1997. Recent advances in numerical methods for
fluid flows. Tech. Rep. HE-43/97/023/A, Electricité de France (EDF-DER), Département
Laboratoire National d’Hydraulique.

[75] Higdon, R., 1986. Absorbing boundary conditions for the numerical simulation of waves.
Mathematics of Computation, 47(176), 437–459.
228 Bibliography

[76] Higdon, R., 1986. Numerical absorbing boundary conditions for the wave equation.
Mathematics of Computation, 49(179), 65–90.

[77] Hinkelmann, R., 1997. Parallelisierung eines Lagrange-Euler-Verfahrens für Strömungs-


und Stofftransportprozesse in Oberflächegewässern. PhD thesis, Universität Hannover.
Bericht Nr. 51/1997, Institut für Strömungsmechanik und ERiB.

[78] Hirt, C., and Nichols, B., 1980. Adding a limited compressibility to incompressible
hydrocodes. Journal of Computational Physics, 34, 390–400.

[79] Hirt, C., and Nichols, B., 1981. Volume of Fluid (VOF) Method for the dynamics of free
boundaries. Journal of Computational Physics, 39, 201–225.

[80] Huerta, A., and Liu, W., 1988. Viscous flow with large free surface motion. Computer
Methods in Applied Mechanics and Engineering, 69, 277–324.

[81] Hughes, T., Ferencz, R., and Hallquist, J., 1987. Large-scale vectorized implicit calcu-
lations in solid mechanics on a Cray X-MP/48 utilizing EBE preconditioned conjugate
gradients. Computer Methods in Applied Mechanics and Engineering, 61, 215–248.

[82] Hughes, T., Franca, L., and Balestra, M., 1986. A new finite element formulation for
computational fluid dynamics: V. Circumventing Babuska-Brezzi Condition: A stable
Petrov-Galerkin formulation of the Stokes problem accomodating equal-order interpola-
tion. Computer Methods in Applied Mechanics and Engineering, 59, 85–99.

[83] Janin, J.-M., Lepeintre, F., and Pechon, P., 1992. TELEMAC-3D: A finite element code
to solve 3D free surface flow problems. Tech. Rep. HE-42/92.07, Electricité de France
(EDF-DER), Département Laboratoire National d’Hydraulique.

[84] Jankowski, J., Malcherek, A., and Zielke, W., 1996. Numerical modelling of suspended
sediment due to deep-sea mining. Journal of Geophysical Research, 101(C2), 3545–3560.

[85] Jankowski, J., and Zielke, W., 1995. Mesoskalige Stofftransporte im Pazifik als Folge des
Tiefseebergbaus. Final report, Institut für Strömungsmechanik und ERiB, Universität
Hannover. 86 pp.

[86] Jiang, C., and Kawahara, M., 1993. The analysis of unsteady incompressible flows by a
three-step finite element method. International Journal for Numerical Methods in Fluids,
16, 793–811.

[87] Johnson, C., 1992. Numerical solution of partial differential equations by the finite
element method. Cambridge Univ. Press, London and New York. 278 pp.

[88] Kármán, T., v., 1940. The engineer grapples with non-linear problems. Bull. Am. Math.
Soc., 46, 615–683.

[89] Kawahara, M., and Miwa, T., 1984. Finite element analysis of wave motion. International
Journal for Numerical Methods in Engineering, 20, 1193–1210.

[90] Kheshgi, H., and Scriven, L., 1984. Penalty finite element analysis of unsteady free
surface flows. Finite Elements in Fluids, 5, 393–434.

[91] Kofler, M., 1994. Maple V Release 2. Addison-Wesley, Bonn. 513 pp.
Bibliography 229

[92] Krauss, W., Ed., 1973. Methods and results of theoretical oceanography. Vol. 1: Dynam-
ics of the homogeneous and the quasihomogeneous ocean. Gebrüder Borntraeger, Berlin.
302 pp.

[93] Kwak, D., Chang, J., Shanks, S., and Chakravarthy, S., 1986. A three-dimensional
incompressible Navier-Stokes Flow solver using primitive variables. AIAA Journal, 24(3),
390–396.

[94] Lafaurie, B., Nardone, C., Scardovelli, R., Zaleski, S., and Zanetti, G., 1994. Modelling
merging and fragmentation in multiphase flows with SURFER. Journal of Computational
Physics, 113, 134–147.

[95] Laitone, E., 1960. The second approximation to cnoidal and solitary waves. Journal of
Fluid Mechanics, 9, 430–444.

[96] Lamb, H., 1932. Hydrodynamics. Cambridge Univ. Press.

[97] Lang, G., 1990. Zur Schwebstoffdynamik von Trübungszonen in Ästuarien. PhD thesis,
Universität Hannover. Bericht Nr. 26/1990, Institut für Strömungsmechanik und ERiB.

[98] Lavelle, J., and Baker, E., 1994. A numerical study of local convection in the benthic
ocean induced by episodic hydrothermal discharges. Journal of Geophysical Research,
99(C8), 16065–16080.

[99] Le Méhauté, B., 1976. Un introduction to hydrodynamics and water waves. Springer
Verlag, New York. 315 pp.

[100] LeBlond, P., and Mysak, L., 1978. Waves in the ocean. Elsevier Oceanography Series
20. Elsevier, Amsterdam–Oxford–New York. 602 pp.

[101] Lehfeldt, R., 1991. Ein algebraisches Turbulenzmodell für Ästuare. PhD thesis, Univer-
sität Hannover. Bericht Nr. 30/1991, Institut für Strömungsmechanik und ERiB.

[102] Lemos, C., 1992. Wave breaking - a numerical study. No. 71 in Lecture Notes in Engi-
neering. Springer–Verlag, Berlin Heidelberg New York.

[103] LeNormant, C., Lepeintre, F., Teisson, C., Malcherek, A., Markofsky, M., and Zielke, W.,
1993. Three dimensional modelling of estuarine processes. In MAST Days and Euromar
Market, Project Reports Volume 1, K.-G. Barthel, M. Bohle-Carbonell, C. Fragakis, and
M. Weydert, Eds., Brussels.

[104] Leschziner, M., and Rodi, W., 1979. Calculation of a strongly curved open channel flow.
J. of the Hydraulics Division, 105(HY10), 1297–1314.

[105] Lilek, Ž., 1996. Ein Finite-Volumen-Verfahren zur Berechnung von inkompressiblen und
kompressiblen Strömungen in komplexen Geometrien mit beweglichen Rändern und freien
Oberflächen. PhD thesis, Universität Hamburg. Bericht Nr. 563, Institut für Schiffbau.

[106] Löhner, R., Morgan, K., Peraire, J., and Zienkiewicz, O., 1985. Recent developments in
FEM–CFD. Lecture Notes in Physics, 238, 236–254.

[107] Madsen, O., and Mei, C., 1969. The transformation of a solitary wave over an uneven
bottom. Journal of Fluid Mechanics, 39(4), 781–791.
230 Bibliography

[108] Mahadevan, A., Oliger, J., and Street, R., 1996. A nonhydrostatic mesoscale ocean
model. Part I: Well-posedness and scaling. Journal of Physical Oceanography, 26(9),
1868–1880.

