Download as pdf or txt
Download as pdf or txt
You are on page 1of 138

Numerical Simulation of Free-Surface Flow

with Moving Rigid Bodies

Geert Fekken
Rijksuniversiteit Groningen

Numerical Simulation of Free-Surface Flow


with Moving Rigid Bodies

Proefschrift

ter verkrijging van het doctoraat in de


Wiskunde en Natuurwetenschappen
aan de Rijksuniversiteit Groningen
op gezag van de
Rector Magnificus, dr. F. Zwarts,
in bet openbaar te verdedigen op
vrijdag 5 maart 2004
orn 16.15 uur

door

Geert Fekken

geboren op 25 mei 1975


te Scharsterbrug
Promotor: Prof. dr. A.E.P. Veidman
Beoordelingscommissie: Prof. dr. ir. A.J. Hermans
Prof. dr. ir. J.A. Pinkster
Prof. dr. ir. G.S. Stelling
The research presented in this thesis was funded by the Maritime Research Institute
Netherlands (MARIN) Wageningen.
Contents

i Introduction i
1.1 Motivation 1
1.2 Moving boundary methods 2
1.2.1 Lagrangian methods 3
1.2.2 Eulerian methods 4
1.2.3 Combined Lagrangian-Eulerian methods 5
1.3 Outline 7

2 Computational method 9
2.1 Equations of flow including moving rigid bodies 9
2.1.1 Equations of fluid dynamics 9
2.1.2 Free surface 10
2.1.3 Extension to moving bodies 11
2.2 Spatial discretisation 11
2.2.1 Mesh generation 11
2.2.2 Discretisation of the geometry 12
2.2.3 Discretisation of the continuity equation 20
2.2.4 Discretisation of the momentum equations 22
2.2.5 Discretisation of the free surface 35
2.3 Temporal discretisation 40
2.3.1 Poisson equation for the pressure 41
2.3.2 Stability 41
2.4 Behaviour of the pressure 47
2.4.1 Velocity boundary conditions for the free surface 48
2.4.2 Pressure boundary conditions at the free surface 59
2.4.3 Moving bodies with sharp corners 61
2.4.4 Pressure positioning 65
2.4.5 Other possible reasons for spikes 70
2.5 Coupled motion between moving bodies and fluid 71
2.5.1 Basic idea 71
2.5.2 Numerical stability 71

3 Results 75
3.1 Validation tests without free surface 75
3.1.1 Free stream consistency tests 75
3.1.2 Added mass coefficients 76
v2zi Contents

3.2 Oscillation of a floating body 80


3.3 Motion of a cylinder moving through a free surface 83
3.3.1 Cylinder rising to a free surface 83
3.3.2 Cylinder sinking from a free surface 86
3.3.3 Cylinder sinking through a free surface 89
3.3.4 High speed water entry of a cylinder 91
3.3.5 Concluding remarks 93
3.4 The wedge entry phenomenon 93
3.4.1 Model tests Marin 93
3.4.2 Impact forces on sections of a planing craft 94
3.4.3 Water entry of different-shaped two-dimensional sections 97
3.4.4 Impact forces on a falling life-boat 100
3.5 Equilibrium state of various floating objects 102
3.5.1 Floating circular cylinder 103
3.5.2 Floating rectangular cylinder 104
3.6 Sideways launching of a ship 111

4 Summary and conclusions 115

Bibliography 119

Samenvatting 129

Dankwoord 133
Chapter 1
Introduction

1.1 Motivation
The subject of this thesis finds its main ground in the motion of ships and offshore
structures at sea. In particular, the phenomenon of 'green water loading' on the deck of
a ship, studied extensively by Buchner [6], was an important motivation. When a ship is
in a heavy storm, waves become very large. Sometimes, they even exceed the deck level,
resulting in a flow of water onto the deck. This problem is called 'green water loading',
where the 'green water' indicates the solid green seawater on the deck, rather than the
spray flying around (figure 1.1). During this green water loading, serious damage can be
done to deck structures on the ship, e.g. see [18, 62]. In the North Sea, sometimes wave
heights higher than 30 meter occur [19]. Also offshore platforms may get damaged by
wave induced loads.

Figure 1.1: Picture of wave loading on a ship and the resulting spray. This picture is
taken from http://www. btintern et. com/ derek. mackay/offshore/images/vessels/.

To avoid these incidents in the future, it is important to increase the knowledge of the
flow behaviour and the wave induced motions of ships and offshore structures. Especially,
the forces and pressures that are exerted by the impact loads on the bow and the deck
Chapter 1. Introduction

structures are of importance. Experimental study is one of the means to achieve more
insight in the flow behaviour and estimate impact forces. Another useful tool in the
prediction of green water is numerical simulation. This area is developing and significant
progress is still made. A historical overview of green-water research can be found in [6].
In this study, the interest will be focused on the design of a numerical simulation tool
for the dynamics of moving bodies in free-surface flows in general, and for green-water
loading and other impact phenomena in particular. A pilot study of green-water loading
without the influence of the ship motion has shown promising results [21]. Recently,
participation in the SAFE-FLOW project about green-water loading has resulted in [42,
43] and is still going. The computational method that was used in [21] has also shown its
capabilities in simulating other applications of free-surface flows, such as anti-roll tank
simulations [16, 17] and free-surface flow in micro-gravity environments [23-25]. The
method is also able to simulate flows without a free surface, even in interaction with
elastic boundaries with as main application blood flow through arteries. Details caii be
found in Loots' PhD. thesis [54] and [55, 56].
As a continuation of the green-water study, in order to simulate green water more
realistically, the motion of the ship has to be simulated as well. Therefore, a method
has to be developed that is able to handle freely moving bodies. With this ability, more
realistic situations can be simulated where ships and offshore structures are freely floating
in high waves. Moreover, besides marine applications lots of other industrial applications
can be imagined for simulating moving bodies in fluids. More precise, the main objective
of this thesis is to develop a method that satisfies the following properties:
The ability of simulating fluid flow with arbitrary moving rigid bodies both with
prescribed motion and moving interactively with the fluid.
The ability of simulating fluid flow with large deforming free surfaces.
Realistic simulation of wave impact phenomena like green-water loading, including
computation of hydrodynarnic forces and pressures.
Since the method in [21], which is implemented in a computer program called ComFlo,
has already shown to be capable of handling a large deforming free surface and has
shown promising results for green-water loading, this method is chosen to be improved
and extended with the above mentioned properties in this thesis. A summary of moving
boundary methods is given in the next section.

1.2 Moving boundary methods


The treatment of moving boundaries in a computational flow model is not an easy task.
In free-surface flow models including moving rigid bodies, two types of moving boundaries
are present. The separation between two fluid phases, e.g. between water and air, whose
location is not known a priori and is able to move freely, and the boundary of the rigid
body which follows the rigid body motion. According to the physical situation, the
moving boundaries should be treated as sharp interfaces between the different phases
without smearing the information in the neighbourhood of the interfaces. The advection
of a free surface is more difficult than the advection of a rigid body boundary, since a free
1.2 Moving boundary methods 3

surface may undergo large deformations and even topological changes, while in principle
the location of the boundary of a rigid body is completely determined by its gravitational
center and a rotation. For tracking moving interfaces, several techniques exist, each with
its own strengths and weaknesses [81]. They can be classified in three categories [82]:

Lagrangian methods
Eulerian methods
Combined Lagrangian-Eulerian methods

1.2.1 Lagrangian methods


In Lagrangian methods, also called moving grid methods, the boundary is treated ex-
plicitely without smearing the information at the interface. This is done by fitting the
computational grid with the boundaries, by means of a boundary conforming curvilinear
grid [96], a block-structured domain decomposition [31,80] or overset meshes [34, 84].
The governing equations are solved on this boundary fitted grid. The grid is constantly
regenerated to conform to the varying physical domain. Boundary conditions can be
applied at the exact location of the free surface or rigid boundary. When the geometric
variation is quite modest, this can be a very efficient technique. This is demonstrated for
a soft contact lens problem in Shyy et al. [82]. Liu and Kawachi [52,53] have applied the
moving grid technique successfully to compute flapping wing of insect flight and undula-
tory swimming. However, with a strongly deforming free surface and arbitrarily moving
bodies, it will be very difficult and time consuming to conform the grid to the different
boundaries. Also difficulties in the form of grid skewness may arise and additional nu-
merical dissipation may be a consequence of the redistribution of the field variables in
the vicinity of the interface. When very complex moving bodies are used, it will be even
more complicated. In green-water flow on complex offshore structures, where the free
surface is subjected to large deformations (breaking waves etc.) and possible topological
changes, it may be clear that these moving grid techniques are not applicable.
A special case of the Lagrangian method is the boundary element method (BEM),
or boundary integral method. Here the problem is reduced to a discretisation only over
the boundary of the domain. The flow is assumed to be incompressible, inviscid and
to be governed by the Laplace equation (potential flow). This method has been used
extensively in interface tracking. Buchner and Cozijn [7] investigated the possibilities for
simulation of green-water loading of the two-dimensional non-linear boundary integral
method developed by Romate [74] and Van Daalen [14]. The main restriction of bound-
ary element methods is the fact that the computations break down when the interface
undergoes topological changes, making green-water simulation a difficult task.
A method becoming increasingly popular that can treat large deforming interfaces and
topological changes is the Smoothed Particle Hydrodynamics (SPH) method originally
developed in [26,57]. This is a meshless method, based on a representation of the fluid
body by large particles of fluid that are subject to Newton's Second Law [73]. Spatial
derivatives are estimated by analytical expressions and do not need a fixed computational
grid. This method has been extended to complicated free-surface flows, e.g. solitary wave
propagation over a planar beach [61], plunging breakers [90] and dam break simulations
Chapter 1. Introduction

[60]. A disadvantage of the method is, that it inherits difficulties of the correct modeling
of boundaries [73].