[109] Mahadevan, A., Oliger, J., and Street, R., 1996. A nonhydrostatic mesoscale ocean
model. Part II: Numerical implementation. Journal of Physical Oceanography, 26(9),
1881–1900.

[110] Malcherek, A., 1995. Mathematische Modellierung von Strömungen und Stoff-
tranportprozessen in Ästuaren. PhD thesis, Universität Hannover. Institut für
Strömungsmechanik und ERiB.

[111] Malcherek, A., Markofsky, M., Zielke, W., Peltier, E., Le Normant, C., Teisson, C., Cor-
nelisse, J., Molinaro, P., Corti, S., and Grego, G., 1996. Three dimensional numerical
modelling of cohesive sediment transport processes in estuarine environments. Final Re-
port to the EC Contract MAS2-CT92-0013, Institut für Strömungsmechanik und ERiB,
Universität Hannover, Electricité de France (EDF), Delft Hydraulics, ENEL.

[112] Marcos, F., and Janin, J.-M., 1997. Nouveaux developpements dans l’etape de convection-
diffusion de TELEMAC-3D. Tech. Rep. HE-42/94/025/A, Electricité de France (EDF-
DER), Département Laboratoire National d’Hydraulique.

[113] McCowan, J., 1894. On the highest wave of a permanent type. Phil. Mag. (Ser 5), 38,
351–358.

[114] Mellor, G., and Ezer, T., 1995. Sea level variations induced by heating and cooling: An
evaluation of the Boussinesq approximation in the ocean models. Journal of Geophysical
Research, C10, 20,565–20,577.

[115] Miyata, H., 1986. Finite-difference simulation of breaking waves. Journal of Computa-
tional Physics, 65, 179–214.

[116] Mizukami, A., and Hughes, T., 1985. A Petrov-Galerkin finite element method for
convection-dominated flows: an accurate upwinding technique for satisfying the max-
imum principle. Computer Methods in Applied Mechanics and Engineering, 50, 181–193.

[117] Nakayama, T., and Mori, M., 1996. An Eulerian finite element method for time-dependent
free surface problems in hydrodynamics. International Journal for Numerical Methods
in Fluids, 22, 175–194.

[118] Navier, C., 1823. Mémoire sur les lois du mouvement des fluides. Mém. d’Acad. de
Science, 6, 389–416.

[119] Nieuwstadt, F., Eggels, J., Janssen, R., and Pourquié, M., 1994. Direct and large-eddy
simulations of turbulence in fluids. Future Generation Computer Systems, 10, 189–205.

[120] Nihoul, J., and Jamart, B., Eds., 1987. Three-dimensional models of marine and estuarine
dynamics, vol. 45 of Elsevier Oceanography Series. Elsevier, Amsterdam. 629 pp.

[121] Noh, W., and Woodward, P., 1976. SLIC (Simple line interface calculation). In Pro-
ceedings of the Fifth International Conference on Numerical Methods in Fluids, A. van
Dooren and P. Zandbergen, Eds., no. 59 in Lecture Notes in Physics, Plenum Press,
pp. 330–340.
Bibliography 231

[122] Orlanski, I., 1976. A simple boundary condition for unbounded hyperbolic flow. Journal
of Computational Physics, 21, 251–269.

[123] Patankar, S., 1980. Numerical heat transfer and fluid flow. Hemisphere, New York.

[124] Patankar, S., and Spalding, D., 1972. A calculation procedure for heat, mass and mo-
mentum transfer in three-dimensional parabolic flows. Int. J. Heat Mass Transfer, 15,
1787.

[125] Phillips, N., 1957. A coordinate system having some special advantages for numerical
forecasting. Journal of Meteorology, 14, 184–185.

[126] Pond, S., and Pickard, G., 1983. Introductory dynamical oceanography. Pergamon Press,
Oxford. 329 pp.

[127] Press, W., Flannery, B., Teukolsky, S., and Vetterling, W., 1989. Numerical recipes - the
art of scientific computing. Cambridge University Press.

[128] Puzyrewski, R., and Sawicki, J., 1987. Podstawy mechaniki plynów i hydrauliki.
Państwowe Wydawnictwo Naukowe, Warszawa. 332 pp. (in Polish).

[129] Quartapelle, L., 1993. Numerical solution of the incompressible Navier-Stokes equations.
Birkhäuser, Basel, Boston, Berlin.

[130] Ramaswamy, B., 1987. Arbitrary Lagrangian-Eulerian finite element method for un-
steady, convective, incompressible viscous free surface fluid flow. International Journal
for Numerical Methods in Fluids, 7, 1053–1075.

[131] Ramaswamy, B., 1987. Lagrangian finite element analysis applied to viscous free surface
fluid flow. International Journal for Numerical Methods in Fluids, 7, 953–984.

[132] Ramaswamy, B., 1990. Numerical simulation of unsteady viscous free surface flow. Jour-
nal of Computational Physics, 90, 396–430.

[133] Ramshaw, J., and Messina, G., 1991. A hybrid penalty–pseudocompressibility method
for transient incompressible fluid flow. Comput. Fluids, 20, 165–175.

[134] Reynolds, O., 1895. On the dynamical theory of incompressible viscous fluids and the
determination of the criterion. Philosophical Transactions of the Royal Society of London,
A186, 123–164.

[135] Rudman, M., 1997. Volume tracking methods for interfacial flows. Int. J. Numer. Methods
Fluids, 24, 671–691.

[136] Sander, B., Wolf-Gladrow, D., and Olbers, D., 1995. Numerical studies of open ocean
convection. Journal of Geophysical Research, 100(C10), 20579–20600.

[137] Sani, R., and Gresho, P., 1994. Résumé and remarks on the open boundary condition
minisymposium. International Journal for Numerical Methods in Fluids, 18, 983–1008.

[138] Sani, R., Gresho, P., Lee, R., Griffiths, D., and Engelmann, M., 1981. The cause and
cure(?) of spurious pressures caused by certain FEM solutions of the incompressible
Navier-Stokes equations. International Journal for Numerical Methods in Fluids, 1, 171–
204.
232 Bibliography

[139] Schaper, H., 1985. Ein Beitrag zur numerischen Berechnung von nichtlinearen kurzen
Flachwasserwellen mit verbesserten Differenzenverfahren. PhD thesis, Universität Han-
nover. Bericht Nr. 21/1991, Institut für Strömungsmechanik und ERiB.

[140] Schröder, P.-M., 1997. Zur numerischen Simulation turbulenter Freispiegelströmungen


mit ausgeprägt dreidimensionaler Charakteristik. PhD thesis, Rheinisch-Westfälische
Technische Hochschule Aachen. Mitteilungen Band 108, Lehrstuhl und Institut für
Wasserbau und Wasserwirtschaft.

[141] Schröter, A., 1991. Das numerische Seegangsmodell BOWAM2 1990 - Grundlagen und
Verifikationen -. Bericht Nr. 31/1991, Institut für Strömungsmechanik und ERiB, Uni-
versität Hannover.

[142] Schröter, A., 1995. Nichtlineare zeitdiskrete Seegangssimulation im flachen und tief-
eren Wasser. PhD thesis, Universität Hannover. Bericht Nr. 42/1991, Institut für
Strömungsmechanik und ERiB.

[143] Shen, J., 1993. A remark on the projection-3 method. International Journal for Numer-
ical Methods in Fluids, 16, 249–253.

[144] Shyy, W., Udaykumar, H., Rao, M., and Smith, R., 1996. Computational fluid dynamics
with moving boundaries. Series in computational and physical processes in mechanics
and thermal sciences. Taylor & Francis, Washington. 285 pp.

[145] Spiegel, E., and Veronis, G., 1960. On the Boussinesq approximation for a compressible
fluid. Astrophys. J., 131, 442–447.

[146] Stokes, G., 1845. On the theories of the internal friction of fluids motion and of the
equilibrium and motion of elastic solids. Trans. Cambridge Phil. Soc., 8, 287–305.

[147] Stokes, G., 1847. On the theory of oscillatory waves. Trans. Cambridge Phil. Soc., 8,
441–455.

[148] Tamamidis, P., Zhang, G., and Assanis, D., 1996. Comparison of pressure-based and
artificial compressibility methods for solving 3D steady incompressible viscous flows.
Journal of Computational Physics, 124, 1–13.

[149] Tezduyar, T., 1991. Stabilized finite element formulations for incompressible flow com-
putations. Advances in Applied Mechanics, 28, 1–44.

[150] Tezduyar, T., Behr, M., and Liou, J., 1992. A new strategy for finite element computa-
tions involving moving boundaries and interfaces – The deforming-spatial-domain/space-
time procedure: I. The concept and the preliminary numerical tests. Computer Methods
in Applied Mechanics and Engineering, 94, 339–351.

[151] Tezduyar, T., Behr, M., Mittal, S., and Liou, J., 1992. A new strategy for finite ele-
ment computations involving moving boundaries and interfaces – The deforming-spatial-
domain/space-time procedure: II. Computation of free-surface flows, two-liquid flows,
and flows with drifting cylinders. Computer Methods in Applied Mechanics and Engi-
neering, 94, 353–371.

[152] Thomas, T., and Leslie, D., 1992. Development of a conservative 3D free surface code.
Journal of Hydraulic Research, 30(1), 107–115.
Bibliography 233

[153] Thomas, T., Leslie, D., and Williams, J., 1995. Free surface simulations using a conser-
vative 3D code. Journal of Computational Physics, 116, 52–68.

[154] Thompson, K., 1987. Time dependent boundary conditions for hyperbolic systems. Jour-
nal of Computational Physics, 68, 1–24.

[155] Thompson, K., 1990. Time dependent boundary conditions for hyperbolic systems, II.
Journal of Computational Physics, 89, 439–461.

[156] Tomé, M., and McKee, S., 1993. GENSMAC: An updated marker and cell technique for
free surface flows in general domians. PhD thesis, University of Strathclyde, Department
of Mathematics. Strathclyde Mathematics Research Report (1993) 11, March 1993.

[157] Tomé, M., and McKee, S., 1994. GENSMAC: A computational marker and cell method
for free surface flows in general domains. Journal of Computational Physics, 110(1),
171–186.

[158] Tu, J., and Fuchs, L., 1995. Calculation of flows using three-dimensional overlapping grids
and multigrid methods. International Journal for Numerical Methods in Engineering, 38,
259–282.

[159] Turner, J., 1973. Buoyancy effects in fluids. Cambridge Univ. Press. 367 pp.

[160] Unesco, 1987. International Oceanographic Tables, Vol. 4. Unesco technical papers in
marine science, No. 40.

[161] Versteeg, H., and Malalasekera, W., 1995. An introduction to computational fluid dy-
namics. The Finite Volume Method. Longman Scientific & Technical.

[162] Vreugdenhil, C., and Korren, B., Eds., 1993. Numerical methods for advection-diffusion
problems, vol. 45 of Notes on Numerical Mechanics. Vieweg, Braunschweig. 373 pp.

[163] Walker, S., and Watts, R., 1995. A three-dimensional numerical model of deep ventilation
in temperate lakes. Journal of Geophysical Research, 100(C11), 22711–22731.

[164] Wall, W., Bischoff, M., and Ramm, E., 1998. Stabilization techniques for fluid and
structural finite elements. In Computational Mechanics – New Trends and Applications,
World Congress on Computational Mechanics, Buenos Aires, Argentina, 29 June – 2
July 1998, S. Idelsohn, E. Oñate, and E. Dvorkin, Eds., CIMNE, Barcelona, Spain.

[165] Wambersie, O., and Crochet, M., 1994. Transient finite element method for calculating
steady state three-dimensional free surfaces. International Journal for Numerical Methods
in Fluids, 14, 343–360.

[166] Wei-Yan, T., 1992. Shallow water hydrodynamics : mathematical theory and numerical
solution for a two-dimensional system of shallow water equations, vol. 55 of Elsevier
Oceanography Series. Elsevier. 434 pp.

[167] Yih, C.-S., 1980. Stratified flows. Academic Press, London, New York. 418 pp.

[168] Yost, S., 1995. Three-dimensional non-hydrostatic modeling of free-surface flows and
transport of cohesive sediment. PhD thesis, University of Michigan.
234 Bibliography

[169] Youngs, D., 1982. Time.dependent multi-material flow with large fluid distortion. In
Numerical methods for fluid dynamics, K. Morton and M. Baines, Eds. Academic Press,
New York, pp. 273–285.

[170] Zalesak, S., 1979. Fully multi-dimensional flux-corrected transport algorithms for fluid
flow. Journal of Computational Physics, 11, 38–69.

[171] Zielke, W., 1992. Strömungsmechanik für Bauingenieure. Institut für Strömungsmechanik
und ERiB, Universität Hannover. Lecture Script.

[172] Zielke, W., 1996. Wegweiser durch numerische modelle. In Numerische Modelle von
Flüssen, Seen und Küstengewässern, Deutscher Verband für Wasserwirtschaft und Kul-
turbau (DVWK). Workshop 23.–24.09.1996.