1.2.2 Eulerian methods


Eulerian methods, or fixed grid methods, are particularly suited for surfaces that undergo
large deformations such as a free surface. The interface is not explicitely tracked but is
reconstructed from the field variables on a fixed grid. The location of the interface
can be determined based on volume fraction information, unavoidably resulting in an
uncertainty of about one grid cell. It is difficult to apply boundary conditions at the
exact location of the boundary, but discretisation is easy, since no remeshing procedures
have to be performed. Topological changes are easier to handle, but some accuracy can
be lost when details of the interface can not be covered by the grid. Computations can
be performed over different types of fixed meshes. The most popular Eulerian methods
are the volume-of-fluid method and the level-set method which are described briefly now.

Volume-of-Fluid (VOF) methods


A very popular method for advecting an interface is the volume-of-fluid method originally
developed by Hirt and Nichols [32]. This method makes use of a fluid fraction variable
F, with values between zero and one, indicating the fraction of each computational cell
that is filled with a fluid phase. This VOF function F is then advected according to the
local flow velocity. More accurate schemes for advecting the volume fractions have been
developed after that [2,4,49,51, 76, 110]. A comprehensive review is given by Rider and
Kothe [72]. In general, two classes of methods are used, namely piecewise constant and
piecewise linear reconstructions of the interface, dependent on the required accuracy of the
reconstruction. Recently, some more strategies to sharpen the interface in VOF methods
have been developed. Ubbink and Issa [92] proposed a high-order discretisation scheme
with bounding treatments. In Gerrits [23], a piecewise constant interface reconstruction
combined with a local level-set function is described that resulted in improved interface
definitions. Gerrits concluded that, for a velocity field that is solved from the governing
Navier-Stokes equations, this method gives the best results compared to piecewise linear
reconstructions, even in combination with a local level-set function.
The strength of the VOF methods is that they can handle large distortions of the
interface, and do not have serious problems with respect to mass conservation. Among
others, they have been applied to complex interfacial phenomena, including droplet dy-
namics and breakup, morphological instabilities in crystals [83], spray dynamics [51],
micro-gravity environments including the effects of surface tension and contact angles
[23] and in the free-surface definition of the before mentioned pilot study of green water
loading [21].

Level-set methods
In the original level-set Hamilton-Jacobi formulation [65, 78, 79] a distance function (z, t)
is introduced denoting the distance from x to the initial interface location at t = 0. The
interface corresponds to the contour = O at any instant, while a particle on the level
1.2 Moving boundary methods 5

set evolves in time by


+uVI=0. (1.1)
Here u, is the normal velocity of the interface. In this way, highly distorted interfaces
can be treated, also topology changes are incorporated automatically. For incompressible
two-phase flow, Sussman et al. [87] have presented results based on the level-set approach.
However, the interface is still of a finite thickness. The method has serious problems with
respect to mass conservation [23]. Several strategies have been studied to overcome these
problems, e.g. a combination with a VOF method has been used [86] or a re-distancing
algorithm [85].

Other Eulerian methods


Other Eulerian methods for moving interfaces than VOF and level-set methods also
exist of course. The phase field method [44, 103], based on expressing the free energy
of the system as a Cahn-Hilliard functional [9], is especially designed for generating
realistic solidification microstructures. The enthalpy in solidification problems developed
by Voller and Prakash [80, 100]. Xiao [105] developed a three-dimensional method for
rigid bodies in viscous flows, where the interface definition is sharpened by employing a
tangent function transformation along with Colella and Woodwards' piecewise parabolic
method (PPM) [13].

1.2.3 Combined Lagrangian-Eulerian methods


The Combined Lagrangian-Eulerian methods combine some of the properties of the La-
grangian and the Eulerian methods. The computations are done on a fixed grid and
the interface is tracked explicitely as a curve (in two dimensions) or a surface (in three
dimensions). In this way, the interface definition stays sharp. Difficulties may arise when
interfaces are deforming and intersecting, especially in three dimensions. These methods
are therefore especially suitable for applications with moving rigid bodies, because their
interfaces do not deform. An overview of the most popular existing Lagrangian-Eulerian
methods is given with their strengths and weaknesses.

Immersed boundary method


The immersed boundary method was originally used by Peskin [71] to simulate blood flow
in the heart. It was also used for modeling swimming organism [20]. Later, the method
has been extended and used by others [36-39,93,95].
In the method, the domain is covered by a fixed Cartesian grid. The interface can be
represented by a curve (2D) or surface (3D) that is defined by a polynomial fitting to a
set of marker particles. The material properties are assigned based on a Heaviside step
function, which allows a smooth distribution of the material properties over a transition
zone d, which is usually taken about twice the grid spacing h (d = 2h):
¡3 = ßi + (ß2 - ßi)H(z - Xk). (1.2)

Here fi could be any material property, with flu in material i and ¡32 in material 2.
x denotes the grid coordinate and Xk denotes the interfacial marker coordinate. The
6 Chapter 1. Introduction

interface can thus be considered to have finite thickness d. The velocity field is stored
at the cell centers, and is interpolated to obtain the velocity at the interfacial points.
The interface force acting on the marker points is spread to the nearby grid points by a
discrete Delta function, which is the derivative of the Heaviside step function. This delta
function typically spreads over 4h. The interface force is incorporated in the momentum
equation by means of an integral source term. The advection of the interface is done by
displacing the marker points following x1 = x + ¿Xtu where u is the velocity of the
interfacial marker k at time n.
The method of Unverdi and Tryggvason [95] represents the boundary as a two-
dimensional triangular grid (in 3D). The method is suitable for viscous multi-fluid flows,
thus the interface may deform. Interaction between interfaces is automatically accounted
for. The thickness of the interface is of the order of the mesh size. The heaviside step
function that is used to distribute the material properties smoothly is derived from an
indicator function I, that is i or O dependent on which side of the interface I is applied.
The method has manly been used for calculating bubbles including the effect of surface
tension.

Fictitious domain method


The more recently developed fictitious domain method by Glowinski [27,28], was es-
pecially developed for incompressible flows with moving rigid bodies. The method is
particularly suited to the simulation of particulate flow, such as the flow of mixtures of
rigid solid particles and incompressible viscous fluids.
The method of Glowinski is based on computing the flow on a fixed space region con-
taining the moving bodies. The Navier-Stokes equations are approximated by a finite-
element method. The idea is that the complete space region is filled with an incom-
pressible fluid, thus the moving rigid bodies are filled with the surrounding fluid. The
fluid inside the body is assumed to have a rigid body motion. The rigid body motion
inside each moving body is forced via a Lagrange multiplier distributed over the body
and a variational formulation involving Lagrange multipliers is derived. These Lagrange
multipliers can be seen as a gluon to match the velocity fields of two continua. Other
formulations of the fictitious domain method are also used. Khadra et al. [41], presented
a formulation based on the permeability of different media. Patankar et al. [66] used an
approach where the deformation rate tensor is constrained to be zero at points in the
fluid occupied by rigid solids.

Cut cell method


In both the immersed boundary method and the fictitious domain method, the influence
of the interface is acquired by body forces. The boundary is treated as a special region in
a single fluid, covering the entire computational domain. The interface does not clearly
demarcate two different regions and discontinuities across the interface are somewhat
smoothed. These methods belong to the category of so-called diffuse interface methods.
For convection-dominated flow, like the green-water phenomenon, the interface between
the rigid body and the fluid should be treated as a sharp interface and the discretisa-
tion of the flow equations should also treat it as such. True sharp interface methods are
rare. Almgren [1] used a variational formulation, Leveque et al. [50] have introduced an
1.9 Outline 7

immersed interface method for tracking interfaces explicitely, maintaining sharp discon-
tinuities, for problems involving elliptic PDEs. Udaykumar et al.[94] introduced a cut
cell method. Also a cut cell method for compressible flow was used by Yang, Causon and
Ingram [106]. Some basics of the cut cell method by Udaykumar [93] are given now.
The interface is represented by marker particles as in the immersed boundary method
and is tracked in the same way. The interface is assumed to be built of piecewise linear
segments. The underlying Cartesian grid with a cell centered storage of the flow vari-
ables, is then cut by the interface and the intersections with the grid are determined.
Cells that are cut by the interface are reshaped by discarding the part that lies in the
solid. To avoid small cells, a cell merging with neighbouring cells takes place when cell
centers of cut cells lie in the solid. Trapezoidal shaped control volumes can be formed in
this way. The second-order accuracy of the flow solver will be reserved when evaluating
mass, convective and diffusive fluxes and pressure gradients on the cell faces with suffi-
cient accuracy. A complex two-dimensional polynomial interpolating function is used in
combination with a finite-volume formulation to achieve this accuracy and to guarantee
the conservation property of the underlying algorithm.

Verstappen and Veldman [97,99], developed a cut cell method, originally designed
for the direct numerical simulation of turbulent flow. The method differs slightly from
the cut-cell method by Udaykumar. A staggered storage of the flow variables is used,
guaranteeing exact mass conservation, no pressure decoupling problems and no complex
interpolation procedures. Further, the equations are discretised in such a way that there
is no need to avoid small cells, thus no merging with other cells is required.

In this thesis, a fixed Cartesian grid is used. The free surface is modeled with a volume-
of-fluid method in combination with a local level-set function [23] to maintain a sharp
interface, while allowing a large deforming free surface. For the moving rigid boundary,
the cut-cell method by Verstappen and Veldman [97, 99] is used and extended to moving
bodies. The method is based on the Navier-Stokes equations that are discretised by
a finite-volume method. The discretisation is done in such a way that the symmetry
properties of the underlying continuous differential operators are preserved. This appears
to be advantageous for the stability of the numerical method [97].