[173] Zienkiewicz, O., and Taylor, R., 1991. The finite elemant method. McGraw-Hill, London,
New York. Fourth edition, Vol. I, Vol. II.
Nomenclature 235

Nomenclature
Abbreviations
Symbol Meaning
1D, 2D, 3D one-, two-, three-dimensional
BC boundary condition
EBE element-by-element (method)
EDF Electricité de France
eq. equation
FE, FD, FV finite elements, finite differences, finite volumes
FEM, FDM, FVM finite element (difference, volume) method
IC initial condition
LBB Ladyzhenskaya-Babuška-Brezzi (condition)
LNH Laboratoire National d’Hydraulique (EDF)
NRBC non-reflecting boundary condition
PSU practical salinity unit
SUPG streamline upwind Petrov-Galerkin (scheme)

Geometrical coordinates
Symbol Meaning
i imaginary unit
x, y, z Cartesian coordinates, x eastwards, y northwards, z vertical upwards
x = (x1 , x2 , x3 ) Cartesian coordinate vector
z∗ Vertical coordinate in σ transformation
B(x, y) Bottom coordinate (time-independent)
S(x, y, t), s(x, y, t) Position of the free surface (time-dependent)

Mathematical and numerical symbols, variables and constants


Symbol Meaning
a(u, v) bilinear form
ch(x) hyperbolic cosine, ch(x) = (ex + e−x )/2
f a variable (in algorithms)
f˜ intermediate value of variable f
f a, f d f variable value after addvection (a) or diffusion (d)stage
fi a value of the f variable with the index i
fn f variable value at the n-th time-step
∂f /∂n gradient of a variable normal to the boundary
n unit vector normal to a surface
sech(x) hyperbolic secant, 1/ch(x)
sh(x) hyperbolic sine, sh(x) = (ex − e−x )/2
Continuation on the next page
236 Nomenclature

Continuation from the previous page


C Courant number
H k (Ω) Sobolev space
H −m (Ω) dual space for a Sobolev space H m (Ω)
I unit tensor
|J| Jacobian (of a transformation)
L2 (Ω) space of all square-integrable functions f → R
R3 real, three-dimensional space
V′ dual space to V
Vh finite-dimensional subspace of V
α, β, γ local coordinates in prismatic elements
θ implicitness (Crank-Nicholson) coefficient
δΩ, Γ domain boundary
ξ, η local coordinates in 2D elements
λi eigenvalues
ϕi interpolation functions (finite elements)
ψi test or weighting functions (finite elements)
∆t time-step
∆x spacing in the x-direction
Ω domain
Ω0 reference domain, reference element volume, basic mesh
|| · || norm
(·, ·) scalar product
A, L, P, Q operators
A, u matrices (tensors), vectors
aij , ui elements of matrices (tensors), vectors

Physical variables and constants


Symbol Unit Meaning
c kg/m3 tracer concentration (e.g. pollutant)
c m/s wave celerity, or concentration
c0 m/s long (shallow water) wave celerity
f N kg−1 force per mass
fH , fV rad/s horizontal, vertical Coriolis coefficients
g, g m/s2 gravitational acceleration, its value
h m water depth, h = S − B
k m−1 wave vector
p N/m2 global pressure
pH N/m2 hydrostatic pressure
p0 N/m2 equilibrium pressure (wave theories)
p′ N/m2 disturbance pressure (wave theories)
Continuation on the next page
Nomenclature 237

Continuation from the previous page


3
q m /s fluid flux
s PSU salinity
t s time
u = (u, v, w) m/s velocity and its components
un m/s normal velocity component
ut m/s tangential velocity component
D m/s2 Navier-Stokes (deformation rate) tensor
Fr 1 Froude number
H m equilibrium water depth
H, h m solitary wave: wave height and undisturbed water depth
N rad/s buoyancy (Brunt-Väisälä-) frequency
R m Earth’s radius
R N/m2 Reynolds stress tensor
Re 1 Reynolds number
o
T C temperature
Ti kg/m3 tracer concentration
T s wave period
η m free surface variation from the equilibrium position
η0 m wave amplitude
λ m wave length
µi kg/ms dynamic viscosity (component), i = x, y, z
ν m2 /s kinematic viscosity (or diffusivity) tensor
νi m2 /s kinematic viscosity (or diffusivity) (component), i = x, y, z
π N/m2 hydrodynamic pressure
̺ kg/m3 fluid density
̺0 kg/m3 reference fluid density
̺∗ kg/m3 mean fluid density (Boussinesq)
∆̺ kg/m3 density variation (̺ − ̺0 )
σ N/m2 stress tensor
τ N/m2 shear stress
τxy N/m2 stress force per surface xy
φ rad phase
ω rad/s angle velocity or radial frequency
ω s−1 vorticity
Ψ m2 /s stream function
Ω rad/s rotation velocity of the Earth, 2π/86164 rad/s
Index

acceleration free surface, 15, 24, 55, 58


centripetal, 208 gradient, 20
Coriolis, 208 lateral, 15, 58
due to gravity, 208 normal vector, 59, 88
gravity, 14, 208 open, 70
reduced gravity, 139, 145 stress, 62
vertical, 2, 19–21, 72, 197 bottom shear, 61, 63, 107
algorithm wind shear, 61, 63, 107, 111, 134
hydrostatic, 22, 26 stress tensor, 59
advection step, 23 boundary condition, 4, 57
diffusion step, 23 absorbing, 68, 200
pressure – free surface – continuity numeric realisation, 103
step, 24, 55, 198 advection step, 97, 100
non-hydrostatic, 41, 71, 77, 79 cyclic, 65
advection step, 74, 96, 98 dynamic, 59, 70
continuity step, 75, 113 hydrodynamic pressure, 69, 70, 114,
diffusion step, 74, 89, 107 198
free surface stabilisation, 105, 162, impermeability, 61, 70, 89
165 inflow, 63, 164, 169
free surface step, 76, 100, 117, 199 inviscid free surface normal stress, 61
Poisson equation for hydrodynamic kinematic, 8, 54, 57, 76, 117, 162, 199,
pressure, 75, 86, 198, 199 210
source terms, 74, 94, 107, 198 numeric realisation, 100
velocity projection, 75, 88, 199 numeric realization, 100
regula falsi, 115 no-flux, 69
SIMPLE, 44 no-slip condition, 62
approximation non-reflecting, 65, 169, 175, 200
β-plane, 209 Higdon, 66
f -plane, 209 Orlanski, 66
Boussinesq, 17, 62 Sommerfeld radiation condition, 65
hydrostatic, 2, 5, 19, 21, 70, 139, 142, Thompson, 67
208, 209 numeric implementation
incompressibility, 15, 42, 205, 206 advection with characteristics, 97
validity conditions, 206 advection with SUPG, 100
rigid lid, 8, 9, 62 diffusion step, 94
Dirichlet, 88, 100, 102
boundary, 15, 58 Neumann, 87
bottom, 15, 24, 55, 58 open, 63
gradient, 3, 20, 21 outflow, 64, 164, 169
bottom gradient, 164 periodic, 65