1.3 Outline
The contents of the remainder of this thesis is divided in three chapters, describing the
computational method, results and conclusions respectively.
Chapter 2 starts with the governing equations for flow including moving bodies and
a free surface with its boundary conditions in section 2.1.
In section 2.2 the discretisation in space is explained. The Cartesian grid in com-
bination with the discretisation of the complex geometry will result in cut cells defined
by apertures. The definition and tracking of the moving rigid bodies in two dimensions
and in three dimensions is explained. A labeling system is introduced to distinguish
boundary cells and flow cells. The discretisation of the continuity equation and the mo-
mentum equations using a finite volume method are explained. Another labeling system
8 Chapter 1. Introduction

is introduced for the discretisation of the free surface, distinguishing air and fluid. A
volume-of-fluid function is introduced to assign the filling ratio of the cells. Boundary
conditions for the pressure and velocity at the free surface are discussed and the volume-
of-fluid method in combination with a local height function is explained for displacing
the free surface.
Section 2.3 describes the temporal discretisation. The Poisson equation for the pres-
sure that has to be solved and some additional difficulties regarding small cells in combi-
nation with moving boundaries are treated.
In section 2.4 the difficulties regarding cut cells in combination with a free surface
are explained. When the free surface is not properly treated, a spiky pressure signal may
result. Also, when moving bodies with sharp corners are used, it is difficult to maintain a
smooth pressure signal. The different origin of the spikes are discussed and some solutions
are proposed to improve the results.
A first attempt to the treatment of the interactive coupling of the motion between
the fluid and the moving bodies is explained in section 2.5. The numerical stability of
the coupling is discussed and an implicit coupling algorithm is proposed.
In chapter 3, various results are given. First some validation tests are done, including
free stream consistency tests and added mass validation tests with different moving bodies
without a free surface (section 3.1).
In section 3.2 a comparison with a method of Yeung and Anantakrishnan [109] has
been made for the oscillation of a floating cylinder with prescribed motion. Different fre-
quencies and amplitudes of oscillation are used and hydrodynamical forces are compared
for both methods. These results are also compared with linear potential results.
Section 3.3 gives an extensive comparison for free-surface evolutions and velocities for
rising and sinking cylinders with a potential method of Moyo and Greenhow [63]. The
motion of the cylinders is not prescribed but follows from the forces exerted by the fluid.
In section 3.4 the wedge entry phenomenon is discussed. Various comparisons are
made with experimental data including experiments with falling life boats. Impact forces
and pressures are calculated and compared.
In section 3.5 the ability of the method to predict the correct equilibrium state under
interactive motion of different shaped moving bodies with different densities is validated
in two and in three dimensions.
Section 3.6 shows some numerical results of the sideways launching of a ship, carried
out by de Jong [35] with the present method.
In chapter 4, finally, this thesis ends with a summary and conclusions.
Chapter 2
Computational method
In this chapter the computational method for the simulation of the fluid flow is explained.
First, the mathematical equations are stated that describe the flow including the motion
of the moving bodies and the free surface. Second, the discretisation in space of the
geometry, free surface and the governing equations is discussed. Finally, the temporal
discretisation is explained.

2.1 Equations of flow including moving rigid bodies


This section describes the equations that are required for computing the flow. First
the general equations for an incompressible fluid are treated, then the equations and
boundary conditions for the free surface. Finally the extension is made to the case of
moving rigid bodies.

2.1.1 Equations of fluid dynamics


The mathematical equations, governing the motion of fluid flow in general can be writ-
ten in integral form, considering an arbitrary control volume V with boundary S. The
continuity equation, demanding conservation of mass, then becomes

fdv+f(u.n)ds=o,
V S
(2.1)

where u denotes the velocity vector (u, y, w), p the density of the fluid, t the time and n
the outward-pointing normal vector of S. In this thesis we only consider incompressible
flow, so the density of the fluid will be constant. The continuity equation (2.1) can then
be rewritten as
f(u n) dS = 0. (2.2)

Conservation of momentum, with respect to control volume V can be written in its


incompressible form, also known as the incompressible Navier-Stokes equations

fu dV + u (u n) dS = (f (Vu . n) - pn dS + J F dv). (2.3)


¡
10 Chapter 2. Computational method

Here p denotes the pressure, ¡. the molecular viscosity of the fluid and F = (Fi, F, F)
is an external force like gravity.
A more commonly used notation for equations (2.2) and (2.3) is the notation in
differential form
Vu=O
and

j+u.Vu= .-(uVp+F).
This formulation can be derived from the equations in integral form by replacing the
surface integrals into volume integrals by Gauss' integral theorem and stating that it holds
for any volume V. However, the integral form (2.2) and (2.3) is more physical, since it is
formulated as a conservation law. In a finite volume discretisation, this conservation law
is transformed into a discrete conservation law. It turns out that this discretisation has
nice numerical properties and yields advantages with respect to conservation of analytical
properties. Therefore, in this thesis, the integral form of the Navier-Stokes equations will
be the basis for the discretisation.

2.1.2 Free surface


The type of flow which is subject of this thesis, contains a separation of the fluid with the
air. The separation surface is called the free surface. This free surface can be described
by an equation S(x, t) = O. The position of the free surface will vary in time, and its
motion can be described by the equation

=+uVS=O. (2.4)

Further, boundary conditions are required to satisfy the balancing forces at the free
surface. These consist of a tangential stress condition

ôu aU
(2.5)

and a normal stress condition


aun
p+2p-- = p0+ak. (2.6)

Here n is the normal of the free surface, t is a direction tangential to the free surface, Po
is the atmospheric pressure, o denotes the surface tension of the fluid and i is the mean
curvature of the free surface which is computed by = V n = V (ja). In this study,
the flow will mostly be dominated by the gravitational force and surface tension effects
play a very minor role. In micro-gravity environments however, the surface tension will
be the driving force and this situation is extensively described in [23], including the effect
of contact angles.
2.2 Spatial discretisation 11

2.1.3 Extension to moving bodies


Allowing moving rigid bodies in the flow domain, one can look at the continuity equation
again. Consider an arbitrary control volume V including (part of) a moving body in
(figure 2.1). The shaded area Vb represents the part of the body, while V1 is the fluida!
part in the control volume. The boundaries Sb, S and Sb! are defined as Sb = 3Vb \ 0V1,
= DV1 \ 0V, and Sb! 0Vb n DV1.

Figure 2.1: Example of an arbitrary control volume including part of a moving rigid body.

The continuity equation for incompressible flow can be written over the complete
boundary of V,

f(u. n) dDV =f(u . nj) dS1 +f(ub flb) dSb = 0, (2.7)

or over the boundary of V1,

J(u.n1) doy1 = f(u.nj) dS1+ f (ub.nf) dSbf = 0. (2.8)


0V1 Sf Sbf

The complete expression for the momentum equations can be written as

fu dV1 = - f u (u. n1) dOy1 + --(f (Vu. n1) - pn1 dOy1


+
f F dV1) (2.9)
V1 0l' ôVj V1

2.2 Spatial discretisation


In this section the mathematical model, presented in the previous section, will be discre-
tised in space. The grid that is used, the discretisation of the geometry in two and three
dimensions, the discretisation of the continuity equation and the momentum equations,
and the discretisation of the free surface is discussed.

2.2.1 Mesh generation


To discretise the mathematical equations in section 2.1, the flow domain must be covered
by a computational mesh. In this study, a Cartesian grid approach is used. The com-
plete flow domain including the geometry of a moving body is covered by a rectangular
12 Chapter 2. Computational method

grid. This approach allows a complex shaped geometry without a time consuming grid
generation as for a boundary-fitted grid. Since the grid is structured and orthogonal,
bookkeeping will be very easy: There are only three cell face directions in three dimen-
sions and all neighbours of every cell are known directly. Even when a moving body is
treated, no regridding is required. There is also a lot of experience with Cartesian grids
for free-surface flow, which is an important feature of the applications of the method in
this thesis. Boundary conditions however are more difficult in a Cartesian grid; especially
when the geometry is not boundary fitted special care is required.

2.2.2 Discretisation of the geometry


In the method used in this thesis, the geometry of the flow domain as well as the geometry
of the moving bodies can obtain arbitrary shapes. To avoid staircase-type geometries due
to the Cartesian grid, the geometry must be allowed to cut through the grid arbitrarily,
thus creating so-called 'cut' cells. Moreover, in case of a moving body the position of
the body has to be updated every time step and therefore the 'cut' cells have to be
recomputed every time step. The way this discretisation is performed in the method is
explained below.

Geometry apertures and labels


To distinguish the flow area and solid geometry in every Cartesian grid cell information
about the solid geometry is stored. This information consists of the amount of solid
geometry in the cell and the fraction of each cell face that is open for flow. The value
Fb denotes the volume fraction of each cell that is open for flow, the so-called 'volume'
apertures and the 'edge' apertures A, A and A denote the fractions of the cell faces
that are open in z-, y- and z-direction respectively (see figure 2.2 for a two-dimensional
example).
Ah
L

Ah

Ah

Figure 2.2: Two-dimensional illustration of the definition of volume apertures and edge
apertures. Extension to three dimensions is straightforward.

This is a rather simple way to store information of the geometry at cell level and these
apertures will be sufficient for the discretisation of the equations of flow. Any additional
information, if necessary, can be calculated later, like computation of normals to the
.2 Spatial discretisation 13

boundary.

Once the volume apertures are known, the cells can be labeled as F(low) cells, where
the equations of flow are computed, B(oizndary) cells, where boundary conditions can be
prescribed, or e(X)terior cells, where nothing is done. F cells are defined as cells with
F > O, B cells are all cells adjacent to F cells, and X cells are all remaining cells.

FFFF
F\F F F F
BF FF
XB
XXB BB

Figure 2.3: Labeling of F, B and X-cells.

Computation of apertures
The computation of the apertures defined above has to be done very carefully, since this
information will play a crucial role in the discretisation. Especially at the boundary of
a moving body care is required, since there the geometry apertures will vary in time
provoking numerical irregularities. These irregularities will appear mostly in the pressure
signal, displaying a spiky behaviour as will be seen later. Therefore it will be important
to develop a geometry reconstruction procedure that will compute the apertures very
accurately. The most accurate computation will be of course an exact computation of
the apertures.

In two dimensions it is not very difficult to construct a procedure that computes the
apertures almost exactly. Suppose a moving body is represented by a number of marker
points in space that are connected with straight lines forming a closed polygon. For
non-curved geometries this representation will be exact. For curved-shaped geometries
the accuracy can be increased by increasing the number of marker points approaching
the exact geometry (figure 2.4).
Walking in a counter clockwise direction along the connected marker points that
define the polygon, the solid body area is always on the left and the flow area is always
on the right. This determines whether to be inside or outside the solid geometry. The
constructed polygon cuts through the Cartesian grid in various ways, therefore an efficient
and general procedure is required for the computation of the apertures. The Sutherland-
Hodgman clipping algorithm [22] suits this very well. The idea is that the intersection
points of the polygon with the spans of the four cell faces are stored and the volume of the
resulting polygon that is created by connecting the intersection points is calculated. The
complete procedure is illustrated in figure 2.5, where the non-curved polygon of figure
2.4 is clipped against a Cartesian grid cell (i,j) with cell-face coordinates x(i - 1), x(i),
y(j - 1) and y(j). First, the polygon is clipped against the line z = x(i), the span of
14 Chapter 2. Computational method

Figure 2.4: Example of a polygon describing a non-curved geometry (left) and approaching
a circular geometry (right).

the right cell face: All marker points are stored in the list of marker points. Starting
with marker point 1, every marker point p satisfying x(p) x(i) is added to a 'clip' list.
Further, every intersection point of the polygon with line x(i) that is found is added to
the list. After treating the last marker point, the clip list only contains the marker points
with x(p) <x(i) and some additional intersection points Ps with x(p8) = x(i). The points
in the clip list form a new polygon that is 'clipped' against the line x = x(i). Now the
new polygon is clipped against the line y = y(j) following the same procedure. Doing
the same for the lines x = x(i - 1) and y = y(j - 1), the remaining polygon is contained
in cell (i,j).
To compute the volume aperture, the area of the clipped polygon needs to be com-
puted. A simple formula exists for the area of a polygon with points p = 1, ..., n:

A0190 = (x(p)y(p+1) x(p+1)y(p))

which can be seen as summing the signed areas of the triangles (1,p,p + 1) for p =
1, ..., n - i or half the sum of the cross products (x(p) - x(1), y(p) - y(l)) x (x(p +
1) - x(1), y(p + 1) - y(p)) for p = 2, ..., n - 1. Finally the aperture becomes Fb(i, i) =
1.0 - Ac1ippedpo1ygo?/hxhv.

The value of the edge apertures can also be obtained from the clipped polygon, using
the intersection points with the cell faces and the counter clockwise numbering direction
of the polygon.

In three dimensions, in principle, something similar can be done. However, the com-
putation of the cut cells can become very complex now. Imagine that instead of a polygon
in two dimensions, the geometry will be represented by a polyhedron in three dimensions.
Allowing arbitrary shapes, this may be a very complex polyhedron and it can become
very complicated to compute the intersections with the Cartesian grid. In every cell the
part of the polygon inside that cell has to be computed. This means that the intersec-
tions of the polyhedron with the cell faces must be determined, to construct the new
polyhedron in the cell. In [58] for instance the surface of the geometry is described by
triangles, and the intersections with the grid can be calculated by analytical geometry
2.2 Spatial discretisation 15

5 i

Figure 2.5: Clipping procedure for the computation of the solid volume in a Cartesian
grid cell.

techniques. A three-dimensional clipping technique is used to 'clip' away the geometry


that lies outside the boundaries of the Cartesian cell. This results in a list of polygons
defining the part of the geometry within the cell. These polygons form a polyhedron, of
which the volume has to be determined This can be done similar to the two-dimensional
situation, by summing the signed tetrahedra that make up the polyhedron, instead of the
signed triangles.

Figure 2.6 illustrates the complex nature of the exact volume computation of cut
cells in three dimensions. The procedure described in [58] consists of first finding the
triangles that span the dimensions of the Cartesian cell. The triangles ¡2ABC, ACD,
LADE, AAEF, /AFG and LAGB intersect the Cartesian grid cell in the figure.
To find the intersection points with the cell faces and the cell edges the triangles are
clipped to the cell extents. The result is a list of polygons that define the portion of
16 Chapter 2. Computational method

/
/

Figure 2.6: Intersection of a geometry described by surface triangles with a Cartesian grid
cell.

the surface geometry enclosed within the cell. These polygons consist of face-polygons
at the faces of the cell and surface-polygons inside the cell. These different polygons
have to be linked to each other to determine their common vertices and edges, so the
proper volume can be defined. In the figure the bold line makes up the volume within the
cell of the intersected geometry. Needless to say is that a lot of bookkeeping is involved
in this procedure. This exact computation of three-dimensional cut cells is where the
Cartesian-grid method becomes really difficult and is a lot more complicated then the
two-dimensional procedure described above. However, for the applications in this thesis,
all these geometrical details about the surface triangles at cell level are less important
than for example high-order fluid flow calculations. A more approximate and simpler
method for describing the geometry and computing the cut cells would be more useful
now, keeping in mind that the exact computation can be used in a later phase or when
other applications come within sight. Though, the accuracy of the computation of the
volume- and edge apertures may not become too low, since the information of volume-
and edge apertures is the only information that is available at cell level and plays an
important role in the numerical procedure. Especially with moving geometries care must
be taken, since the apertures become time dependent and the pressure is very sensitive to
these aperture changes, as stated before. One necessary property is that the computation
must be volume conservative, or almost volume conservative. Otherwise, unphysical holes
or pressure waves can be created in the fluid.

Selected approach
The method that we have selected consists of filling the domain with a number of markers
that are uniformly spaced in every cell. Every marker is the center of a virtual rectangular
22 Spatial discretisation 17

box. Before the simulation is started, every marker that lies outside the solid geometry
is thrown away, and the solid volume of the cell is determined by the sum of the vol-
umes of the virtual boxes around the remaining markers. This can be seen as kind of a
Marker-And-Cell method for the moving geometry instead of for the fluid. When a cell
is completely filled with geometry (Fb O), this sum is equal to the volume of the cell
The number of markers in every direction is a measure for the accuracy of the
geometry description. If this number is called md, then the volume belonging to a marker
in cell with dimensions h, h and h is equal to hhh/nd3. If d = 1 the virtual box
is exactly the Cartesian grid cell. In figure 2.7 this is illustrated by a two-dimensional
example of a circular geometry. Three different accuracy levels are shown, rid = 1,2 and 4.

=i =2 =4

rivau T1Ii
P,
L
rJ

Figure 2.7: Two-dimensional example of a circular solid geometry described by a La-


gran gian subgrid with three different levels of accuracy.

Without moving bodies, the apertures are only computed at the start of the simula-
tion, so the markers do not need to be stored. With moving bodies however, apertures
change in time. Since the bodies are assumed to be rigid, the markers can be seen as
fixed points in the moving body, following the rigid body motion. This means that the
virtual boxes also follow the rigid body motion. The positions of the markers inside the
moving body are stored and also the size of their virtual boxes. This can be seen as
sort of a Lagrangian subgrid called the subgrid for simplicity hereafter and the virtual
boxes are called the subgrid cells. First, considering only translation, the orientation of
the subgrid with respect to the Cartesian grid remains the same. The subgrid follows
the translation and will move through the grid. In general, not all markers have their
surrounding box in one cell, but the box cuts through the cell faces and it will be con-
tained in more than one cell. Since their orientation is the same, the 'cut' volumes of
these subgrid cells in different grid cells will also be rectangular boxes. Given the position
of the marker and the size of its rectangular subgrid cell there is no problem in calcu-
lating the 'cut' volumes. Figure 2.8 is a two-dimensional illustration of a solid square
initially covering four grid cells with only one marker per cell (rid = 1). This square is
translated diagonally through the grid. The grid cells are cut by the subgrid and the cut
volumes are easily calculated given the position of the markers and dimensions of their
boxes. Volume apertures are now simply calculated by summing the cut volumes in every
cell of all the markers, in general for d > 1. Volume conservation is assured in this way
as well as accuracy of the geometry description to a level that can be adjusted by the user.
18 Chapter 2. Computational method

Figure 2.8: Translation of a solid square covering four grid cells with rid = 1.

Considering not only translating, but also rotating bodies, the situation becomes
somewhat more complicated. Rotating the subgrid changes the orientation with respect
to the Cartesian grid. Consequently the cut volumes will no longer be easily treated rect-
angular boxes but complicated shaped polyhedra in three dimensions. Exact calculation
of these volumes requires again a complex intersection and volume reconstruction algo-
rithm as described earlier in this subsection for exact volume computation as in figure 2.5.
Instead, to keep the orientation of the subgrid the same, the subgrid cells are not rotated
but only translated according to the motion of the markers. This will reduce the volume
aperture computation to an identical problem as of a solely translating body. However,
in the interior some holes and overlapping areas may be created in the geometry, but
these overlapping areas mostly cancel the holes (figure 2.9). Some care has to be taken
here, assuring that interior cells always have Fb = 0, but in practice the discrepancies
turn out to be smaller than 0.01%.

ríu
Figure 2.9: Rotation of a solid square covering four grid cells with d = 1. left: original
position, middle: exact rotation, right: rotation keeping the orientation of the subgrid
according to the Cartesian grid orientation. The dark shaded areas in the right figure
represent overlapping of the subgrid cells.