238
Index 239

rigid lid, 8, 9, 62, 76, 137 conservation, 153


slip condition, 62 kinematic, 145
tracer, 69 kinetic, 155
potential, 145, 155
co-ordinate equation
inertial system, 207 advection
isopycnic, 7 σ-transformed, 97
non-inertial system, 14, 18, 206 Bernoulli, 164
rectangular system, 209 conservative free surface, 24, 54, 55, 57,
system 76, 126, 142, 163, 199
transformed, 57 explicit numeric realisation, 103
coefficient semi-implicit numeric realisation, 104
Coriolis, 18, 108, 208 continuity, 18
condition incompressible, 16, 46
for boundedness and continuity, 35 integrated, 25
incompressibility, 38, 40 conventional pressure eq., 41
inf-sup, 35 derived pressure eq., 41
LBB, 39, 40, 197 diffusion
convection σ-transformed, 89
deep ocean, 9 geostrophic, 108
current global system, 35, 36
buoyancy-driven, 3, 21, 95, 141 element-by-element, 37
drift, 107, 110, 111 matrix assembly, 36
gravity, 3, 21, 95, 141 solution, 36
in geostrophic equilibrium, 107 hydrodynamic set, 14
mesoscale ocean, 9, 209 for incompressible flows, 18, 71
secondary, 3, 8 in characteristic form, 67
slope, 107, 108, 113 Laplace, 114, 143, 210
wind-driven, 11, 110, 111, 134 mass conservation, 14, 18
deep water renewal, 8 momentum conservation, 14, 18
density with Boussinesq approximation, 18
average, 17 Navier-Stokes, 2, 14
equation of state, 14, 18, 139, 146 Reynolds-averaged, 16
EOS–80, 213 Navier-Stokes eqs., 15
sediment influence, 217 of state, 14, 15, 18, 139, 146
simple forms, 214 EOS–80, 213
reference, 17 Poisson, 33, 113
variance, 17 Poisson equation
domain numeric formulation, 86
computational, 15, 79 Poisson for hydrodynamic pressure, 47,
variable extents, 7, 56 198
triangulation, 79 Poisson for pressure, 41, 44
down-welling, 7, 8 Poisson for pressure from fractional for-
mulation, 45, 198
eigenvalue, 67, 115 shallow water eqs., 2, 21, 22
eigenvectors, 67 shallow water eqs., 2D, 24, 76, 199
Ekman spiral, 11 St.-Venant, 24, 76
energy tracer transport, 14, 18
240 Index

transport, 14 test f., 30


wave eq., 65, 143 weighting f., 30

finite element, 32 hydrothermal discharges, 8


2D reference element, 84
conform, 33 initial condition, 4, 57, 58
iso-parametric, 32
reference element, 81 matrix
flow advection, 99, 102
around structures, 21 diffusion, 88, 91, 92
channel, 11 mass, 88, 91, 99, 102
inviscid, 156, 164, 168 SUPG, 99, 102
stationary, 165, 166 mesh
subcritical, 164, 166 σ-mesh, 7, 11, 57, 79, 197, 200
supercritical, 164, 190, 191 σ-transformation, 7, 23, 57, 79
fluid diffusion step, 90
incompressible, 15, 205 Jacobian, 80, 84, 92
Newtonian, 14 method
force advection
buoyancy, 2, 17, 20, 24, 49, 95, 198 characteristics, 6, 23, 25, 67, 74, 76,
Coriolis, 9, 18, 20, 24, 96, 107, 198, 206 96, 100, 119, 123, 162
other forcing, 96 SUPG, 23, 25, 74, 76, 98, 119, 123,
form 162
bilinear, 34 Arbitrary Lagrangian-Eulerian (ALE),
V -elliptic, 35 7, 10, 56
formula deforming space-time FEM, 57
Green’s, 34, 87, 91 element-by-element, 6, 37
formulation for Navier-Stokes eqs., 27
weak, 30, 32–35 artificial compressibility, 7, 29
derivative, 89, 94, 103 coupled, 6, 10, 28
free surface, 2, 4, 50 decoupled, 7, 9, 29, 43
conservative equation, 8, 55, 76, 126, direct, 6, 28
142, 163, 199 fractional step, 7, 29
method non-primitive variables, 7, 29
height function, 8, 54, 190 operator splitting, 7, 23, 29, 41, 74,
line segment, 55 85, 197
marker-in-cell (MAC), 8, 50, 200 penalty function, 7, 29
volume of fluid (VOF), 8, 51, 200 pressure correction, 7, 29
representation, 50 projection, 7, 29, 42
tracking, 50 projection-1, 45
function projection-2, 45, 46, 48, 78
approximation f., 30 projection-3, 45
base f., 30 Taylor, 7, 29
Lagrangian interpolation f., 31, 36, 86, general
98, 101 finite difference, 27, 28
sign, 143 finite element, 3, 28, 28, 30
square integrable f., 32 finite element, equal interpolation, 40
step f., 143 finite element, Galerkin, 31, 87, 90
Index 241

finite element, mixed interpolation, projection, 43


38 splitting, 23, 41, 74, 85, 197
finite element, Petrov-Galerkin, 31, stability, 86
98, 101
finite element, standard, 31 parameter
finite element, SUPG, 31, 98, 101 Coriolis, 108, 208
finite volume, 27 depth of frictional influence, 109
space-time FEM, 10 upwind, 32, 102
spectral, 28 pressure
weighted residue, 30 atmospheric, 48, 62, 70, 108
Navier-Stokes eqs., 8 baroclinic, 5, 21, 22, 24, 49, 62, 76, 95,
model, 217 198
application domain, 11, 199, 218 barotropic, 4, 5, 21, 22, 24, 49, 50, 62,
calibration, 1 76, 94, 198
conceptual, 11, 217 decomposition, 9, 47, 198
for 3D shallow water eqs., 5 equation
functionality, 218 boundary conditions, 42, 46, 69, 114,
198
hydrostatic, 2
conventional, 41
non-hydrostatic, 8
derived, 41
parameters
for hydrodynamic pressure, 47, 198
algorithmic, 218
from fractional formulation, 45, 198
software, 218
equilibrium, 209
system, 218
gradient, 22, 48
physical, 1
hydrodynamic, 2, 47, 48, 155, 198, 210
physical system, 217
hydrostatic, 21, 47, 210
three-dimensional, 2
from local density, 49
two-dimensional, 2
from mean density, 48
validation, 219
kinematic, 42
process, 219
perturbation, 142, 154, 155, 210
verification, 1, 129, 219
Poisson equation, 9, 44
application-oriented, 219
event-oriented, 219 river meanders, 3
momentum
conservation, 153 scalar product
Hilbert space, 32
norm Sobolev space, 33
dual space, 34 solution
Hilbert space, 32 classical, 34
Sobolev space, 33 weak, 34
number space
Courant, 118, 119, 119, 126 dual, 34
Froude, 163 for pressure, 38
Reynolds, 15, 20 for velocity, 38
Rossby, 20 Hilbert, 32
Sobolev, 33
operator Stokes problem, 37
inverse, 35 subspace
isomorphic, 35 finite dimensional, 33, 38
242 Index