Using large three-dimensional moving bodies in combination with a very fine grid can
produce a lot of markers when rid > 1. The position and dimensions of the markers
have to be stored, so one can imagine that this may lead to memory problems. When
the complete moving body covers an amount of grid cells in every direction of 0(n),
the total amount of grid cells containing the markers is 0(n3), resulting in a number
of markers of 0(ri3 na3). In the interior of the moving body all cells are filled with
.2 Spatial discretisation 19

d3 markers, while the high accuracy is only required at the surface of the body. To
reduce the number of markers, before the simulation the markers in the cells in the in-
terior of the body (i.e. Fb = 0) are replaced by one single marker in the center of the
cell with body grid cell equal to the Cartesian grid cell. So in the interior the cells are
treated identically as with rid = 1. This leaves the total number of markers at about
O(n2 n3 + n3) = O(( + -) n3 fld3) O(n2 rid3) gaining about one order of n. In figure
2.10 this is illustrated with the circular geometry of figure 2.7 and d = 4. To save even
more memory, the procedure of replacing a number of markers in a square by one single
marker can of course be continued near the surface of the body.

Figure 2.10: Illustration of marker reducing in interior of moving body with d = 4. In


interior cells markers are replaced by one marker.

Until now, in the three-dimensional case only volume apertures have been discussed.
The calculation of edge apertures has to be done in a rather continuous way through time.
The way the subgrid is treated, assures a continuous change of the volume apertures, even
with d = 1. This subgrid is not very suitable for calculation of edge apertures, because
of its bumpy behaviour over the cell edges. Therefore another approach, based on a
piecewise linear reconstruction of the geometry is suggested here. This method is often
used for the reconstruction of a free surface in a Volume-of-Fluid (VOF) method (e.g.
Youngs [111]). The piecewise linear interface reconstruction (PLIC) method based on
filling ratios in a cell and a normal at the surface is found to be a proper alternative
for calculating edge apertures. The open fraction of a cell is called Fb, so the closed
fraction (the geometry of the moving body) is equal to i - Fb. A linear approximation
in every cell nx + fly + nz = c of the body surface is constructed given a normal n
and the value i - Fb. The edge apertures are determined by the fractions of the cell
faces that are cut out by the linear approximation. Since the PLIC-method does not
guarantee a continuous surface reconstruction, i.e. the separate constructed planes do
not necessarily connect continuously, different fractions of a cell face are cut by the two
constructed planes of the cells connected by that cell face. The edge apertures can not
be multi-valued, thus a simple averaging of these two different fractions is performed to
overcome this discontinuity. In figure 2.11 a two-dimensional example is shown. The bold
lines represent the PLIC reconstruction, the lines labeled i and 2 indicate the location
where the cell faces are cut by the body, after averaging the two FLIC values.

To guarantee a smooth behaviour in time of the simulated flow, it will be a nice


property if the edge apertures also behave smoothly. The edge apertures depend on the
20 Chapter 2. Computational method

Figure 2.11: Two-dimensional example of a possible PLIC-reconstruction of a solid body


required for the computation of edge apertures.

volume apertures Fb and the normals n of the PLIC-reconstructed surface of the moving
body. Earlier in this section it is concluded that the volume fractions are computed in a
smooth way. When the computation of the normals is done smoothly as well, the edge
apertures are expected to behave similar. There are different approaches to estimate
normals in a VOF-fleld. Analytically, the normal of the body surface in a cell is given
by V(1 - Fb)/IV(1 - Fb). In [23] this equation is discretised with respect to all the
neighbours of the cell, resulting in a system of linear equations that can be solved. In [30]
a simple eight-point discretisation is used in two dimensions and can be extended to a
twentysix-point discretisation in three dimensions. Rename the solid volume 1 Fb (i, j k)
in a cell by Cj,j,k. Consider the x-value of the normal (nX),3,k, which is discretised by a
weighted average of the two-dimensional x-normals in cells (i, k) in the planes j - 1,j
and j + 1:
(nx)i,j,k
l(()i±1 + 2(nX),k + ()1)
Here (nx)i,k is then discretised as a weighted average of aC/Ox

(n),k = «Ci+I,k+l - Ci_1,k+1) + 2(C+1,k - C_1,k) + (Ci+1,k_1 - Ci_1k_1)).


The discretisation is symmetric, thus the same result is obtained when the discretisation
is done by a weighted average of two-dimensional x-normals in cells (i, j) in the planes
k - 1, k and k+ 1. The discretisation of and (nz),3,k is done in a similar manner,
resulting in a twentysix-point discretisation of In this study, this discretisation
for the normals is executed for the calculation of the edge apertures. Since Fb values
behave smoothly in time, the normals that are only dependent of these values will behave
smoothly as well and therefore the edge apertures A2, A and A2.

2.2.3 Discretisation of the continuity equation


Equation (2.7) or (2.8) represent the continuity equation for incompressible flow, including
the motion of a rigid body. In every F cell, this equation is discretised. Consider the
2.2 Spatial discretisation 21

right Cartesian grid cell in figure 2.12, including a moving wall with velocity 12b The
velocities are staggered, i.e. horizontal velocities are placed at the vertical cell faces and
vertical velocities at the horizontal cell faces and assumed to be constant at the cell face.
n
Vb vn
t

Ub ___,. Ue Ub

Vs
A
ÎVS

Figure 2.12: Grid cells for discretisation of the continuity equation, using the uncut cell
according to equation (2.7) (left) or the cut cell according to equation (2.8) (right) for the
control volume (dashed lines).

Discretising the continuity equation over the volume in the cell that is cut out by the
moving wall according to the finite volume formulation yields

hAUe + hAv'2 - l(u& n) - hAv8 = 0. (2.10)


Here h and h are the sizes of the cell and nb the normal of the boundary. This equa-
tion states that the sum of all volume fluxes, including the volume of the moving wall,
equals zero. Since the conservation law is discretised, discrete conservation of volume is
guaranteed upto machine precision.
Two alternatives based on equations (2.7) and (2.8) can be distinguished.

Alternative 1:

A formulation independent of the configuration of the solid body is acquired when


equation (2.7) is used. The control volume can now be chosen as the uncut grid cell,
thus the cell including the boundary (right grid cell in figure 2.12). The velocities (ub, Vb)
(= Ub) of the solid body are assumed to be situated at the cell faces, in conformity with
the fluid velocities. The general form of the continuity equation for a Cartesian grid cell
with a moving boundary now turns into:

h (A e + (1 - A) ug - A' uv - (1 - A') u°) +


h (A v" + (1 - A) v - A y8 - (1 - A,) v) = 0. (2.11)

However, for reasons that will be explained in the next subsection, this general form is
not used.
22 Chapter 2. Computational method

Alternative 2:

The other alternative is based on equation (2.8). Equation (2.10) is discretised with
velocity u6 assumed to be constant on the boundary to obtain a general form. This
value of Ub = (u6, Vb) can be chosen dependent on the position of the boundary. For
non-rotating bodies, this velocity is just the moving body velocity. For rotating bodies,
the geometrical information that is provided by the apertures may be used to extract
the position of the boundary in the Cartesian cell, then the boundary velocity ub can be
adjusted to this position. The normal n6 is determined by the edge apertures, reading
n6 = (fli,x, n) = (h(AA), h(AA))II. Recognizing
= (h(A - A), h1(A - A,))II, the continuity equation can then be written as
h,,, (A - A u'° ) + h (A y'2 - A y8) -
e

h (A - A) Ub - h (A - A) b = 0. (2.12)

Note that when the moving body is not rotating, (2.11) and (2.12) are equivalent.

2.2.4 Discretisation of the momentum equations


The momentum equations in integral form (2.9) state that the change of momentum in
time in a control volume V1 is equal to minus the outward convected momentum through
the boundary of V1 plus the momentum flux through the boundaries due to the stress
tensor, consisting of a diffusive flux and a pressure term, plus a source term due to the
external forces. According to the staggered grid approach, pressures are defined at cell
centers of the Cartesian grid cells, and velocities at the cell faces. The control volumes
for the momentum equations in x-, y- and z-direction are chosen around the face ve-
locities, consisting of exactly half the volumes of the two adjacent grid cells by bisecting
the two grid cells (figure 2.13).

ti 'L

Figure 2.13: Uncut control volumes (dashed lines) for the x-momentum equations. uni-
form grid (left), non-uniform grid (right).

This also holds for cut cells, bisecting the cut volumes of both adjacent cells (figure
2.14). In this discretisation the boundary is assumed to behave linearly in a cell. Since
the only information that is available in a cell are the volume- and edge apertures, this
linear behaviour is sufficiently detailed. The size of the control volume in figure 2.14 is
easily computed as Vo1ume(ì)= (F'L'hh + Fhh).
2.2 Spatial discretisation 23

Figure 2.14: Cut control volume (dashed line) for the x-momentum equations.

Time derivative
The time derivative term in (2.9)
fa
fat udVf
Vf

is discretised in space using the midpoint rule: the velocity u is multiplied by the volume
of the control volume, resulting in
ra au
udVj=21--, (2.13)
Vf
J
where Ç is the volume of the control volume l/..