Telemac, 201, 202 sediment, 217


Telemac2D, 3, 5, 25, 76 temperature, 214
Telemac3D, 3, 6, 11, 22, 26, 197, 201 transport
tensor sediment, 3, 201
Navier-Stokes, 59 triangulation, 32
turbulent eddy-viscosity, 14 Trim2D, 5
turbulent stress, 14, 16, 59 Trim3D, 5
incompressible, 16 turbulence
test Boussinesq eddy-viscosity, 14, 16, 20,
bottom steepness, 130, 191 59
channel flow
over a bump, 130, 166 up-welling, 7, 8
sub- and supercritical over a ramp,
vectorisation, 37, 89
130
velocity
subcritical and supercritical over a
divergence-free, 44, 46, 78
ramp, 163
Ekman profile, 107
drift current, 110
Ekman spiral, 107
free surface breaking, 130, 190
intermediate, 43–45, 75, 88, 89
Gaussian profile advection, 118
potential, 154
interfacial internal waves, 129, 136 solenoidal, 43, 44, 46, 47, 75, 78, 88
Laplace equation, 114 vertical component, 25, 72
lock exchange flow, 129, 141
Molenkamp, 119 wave
Poisson equation, 115 breaking, 10, 177
rotating cone, 119 dispersion relation, 211, 212
slope current, 108 dispersion relationship, 136
solitary wave, 130, 176 internal
standing surface wave, 129, 153 celerity, 137
waves over underwater channel, 130, dispersive, 137
171 interfacial, 136
waves reflection, 129, 130 long, 137
wind-driven circulation, 129, 134 non-dispersive, 137
theorem short, 137
Fourier, 115 standing, 11
Leibniz, 55 lee wave, 21
time orbital movement, 21, 212
discretisation reflection, 130, 175
Crank-Nicolson, 25 sinusoidal, 211
explicit formulation, 86 small amplitude, 209
implicit, 163 solitary, 176
implicit formulation, 86 soliton, 176
implicitness factor, 86, 98, 101 surface
semi-implicit, 23, 25, 85, 90, 163 celerity, 143, 164, 211
Tisat, 5 deep water, 3, 212
tracer dispersive, 209, 212
active, 14, 214 long, 130, 142, 172, 209, 212
passive, 14 non-dispersive, 209, 212
salinity, 139, 146, 213 shallow water, 212
Index 243

short, 3, 21, 130, 154, 172, 201, 209,


212
standing, 154
standing, long, 155
standing, short, 154
velocity, 154, 212
tsunami, 10
viscous damping, 175
244 Danksagung

Danksagung

Obwohl ich die meisten Zeit als selbständiger Einzelkämpfer mit den Problemen, die mir
diese Arbeit bereitete, beschäftigt war, gibt es eine Reihe von Personen, bei denen ich
mich bedanken möchte. Ich hoffe ich vergesse niemandem.
Zuerst Herr Prof. Werner Zielke: Im Oktober 1991 hätte er mich, zu dieser Zeit einen
perfekten Outsider, nach einem Vorstellungsgespräch wieder vor der Tür setzen können,
was niemand bemerkt hätte. Er hat das aber nicht gemacht, und ist später, sehr konse-
quent Chef und Lehrer geblieben, mit Gewährleistung der Möglichkeit zur Promotion an
seinem Institut inbegriffen und schließlich der Übernahme des Hauptreferats. Prof. Ziel-
ke hat eine Gabe, den Mitarbeitern eine gewisse Prise Freiheit zu geben, auch wenn die
Resultate, wie immer in der Forschung, ohne Erfolgsgarantie sind. Ohne dieser Toleranz
und Freiheit ist keine kreative Tätigkeit möglich, und diese Arbeit wäre wahrscheinlich
nicht entstanden. Danke!
Unbedingt möchte ich Prof. Rainer Helmig (den ich duzen darf!), für die Übernahme des
Korreferates besonders herzlich danken. Er hat es auf besonders natürlich selbstverständ-
liche Weise gemacht und mich dabei ermuntert, ein bißchen selbstsicherer zu werden. Er
hat zu meiner Dissertation so tiefgreifende, lehrreiche und allzu treffliche Anmerkungen
formuliert, daß ich meine Arbeit in diesem strömungsmechanisch–numerischen Dschun-
gel richtig zuordnen und die einfachsten Fragen (Was gibt es bereits? Was mache ich
neu? Wofür ist das?) wirklich ganz klar beantworten konnte.
Den ehrenwürdigen Rest der zahlreicher Mitglieder meiner Prüfungskomission, nämlich
(in alphabetischer Reihenfolge) den Herren Vorsitzenden Prof. Rudolf Damrath,
Dr.-Ing. Olaf Kolditz, Dr.-Ing. Ralf Mahnken, Prof. Mark Markofsky, und Prof. Gu-
stav Rosemeier danke ich für ihr Engagement und eine ehrliche, gründliche und für mich
überraschend sehr angenehme Prüfung.
Andreas Malcherek hat im April oder Mai 1995 mich darauf aufmerksam gemacht, daß
auf der Basis der Telemac-Bibliotheken ein Druck(korrektur)verfahren (in so unpräziser
Weise haben wir das zu dieser Zeit genannt) zu basteln sei, und dabei auf seine sehr
direkte Weise gedroht, daß wenn ich das nicht mache, er das selber tun wird, obwohl er
das Programmieren für eine Sklavenarbeit hält. Für diesen kräftigen Ansporn muß ich
dankbar sein. Andreas hat auch die ursprünglichste Version dieses Textes (von 1996!)
gelesen und seine Anmerkungen habe ich in vielen Fällen dankbar berücksichtigt.
Reinhard ‘Phillip’ Hinkelmann danke ich für seinen unglaublichen Einsatz beim Prüfle-
sen und Formulieren der Vorschläge für Verbesserungen in meiner schriftlichen Arbeit.
Phillip gehört zu den wenigen Menschen, die meine Dissertation wirklich sehr genau und
sorgfältig gelesen haben. Er hat den gesamten Text auf solche konstruktiv kritische Art
gelesen, die sich jeder Doktorand in den heimlichsten Träumen wünscht. Phillip – Danke!
Meine Zimmergenossin (und früher auch HiWi!) Rebekka Kopmann hat mich immer auf
ihre fröhlich ungezwungene Weise in meiner täglicher Arbeit bedinungslos und sofort
unterstützt, insbesondere wenn eine gute Portion weiblicher Intuition notwendig war,
um die für männlichen Wesen schwierigen Probleme zu lösen. Danke!
Danksagung 245