Convection
The discretisation of the convective terms

fu (u . nj)
av1

in equation (2.9), is done with V1 as in figure 2.13 and 2.14. This term represents the
advection of the momentum u by the mass fluxes u n daVf through the boundary of
V1. To avoid confusion between the identical symbol u used for the advected variable
and the advection variable, the momentum u is replaced by the symbol = (, ,, q).
The convective term in x-direction can now be written as

f
OVj
(u nf)

First consider a control volume in the absence of a boundary, as in figure 2.13.


Since the velocities are positioned at the cell faces, they are not directly available
at the boundaries of the momentum control volume. A choice has to be made for the
momentum at 3v1 and the mass fluxes through 8V1. The following simple averaging
24 Chapter 2. Computational method

-.--, tt'

v72w vne

h u -, uc u6

t
V8W V86

u8

Figure 2.15: Control volume and positioning of variables for discretisation of convection.

for ç& is performed:

ç= +Ø), ç=
1
+ = + ), (2.14)

and they are assumed to be constant on each different segment i, r, u and d.


This simple averaging ensures a skew-symmetric convection matrix if the mass fluxes
are properly discretised. Even if the grid is non-uniform in y-direction no weighting is
included in the averaging for this reason. The skew-symmetric property of the continuous
convective operator is thus conserved and this discrete skew symmetry provides nice
numerical properties as will be seen later. The mass fluxes, renamed as m, are also
averaged:

m1 = (u'h + îth), mT = (u6h + uchy),

rn'8 = (v78t11h + v'86h), md = (v88'8h + v86h)

The discrete convective term in x-direction now becomes

f(u n1) döVf m1 - mdçb + mTçb +


aVf
2.2 Spatial discretisation 25

Evaluating this term into coefficients for the face-values of ç, leaves

m1çb - mdq + mTçb + m' =


+ uc)h) - (v8wh + vseh) +
+ uc)h) + (vm0h + vneh) +

- vswh - vh + ueh + vnwh + vh) q.


Here the skew symmetry is recognized: The coefficient for , which appears at the diag-
onal of the coefficient matrix, is identified as the total mass flux through the boundaries
of the two cells which equals zero since conservation of mass holds. The off-diagonal
coefficients for , , and represent the mass fluxes through the western, south-
ern, eastern and northern boundaries respectively and are clearly skew symmetric. As a
consequence of this zero diagonal property of the convective matrix, together with the
simple averaging of the momenta that make up the skew symmetry, the discretisation
of the convection is completely determined by the discretisation of the divergence. This
will be the basis for the discretisation of convection and these properties should hold
generally, therefore they should hold in the vicinity of a moving boundary as well.

Consider a situation as in figure 2.16. Since the discretisation of the divergence will

VgW

uc

r us

se
ve,

Figure 2.16: Positioning of staggered variables for discretisation of convection near a


moving boundary.

determine the discretisation of convection completely, it is important which formulation


of the continuity equation is used in subsections 2.1.3 and 2.2.3. Two alternatives, in
'Note that on a uniform grid this discretisation is identical to a central finite difference scheme.
26 Chapter 2. Computational method

analogon with the alternatives in subsection 2.2.3, may then be distinguished.

Alternative 1:

The first alternative uses the general formulation with equation (2.11) as discrete
form of the continuity equation. The velocities of the moving body can now simply be
considered in staggered position in the Cartesian cells, independent of the configuration
of the solid body. The discretisation of the convective terms then becomes:
_m1 - m'ç + mTçb + m'çb =

_((Auw + (1 _A)u+Auc+ (1 A)ug)h) +

_((Avsw + (1 - A)v)h + (Av + (1 - A)v)h) +

+ (1 A)u + Auc + (1 A)ug)h5) +

+ (1 A)v)h + (Av + (1 A)v)h) +

X
+ (1 - A)u)h (Avsw + (1 - A)v)h +
4

+ (1 - A)v)h + (A:ue + (1 - A)ug)h +

(Av'° + (1 - A)v)h + (Av" + (1 - A)v)h) . (2.15)

Here we recognize the mass fluxes through the boundary, and the central coefficient van-
ishes since this indeed represents the discrete continuity equation (2.11) applied at the
two adj acing cells. However, though equation (2.11) fixes this discretisation, the choice of
the control volumes like in figure 2.14 does not justify this discretisation. The magnitude
of the mass fluxes in (2.15) do not need to be according to the mass fluxes as expected
in the control volume. This can be best illustrated by a situation as in figure 2.17.
Ignoring vertical velocities, the mass fluxes m1 and mr, according to the choice of the con-
trol volume (the bold dashed line), are equal to (Auw +Auc)h and (Auc+Aue)h
respectively. The discretisation (2.15) however, proposes mass fluxes mt = (AuvJ + (1-
A)u + Au' + (1 - A)ug)h5 and mT = (Auc + (1 - Acjug + Aue + (1 -
suggesting a control volume like the dotted line. These mass fluxes appear in the coeffi-
cient matrix for the momentum vector, and in this case they may be disproportional to
the control volume ft In practice the defect may turn out to be small, since UV, UC and
uc will be almost equal to each other and due to to the rigidity of the body the mass
fluxes m1 and mr might almost cancel for the components and . Nevertheless, it is
better to use the more physical approach and keep in mind the dashed control volume.
Note that both approaches do not differ from each other when the velocities of the solid
body are equal to zero, since then the mass fluxes of the solid body do not contribute to
2.2 Spatial diseretisation 27

Figure 2.17: Control volume (dashed line) vs. control volume suggested by discretisation
(2.15) (dotted line).

the mass fluxes m1 and mr.

Alternative 2:

For reasons mentioned above, equation (2.12) is chosen as continuity equation, re-
quiring a constant velocity ub at the physical boundary in a cell instead of using the
staggered grid boundary velocities u, u, v and v.

Consider the control volume in figure 2.18, identical as in figure 2.14, divided in seven
segments and assuming boundary velocity u holds in the entire right cell and velocity
u in the left cell.
The mass fluxes through the different segments of the boundary are computed as follows:

Figure 2.18: Momentum control volume divided in segments


28 Chapter 2. Computational method

In' - EACh
2 xl/ (ue+uc),
m2 = - A)hy(ue + u) +
In3 = '4sehese
m4 = (A - A)hu + (.4 4se)hev,
= (AhyuC + (A - A)h31u) + -
m6 = AhvThw,

mT =

These choices are motivated by estimating the flux through a segment by a velocity times
area and keeping in mind the discrete continuity equation (2.12). Horizontal mass fluxes
in m', m2 and m5 are simply averaged between two known velocities, the mass flux in rn5
is the length of segment 4 multiplied by u n. The vertical mass fluxes in segments 3, 6
and 7 are equal to the staggered vertical velocities multiplied by the length of the segments
and the vertical mass fluxes in segments 2 and 5 depend only on the boundary velocities
and the normals at those segments. One criterium is that the continuity equation also
holds for the momentum control volume. To check this the (signed) mass flaxes can be
summed:

m1 + m2 - rn3 - in4 - in5 + m6 + in7 =


(hy(Auc + Aue) - h(A - A)u - Ah:vse + h(A - A»v -
Ahuc - (A - A)hu, + A"h'v' + Ahv - (AW -

Equation (2.12) applied at the two adjacing Cartesian cells is recognized, so this sum
is equal to zero.

A general form, that can be used in every cell independent of the configuration of the
geometry can be written in terms of mass fluxes through the right, upper, left and lower
boundary, as in the first alternative, appearing as coefficients for the momenta , , ç
and ç respectively:

f(u . nf) d3V1 = m1q - mdq + mTçb +


av1

where the values of are again averaged according to (2.14), using boundary values of
= Ub when the corresponding aperture is equal to zero. The mass flaxes are worked
2.2 Spatial discretisation 29

out below.

= h((AUe + Auc) + (A - A)u),


m' = (h:(Avne + max(O, (A - A)v)) + h(Avnw + max(O, (A - A)v))),
m1 = + Auw) + (A - A)u),
md = + max(O, (AW - A)v)) + h(Avse + max(O, (A -
Considering again figure 2.18, mr consists of the sum of the signed horizontal mass
fluxes through segments 1, 2 and 4 (m', m2 and m4), m1 is formed only by horizontal
mass flux through segment 5 (m5), m' by vertical mass fluxes through segments 6 and
7 (m6 and m7) and m' by vertical mass fluxes through segments 2, 3, 4 and 5 (m2, m3,
m4 and m5). The max-functions that appear in mL and are required to distinguish
whether the boundary fluxes are applied at the upper or lower segments, or in other
words, whether the body moves into the cell or out of the cell. Note that the mass fluxes
m1 and m' in figure 2.17 have the right magnitude now. The contributions of u and 4
vanish and the mass fluxes can be seen as fluxes only through the dashed control volume.

Evaluating integral (2.16), the coefficients for , , , 4 and can be determined,


and the skew-symmetric properties again appear. To show the subtlety the central coef-
ficient is worked out below and is shown to be equal to zero.

The central coefficient CC appears in every term in (2.16), since in every averaging for
, the central term ç appears. Thus the central coefficient is equal to C' = (m1 -
md + mr + mh). Evaluating this sum of mass fluxes leaves
m1 md + mr + mu =
+ Aue) +
+ Av) + h(Av + Avne) +
- A)u + (A - A)u) +
- A)v + h(A -
It is not very difficult to see that this sum is equal to zero, provided that equation (2.12)
holds in both cells, hence CC = O. To see why the max-functions disappear, look at the
term max(O, (A - A)v) - max(O, (A - A)v). At least one of these max-functions
is equal to zero, since A - is positive, negative or equal to zero. Thus the complete
expression can be replaced by the nonzero term. Finally the minus sign can be eliminated
and the term reduces to (A - A)v.
Concluding the discretisation of the convective terms, a discrete skew-symmetric for-
mulation has been found including arbitrarily moving boundaries, thus conserving the
30 Chapter 2. Computational method

symmetry properties of the continuous convective operator. Further, the discretisation


that is found can be applied in a general way, solely with information of the apertures
and velocities of the moving boundary. Moreover, the choice of the control volumes is
justified by the discretisation with mass fluxes proportional to the magnitude of the con-
trol volumes.