Meinen ex-HiWis, Steffi Rüppel, und den beiden gerade Dipl.-Ing. gewordenen Hella
Schwartkop und Martin Kohlmeier danke ich für ihre Unterstützung. Ich habe von Euch
auch viel gelernt!
Ich bedanke mich auch bei meinen Kolleginnen und Kollegen am Institut, die direkt
oder indirekt am Erfolg dieser Arbeit beteiligt waren, und eine ausgesprochen exzellente
Arbeitsatmosphäre.
I would like to thank very much Michael Willey, whom I met by sheer chance in the inter-
net and never in the so-called real world. Mike appreciated my humble advice concerning
his planned bike trip to the Carpathians so much, that he felt obliged to proofread my
thesis from the purely lingual point of view as a native speaker. Everything was ar-
ranged via internet... Finally, I received my LATEX files back from Helsinki transformed
to HTML with all my mistakes corrected in red... Just in one week... What a quality of
proofreader’s workmanship! – Mike, thanks a lot!
Moim rodzicom i siostrze chcialbym podziekować
֒ za niezawodne podtrzymywanie mnie
na duchu, oraz za ciag
֒ la,
֒ szczególnie w trakcie rozmów telefonicznych objawiana֒ troske,
֒
kiedy w końcu wreszcie zabiore֒ sie֒ poważnie za doktorat... Dzialalo bardzo motywujaco!
֒
Dziekuj
֒ e!
֒
Die beiden Frauen, die im Institut als ‘meine Süßen’ oder einzeln als ‘meine Liebste’ und
‘mein Lieblingstöchterchen’ bekannt sind, sollen nicht vergessen werden! Es wird immer
wieder bei den Promotionen den am nähsten stehenden Familienmitglieder für Geduld,
das Vergeben der sündhaften Vernachlässigung des familiären Lebens, und so weiter,
gedankt. Ich bedanke mich auch dafür, aber im besonderen Maße danke ich dafür, daß
meine beiden Süßen in dieser manchmal sehr nervösen Zeit auch so viel gestört haben,
so daß ich ein Gleichgewicht zwischen der Arbeit und dem normalen Leben behalten
konnte.
Man sollte die materiellen Seiten nie vergessen. Ich bedanke mich bei allen Steuerzahlern
in der Budesrepublik Deutschland, die das Budget vom Bundesministerium für Bildung
und Forschung realisiert haben, und den Menschen, die über dies verfügen. Fast die
ganze Zeit konnte ich aus diesem Budget meinen (und meiner Familie) Lebensunterhalt
sichern, mal als Adressat einer Zuwendung, und mal als Bearbeiter eines Tiefseeumwelt-
schutz–Projektes. Danke!
Bei der bedrohten Spezies Yeti (und anderen eindeutig mehr real existierenden und mit
wirklichen Problemen heimgesuchten Himalaja-Völker) möchte ich mich für die heimliche
Unterstützung bedanken.
Am Ende möchte ich allen danken, bei denen ich vergessen habe, mich zu bedanken.

Verpflichtet,

Jacek A. Jankowski

Hannover, Dezember 1998


Liste der bisher erschienenen Institutsberichte
(* = bereits vergriffen)