Diffusion
The discretisation of the diffusive terms
I f (Vu. n1)dOV1
av1

in equation (2.9) requires derivatives of the velocities at the boundaries of the control
volumes. On each segment of the control volume (in figure 2.18 segments i to 7) these
derivatives are computed. The integrand for the x-momentum equation can be written
as íiOn/ôn, where Ou/On is discretised as the difference of the velocity characteristic on
the side of the specific segment and the velocity u divided by the distance between the
location of these velocities which has to be estimated, since these velocities are assumed
to be constant over the entire control volume in the finite volume method. So on segment
k an accurate discretisation of Ou/On becomes
OUUkUc (216
On Inkl
Here Inh is the estimated distance between the location of Uk and u and is computed
by

I
k1 =Vk-,
Ak
where Ak is the area of segment k and Vk is a characteristic volume for segment k. The
precise choice of the volumes Vk is described in detail in [23]. This is done in a way that
conserves the symmetry property of the continuous diffusive operator, i.e. the coefficient
matrix for diffusion is symmetric, moreover the matrix is also negative definite
The above choices of Vk and Ah are required for an accurate description of the diffusive
fluxes. However there are some restrictions that have to be coped with. Using explicit
time integration, some restrictions with respect to stability have to be coped with (see
section 2.3.2).

Pressure
The pressure term in (2.9)
_Ifni dOy1
ÔV1

is treated discretely with the same philosophy as the convective and diffusive terms,
namely that the symmetry of the underlying continuous operators is preserved. Writing
.2 Spatial discretisation 81

the equations in differential form, the continuity equation reduces to V . u = O and the
pressure term reduces to Vp. Analytically, the divergence operator (V.) is minus the
transpose of the gradient operator (V). This analytical property should also hold for the
discrete divergence operator Mh and the discrete gradient operator Gh: M,, = Gr. This
means that at the cell faces, where the velocities are defined the pressure gradient also
has to be defined there, which is automatically satisfied by the choice of the staggered
grid. Moreover, the coefficients of M,, working on u must be of equal magnitude as the
coefficients of G,, working on p. To illustrate this choice and the property Mh = Gr, a
one-dimensional example is shown below.

pl pr
prr
UC

Figure 2.19: One-dimensional example to illustrate the discretisation of the pressure term
and the continuity equation, the areas of the cell faces w, c and e are equal to A'', AC and
A respectively.

The choice of the continuity equation gives as coefficients in matrix form for (u", uC, ue)T:

A° AC O
O _AC A'
( O O A'
Minus the transpose of this matrix equals
lAW
4c
O
AC
o\
O

\ o A' AeJ
which should be the coefficients for (pC7.,p,pt). As can be seen the discretisation of the
pressure term reduces to simply multiplying the area of the cell face with the pressure
difference of the adjacing cells (e.g. A'(p' - p1)). This is also completely justified by
the choice of the momentum control volumes as will be shown now. Consider again the
control volume for the x-momentum equation in figure 2.18. A pressure p' holds in the
complete east Cartesian grid cell and a pressure p1 holds in the complete west cell. The
integral

f pn,
9V1

over the boundary of the control volume can be written as the sum of the contributions
of the segments 1, ..., 7. Segments i and 2 contribute pT h in total, segment 3 gives
zero, segment 4 and 5 contribute pT h (1 - A) and p1 h, A respectively, while the
integrals over segment 6 and 7 equal zero. Summing these contributions completes the
32 Chapter 2. Computational method

pressure term:
AC
f Pflr d8V1 - j_ix y (r - rl), (2.17)
8v1

which is equal to the pressure difference multiplied by the open area of the central cell
face and thus the only geometry information that is required are the apertures A. A h5
is present in the continuity equation of the left cell in figure 2.18 and A, h5 in the
right cell, both coefficients of uc. This shows that the analytic property M,, = G is
preserved.

External forces
The discretisation of external forces in general can be done according to equation (2.9)
using the volume integral
fFdVf. (2.18)

Discretising the x-component of this term with the midpoint rule yields

-
f F dV1 - FCX

where Çf is the volume of the control volume V1. In this study the only external force
working on the fluid will be gravity, and is always pointing in the z-direction. The gravity
term F in (2.9) is equal to (0, 0, p19)T. In the z-momentum equation this force could be
discretised using the midpoint rule as

fF dV1 = P9

However, if gravity is the only force present, and all velocities are equal to zero and
have to stay equal to zero, the pressure term must cancel the gravity term since all
other terms are equal to zero (no convection and diffusion) and should result in the
hydrostatic pressure. Therefore the discretisation of the gravity term has to be adapted
to the discretisation of the pressure term. According to equation (2.9), the hydrostatic
pressure and the z-momentum equation as in figure 2.20, the discretisation of gravity F,,
must result in
p9_jCJ, (ti
z x '.Phydrostatic
d
Phydrostatic

Evaluating (2.19) results in = A h p1ghz, suggesting a control volume of mag-


nitude A hh which is in general not equal to the magnitude of the momentum control
volume ÇJ of V in cut cells. In uncut cells, these volumes are equal. However, using
Gauss' divergence theorem for the vector (0,0, p1gz), (2.18) can be written as a boundary
integral
f Pi 9Z n
OV1
2.2 Spatial discretisation 33

C g

Figure 2.20: Momentum control volume V1 near a boundary for staggered grid location c
to illustrate the discretisation of gravity.

And now this approach will result in the correct discretisation for gravity, by writing this
force p1g as a boundary integral of the hydrostatic pressure potential Pf 9Z (compare
2.17):
F f pgz n d8Vj = A h (p1gzu - p1gzd) = A h pjgh.
=
0v1

Since every conservative force can be written as the gradient of a potential function,
it is recommended that the discretisation of an external force of this nature is derived
from a boundary integral over the momentum control volume of the force potential.

Symmetry-preserving properties
As explained before, the discretisation of the conservation laws is done in such a way
that it preserves the symmetry properties of the underlying continuous operators. The
advantages and consequences are outlined below.

The discrete divergence operator Mh can be split into a contribution M, working on


the interior velocities u and a contribution M working on boundary velocities Ub. The
discrete continuity equation can then be written as
Mh°u = Mub. (2.20)

Since the discrete gradient operator Gh is only working on pressures inside the fluid, the
following symmetry property holds:
M,=G'.
As stated before, the discrete convective operator Ch is skew symmetric (Ch = Ci) and
the discrete diffusive operator Dh is symmetric and negative definite. Together with the
discrete gravity F and the matrix of momentum control volumes the momentum
equations can be written as

= Ch(u, ub)u + -(Dhu - Ghp + Fe). (2.21)


34 Chapter 2. Computational method

To investigate the stability of the spatial discretisation, the evolution of the discrete
energy can be evaluated. The total kinetic energy in the fluid can be written as

= pIuI2 dV1.
Vf

In discrete form this integral is denoted as the inner product

E,,kzfl = (u, pQ1u).

The change of discrete kinetic energy in time can then be evaluated:

dk = ld (u, pílju)lid
= (u, p1u)\ + \U lid pf U

= ii d \
+ii(u,pcu
d

= ((pCu, u) + (Dhu, u) - (Ghp, u) + (F,,, u)) +


((u, pChu) + (u, Dhu) - (u, G,,p) + (u, F,,))
= ((p(C + C,,)u, u) + ((Dh + D,,)u, u)) -
(Ghp,u)+ (Fh9,u).
Since Ch = Cg" and G71 = Mh°, the last expression can be written as

= ((D,, + D) ) +(p,M,u)+ (F,,u).


From this we can derive that, provided the symmetry properties are preserved, convection
does not contribute to the kinetic energy and diffusion dissipates energy since Dh is a
negative definite matrix. Furthermore, without moving bodies, Mub = O. Consequently
Mh°u = O, hence from (2.22) the pressure does not contribute to the energy. However,
in the case of a moving body, the pressure has a contribution to the energy equal to
(p, M,,°u) = (p, Mub), completely controlled by the boundary velocity Ub. The gravity
contribution (F,, u) can be written as (Ghpgz, u), therefore
_M,Tpgz,u) = (pgz,M,u)
(FK,u)
so that the gravitational force is treated completely consistent with the pressure term
and does not contribute to the energy in the absence of moving bodies. In fact, the only
energy increase may be due to

(p, Mh°u) + (F,,, u) = (p - pgz, Mub).


Since the gravity term pgz is already contained in p, the energy change is caused by
pressure variations around the hydrostatic pressure due to the motion of the moving
2.2 Spatial discretisation 35

body, which agrees with our intuitive idea. Thus, in a closed system without a free
surface, the change of discrete energy is equal to

= ((Du + D)u, u) + (p - pgz, Mub).


Since diffusion dissipates energy and the increase due to pressure is completely controlled
by Ub, the spatial discretisation does not give rise to an uncontrolled energy increase. Of
course, free-surface effects may have influence on the energy, but this is discussed later.

2.2.5 Discretisation of the free surface


The most important property of the flow that is studied in this thesis is the presence
of a free surface. The flow is mainly determined by this free surface, therefore a proper
discretisation is required to capture its dynamics.