01/1970 * Holz, K.-P. Ergänzung des Verfahrens finiter Elemente durch


Ecksingularitäten zur verbesserten Berechnung schiefwinkliger Plat-
ten. Dissertation, Techn. Univ. Hannover, 1970
02/1971 Ehlers, K.-D. Berechnung instationärer Grund- und Sickerwas-
serströmungen mit freier Oberfläche nach der Methode finiter Ele-
mente. Dissertation, Techn. Univ. Hannover, 1971
03/1971 * Meissner, U. Berechnung von Schalen unter großen Verschie-
bungen und Verdrehungen bei kleinen Verzerrungen mit Hilfe finiter
Dreieckselemente. Dissertation, Techn. Univ. Hannover, 1971
04/1972 Grotkop, G. Die Berechnung von Flachwasserwellen nach der
Methode der finiten Elemente. Dissertation, Techn. Univ. Hanno-
ver, Sonderdruck aus dem Jahresbericht 1971 d. SFB 79, H. 2, 1972
05/1973 Schulze, K.-W. Eine problemorientierte Sprache für die Dyna-
mik offener Gerinne. Dissertation, Techn. Univ. Hannover, Mitteil.
d. SFB 79, Heft 1, 1973
06/1977 Beyer, A. Die Berechnung großräumiger Grundwasserströmun-
gen mit Vertikalstruktur mit Hilfe der Finite–Element–Methode.
Dissertation, Fortschrittberichte der VDI–Zeitschriften, Reihe 4,
Nr. 34, 1977
07/1977 * Ebeling, H. Berechnung der Vertikalstruktur wind- und gezeiten-
erzeugter Strömungen nach der Methode der finiten Elemente. Dis-
sertation, Fortschrittberichte der VDI–Zeitschriften, Reihe 4, Nr.
32, 1977
08/1977 * Gärtner, S. Zur Berechnung von Flachwasserwellen und insta-
tionren Transportprozessen mit der Methode der finiten Elemen-
te. Dissertation, Fortschrittberichte der VDI–Zeitschriften, Reihe
4, Nr. 30, 1977
09/1977 * Herrling, B. Eine hybride Formulierung in Wasserständen zur
Berechnung von Flachwasserwellen mit der Methode finiter Elemen-
te. Dissertation, Fortschrittberichte der VDI–Zeitschriften, Reihe 4,
Nr. 37, 1977
10/1979 Hennlich, H.-H. Aeroelastische Stabilitätsuntersuchung von
Linientragwerken. Dissertation, Fortschrittberichte der VDI–Zeit-
schriften, Reihe 4, Nr. 49, 1979
11/1979 Kaločay, E. Zur numerischen Behandlung der Konvektions–
Diffusions–Gleichung im Hinblick auf das inverse Problem. Disser-
tation, Univ. Hannover, 1979
12/1980 Januszewski, U. Automatische Eichung für ein- und zweidimen-
sionale, hydrodynamisch-numerische Flachwassermodelle. Disserta-
tion, Univ. Hannover, Fortschrittberichte der VDI–Zeitschriften,
Reihe 4, Nr. 58, 1980
13/1982 * Carbonel Huamán, C.A.A. Numerisches Modell der Zirkula-
tion in Auftriebsgebieten mit Anwendung auf die nordperuanische
Küste. Dissertation, Univ. Hannover, 1982
14/1985 Tuchs, M. Messungen und Modellierung am Deep Shaft. Disser-
tation, Univ. Hannover, 1984
15/1985 Theunert, F. Zum lokalen Windstau in Ästuarien bei Sturm-
fluten – Numerische Untersuchungen am Beispiel der Unterelbe.
Dissertation, Univ. Hannover, 1984
16/1985 Perko, H.-D. Gasausscheidung in instationärer Rohrströmung.
Dissertation, Univ. Hannover, 1984
17/1985 Crotogino, A. Ein Beitrag zur numerischen Modellierung
des Sedimenttransports in Verbindung mit vertikal integrierten
Strömungsmodellen. Dissertation, Univ. Hannover, 1984
18/1985 Rottmann–Söde, W. Ein halbanalytisches FE–Modell für har-
monische Wellen zur Berechnung von Wellenunruhen in Häfen und
im Küstenvorfeld. Dissertation, Univ. Hannover, 1985
19/1985 Nitsche, G. Explizite Finite–Element–Modelle und ihre Natur-
anwendungen auf Strömungsprobleme in Tidegebieten. Dissertati-
on, Univ. Hannover, 1985
20/1985 Vera Muthre, C. Untersuchungen zur Salzausbreitung in Ästu-
arien mit Taylor’schen Dispersionsmodellen. Dissertation, Univ.
Hannover, 1985
21/1985 Schaper, H. Ein Beitrag zur numerischen Berechnung von nicht-
linearen kurzen Flachwasserwellen mit verbesserten Differenzenver-
fahren. Dissertation, Univ. Hannover, 1985
22/1986 Urban, C. Ein Finite–Element–Verfahren mit linearen Ansätzen
für stationäre zweidimensionale Strömungen. Dissertation, Univ.
Hannover, 1986
23/1987 Heyer, H. Die Beeinflussung der Tidedynamik in Ästuarien durch
Steuerung – Ein Beitrag zur Anwendung von Optimierungsverfah-
ren in der Wasserwirtschaft. Dissertation, Univ. Hannover, 1987
24/1987 Gärtner, S. Zur diskreten Approximation kontinuumsmechani-
scher Bilanzgleichungen. Institutsbericht, davon 4 Abschnitte als
Habilitationsschrift angenommen, Univ. Hannover, 1987
25/1988 Rogalla, B.U. Zur statischen und dynamischen Berechnung geo-
metrisch nichtlinearer Linientragwerke unter Strömungs– und Wel-
lenlasten. Dissertation, Univ. Hannover, 1988
26/1990 * Lang, G. Zur Schwebstoffdynamik von Trübungszonen in Ästua-
rien. Dissertation, Univ. Hannover, 1990
27/1990 Stittgen, M. Zur Fluid–Struktur–Wechselwirkung in flexiblen
Offshore–Schlauchleitungen. Dissertation, Univ. Hannover, 1990
28/1990 Wollrath, J. Ein Strömungs- und Transportmodell für klüftiges
Gestein und Untersuchungen zu homogenen Ersatzsystemen. Dis-
sertation, Univ. Hannover, 1990
29/1991 Kröhn, K.-P. Simulation von Transportvorgängen im klüftigen
Gestein mit der Methode der Finiten Elemente. Dissertation, Univ.
Hannover, 1991
30/1991 Lehfeldt, R. Ein algebraisches Turbulenzmodell für Ästuare. Dis-
sertation, Univ. Hannover, 1991
31/1991 Prüser, H.-H. Zur mathematischen Modellierung der Interaktion
von Seegang und Strömung im flachen Wasser. Dissertation, Univ.
Hannover, 1991
dito Schröter, A. Das numerische Seegangsmodell BOWAM2 1990
– Grundlagen und Verifikationen – . Univ. Hannover, 1991
32/1992 Leister, K. Anwendung numerischer Flachwassermodelle zur Be-
stimmung von Wasserlinien. Dissertation, Univ. Hannover, 1992
33/1993 Ramthun, B. Zur Druckstoßsicherung von Fernwärmenetzen und
zur Dynamik von Abnehmeranlagen. Dissertation, Univ. Hannover,
1993
34/1993 Helmig, R. Theorie und Numerik der Mehrphasenströmungen in
geklüftet–porösen Medien. Dissertation, Univ. Hannover, 1993
35/1994 Plüß, A. Netzbearbeitung und Verfahrensverbesserungen für Ti-
demodelle nach der Finiten Element Methode. Dissertation, Univ.
Hannover, 1994
36/1994 Nöthel, H. Statistisch–numerische Beschreibung des Wellen-
und Strömungsgeschehens in einem Buhnenfeld. Dissertation, Univ.
Hannover, 1994
37/1994 Shao, H. Simulation von Strömungs- und Transportvorgängen in
geklüfteten porösen Medien mit gekoppelten Finite–Element– und
Rand–Element–Methoden. Dissertation, Univ. Hannover, 1994
38/1994 Stengel, T. Änderungen der Tidedynamik in der Deutschen
Bucht und Auswirkungen eines Meeresspiegelanstiegs. Dissertation,
Univ. Hannover, 1994
39/1994 Schubert, R. Ein Softwaresystem zur parallelen interaktiven
Strömungssimulation und -visualisierung. Dissertation, Univ. Han-
nover, 1994
40/1994 Alm, W. Zur Gestaltung eines Informationssystems im Küsten-
ingenieurwesen. Dissertation, Univ. Hannover, 1994
41/1994 Benali, H. Zur Kopplung von FEM– und CAD–Programmen
im Bauwesen über neutrale Datenschnittstellen. Dissertation, Univ.
Hannover, 1994
42/1995 Schröter, A. Nichtlineare zeitdiskrete Seegangssimulation im fla-
chen und tieferen Wasser. Dissertation, Univ. Hannover, 1995
43/1995 Blase, Th. Ein systemtechnischer Ansatz zur Modellierung
von Hydraulik, Stofftransport und reaktionskinetischen Prozessen
in Kläranlagen. Dissertation, Univ. Hannover, 1995
44/1995 Malcherek, A. Mathematische Modellierung von Strömungen
und Stofftransportprozessen in Ästuaren. Dissertation, Univ. Han-
nover, 1995
45/1995 Lege, T. Modellierung des Kluftgesteins als geologische Barriere
für Deponien. Dissertation, Univ. Hannover, 1995
46/1996 Arnold, H. Simulation dammbruchinduzierter Flutwellen. Dis-
sertation, Univ. Hannover, 1996
47/1996 Kolditz, O. Stoff- und Wärmetransport im Kluftgestein. Habili-
tation, Univ. Hannover, 1996
48/1996 Hunze, M. Numerische Modellierung reaktiver Strömungen in
oberflächenbelüfteten Belebungsbecken. Dissertation, Univ. Hanno-
ver, 1996
49/1996 Wollschläger, A. Ein Random-Walk-Modell für Schwermetall-
partikel in natürlichen Gewässern. Dissertation, Univ. Hannover,
1996
50/1997 Feist, M. Entwurf eines Modellierungssystems zur Simulation
von Oberflächengewässern. Dissertation, Univ. Hannover, 1997
51/1997 Hinkelmann, R. Parallelisierung eines Lagrange–Euler–Ver-
fahrens für Strömungs- und Stofftransportprozesse in Oberflächen-
gewässern. Dissertation, Univ. Hannover, 1997
52/1997 Barlag, C. Adaptive Methoden zur Modellierung von Strofftrans-
port im Kluftgestein. Dissertation, Univ. Hannover, 1997
53/1997 Saberi-Haghighi, K. Zur Ermittlung der verformungsabhängi-
gen Windbelastung bei Hängedächern. Dissertation, Univ. Hanno-
ver, 1997
54/1998 Krüger, A. Physikalische Prozesse im Nachklärbecken – Modell-
bildung und Simulation. Dissertation, Univ. Hannover, 1998
55/1998 Wolters, A. H. Zur Modellierung des instationären thermohy-
draulischen Betriebsverhaltens von Fernwärmeanlagen. Dissertati-
on, Univ. Hannover, 1998
55/1998 Wolters, A. H. Zur Modellierung des instationären thermohy-
draulischen Betriebsverhaltens von Fernwärmeanlagen. Dissertati-
on, Univ. Hannover, 1998
56/1999 Jankowski, J. A. A non-hydrostatic model for free surface flows.
Dissertation, Univ. Hannover, 1999

View publication stats

Potrebbero piacerti anche