Free surface apertures and labels


The method makes use of a modified volume-of-fluid (VOF) method for the reconstruction
of the free surface. First, an indicator function F5 is introduced, the so called free-surface
apertures, indicating the fraction of a cell that is occupied with fluid. Thus O F5 1.
Then, all F(low) cells are given a free-surface label, to distinguish the air from the liquid:
E(mpty) cells: F3 = O
S(urf ace) cells: all cells adjacent to E-cells
F(luid) cells: all remaining cells
This procedure of labeling is illustrated in figure 2.21.

0.0 0.0 0.0 0.0 0.0 0.0 E E E E E E

0.03 0.0 0.0 0.0 0.18 S E E

072. 0.27 030 076 0 F S S

1.0 10 087 F F F

1.0 0.81 F F

Figure 2.21: Example of free-surface aperture distribution F5 (left) with its corresponding
free-surface labels (right).

In E cells, e.g. in the air, the flow is assumed to be irrelevant to the liquid dynamics,
so the Navier-Stokes equations are not solved here. In S cells a boundary condition for
36 Chapter 2. Computational method

the pressure is applied, derived from equation (2.6). Further, momentum velocities are
computed at the staggered grid locations at F-F, F-S and S-S cell faces. Also velocities
at S-E cell faces and even at E-E cell faces appear in the computation of these momentum
velocities, so velocity boundary conditions at these locations are also required (see figure
2.22). More details about the free-surface boundary conditions are described below.

Figure 2.22: Location of required free-surface boundary velocities (bold arrows) and inter-
nal momentum velocities (non-bold arrows).

Pressure boundary conditions


In every S cell a boundary condition for the pressure is required. As stated before this
boundary condition is based on the normal stress condition (2.6). Regarding the type of
flow studied, viscous effects and surface tension do not play an important role, therefore
the terms 2a/a and aic are neglected here. Now the normal stress condition simply
reduces to p Po at the free surface. A sufficient boundary condition would therefore be
setting the pressure to p in S cells, because the pressure is assumed to be constant in a
cell in the finite volume method. However, since the free-surface apertures F5 lie between
O and 1, the exact location of the free surface can differ in a cell. For a more accurate
approximation of the free-surface boundary condition the pressure in the S cell, which is
located in the center of the cell, can be adjusted by linearly interpolating with a pressure
in an neighbouring F cell that is chosen based on the orientation of the free surface. In
this way the boundary condition p = Po is moved from the cell center to the intended
free-surface location (figure 2.23). The free-surface pressure p15 is linearly interpolated
to the center of the S cell by
Ps + (î - 1)pF = 1Pfs, (2.22)

where i = h/d. Sometimes, an S cell has more than one neighbouring F cell. Therefore,
first the orientation of the free surface is determined based on neighbouring F5 values. In
two dimensions this orientation will be horizontally or vertically. Then, in the direction
that is found, an F cell will be searched to interpolate with. If no F cell is found, also in
other directions F cells will be searched. If no neighbouring F cell is found, the value of
Ps is set to po.
2.2 Spatial discretisation 37

Figure 2.23: Illustration of pressure interpolation in S cells.

When curvature and surface tension effects do play an important role, for example
in micro-gravity environments, the term cric can not be neglected since this will be the
driving force of the flow. A detailed description of the discretisation of this term, including
the effect of contact angles, can be found in [23].

Velocity boundary conditions


S-E velocities

Boundary conditions for the velocity are required at S-E cell faces since they appear
in the momentum discretisation of neighbouring momentum velocities as stated before
(figure 2.22). These velocities are present in the coefficient matrix for convection. Re-
garding the discrete energy, it would be nice if conservation of mass is demanded in S
cells, since then the coefficient matrix for convection will keep its skew-symmetric prop-
erty, required for energy conservation. Other studies almost always apply these mass
conservative boundary conditions, and it is not very difficult to implement, though some
different configurations have to be considered. However, this approach also leaves some
disadvantages especially near cut cells. It turns out that free-surface velocity boundary
conditions are of crucial importance for the flow and imply a high sensitivity with respect
to pressure variations and energy conservation. They must be treated very carefully, even
to prevent the simulation from breaking down [10]. In section 2.4 the discretisation of
these boundary conditions is described in detail, with their effects on the pressure and
energy. A comparison is made between different approaches with their advantages and
disadvantages.

E-E velocities

E-E velocities only appear in the discretised momentum equations for S-S velocities.
These velocities are more or less directed tangential to the free surface, therefore the
tangential boundary condition (2.5) is discretised here to obtain the E-E velocities. The
derivatives and 9u/9t are discretised in Cartesian directions, since it is difficult
98 Chapter 2. Computational method

to compute them for arbitrary directions. Consider the two-dimensional example in figure
2.24, the E-E velocity utm is computed from a central discretisation of

(8u ôv'\
=0,

resulting in
= us - - vto).

where u is a known S-S momentum velocity and Ve and Vto are known S-E boundary
velocities. Hence to compute E-E boundary velocities, it is important to compute first
all S-E boundary velocities. The extension to three dimensions requires one of the three

U
E E

hy
- vwS
7-1- V

Figure 2.24: Two-dimensional situation with an E-E boundary velocity utm that has to be
computed.

following boundary conditions, dependent on the direction of the E-E velocity and the
S-E velocities:
3u t9v
ôy 5x
- + - =0,
av aw
Oz
+=0
ay
and
-aw+ -au=0.
ax az
Since one E-E velocity may have more than i neighbouring S-S velocity, an averaging
has to be performed in this case. More details about this can be found in [23].

Displacement of the free surface


Because time-dependent problems are considered, the free surface position will change
in time. Therefore, every computational time step the free-surface apertures F0 and
free-surface labels have to be recomputed. The free surface is described by the equation
S(x, t) = O and its motion can be described by equation (2.4). The free-surface aperture
or VOF function F5 is advected in time based on the donor-acceptor method, developed
2.2 Spatial discretisation 39

by Hirt and Nichols [32]. The idea is that a flux 5F3 through every cell face is computed
from the velocity u n, the area A of the cell face and the time step ôt:
5F3 = u nAOt. (2.23)

Then the total change F3 of the VOF function F8 in a cell can be computed by summing
the fluxes through the cell faces of the cell:

LFS = (2.24)

where m is the total number of cell faces. For a two-dimensional Cartesian grid with cell
faces e, n, w and s (2.24) becomes:

!F8 = Ot(UeAeh + vThAThh - flWAWh - v5A3h) (2.25)

In the interior of the fluid, in F cells not in the neighbourhoud of the free surface,
LF3 = O. It is easily seen that (2.25) fulfills this property, since the term between paren-
theses represents the discrete divergence, which is equal to zero. Near the free surface
however, the cells are not completely filled with liquid, which makes the computation of
the fluxes more complicated. The fluxes computed by (2.23) may exceed the value of F8
in the donor cell. Consequently, after performing equation (2.24) in these cells, F8 may
adopt values below O or values above 1. Therefore, limiters are necessary to prevent this.
More details about this can be found in [23, 32].

Still, applying the limiters in the original VOF method of Hirt and Nichols, it is
possible that some liquid mass is lost or gained. Furthermore, so-called 'flotsam' or
'jetsam' can be created. This means that small drops of fluid are separated from the
bulk of the fluid during the free-surface advection, which usually is not physical. This
is the reason that a modified VOF method is used, modified in the sense that at the
free surface in S cells, the original advection method of Hirt and Nichols is adapted. A
local height function is introduced in a 3x3 block (in 2D) and a 3x3x3 block (in 3D) of
cells around the concerning S cell. First, the orientation of the free surface is determined
(horizontal or vertical), depending on the values of F3 in this block of cells. Then, the
horizontal or vertical local height in each row or column in this block is determined by
summing the VOF fractions (figure 2.25). The local height function can be represented
by S(x, y, t) = h(x, t) - y = O. The height of the free surface in the center column is now
determined by the advection equation (2.4), resulting in
Oh
at
= V - 0h
n.
Ox

For the reconstruction of the free surface in a cell, the original method of Hirt and Nichols
is based on a horizontal or vertical orientation of the free surface. A more accurate
representation of the free surface can be applied using the method of Youngs [111], based
on a piecewise linear interface reconstruction (FLIC). Fluxes across the cell faces can
then be computed more accurately. In this method it is also possible to advect the free
surface using the local height function described above. A detailed comparison between
40 Chapter . Computational method

h = 1.5

h = 0.9

h = 0.3
Í -x

Figure 2.25: Determination of local height function based on F3 values in a S by S block


of Cartesian grid cells.

the method of Hirt and Nichols and the method of Youngs both with and without local
height function can be found in [23]. The overall conclusion of this comparison is that the
method of Youngs is more accurate than the method of Hirt and Nichols for prescribed
velocity fields. However, it turns out that when the velocity field is computed from the
Navier-Stokes equations, the method of Hirt and Nichols combined with a local height
function gives the best results, and the latter will thus be used for the reconstruction and
the advection of the free surface.

2.3 Temporal discretisation


The discrete continuity equation (2.20) is performed at the new time level n + i to ensure
a divergence free velocity field at the new time level:
(M)Th+un+l = (M)"-4ur'. (2.26)

The discrete momentum equation (2.21) is discretised in time using an explicit forward
Euler method for the time derivative. All diffusive terms and external forces are discre-
tised at the old time level n. The pressure term is discretised at the new time level. This
results in
U
= C(uTh,u)uTh + (Dur - G»p1 + (Fu'). (2.27)

Note that both in (2.26) and (2.27) the divergence and gradient operator are also written
with a superscript n + 1 since these operators depend on the geometry, that changes in
time when moving bodies are used. The divergence and gradient operator have to be
treated at the same time level to ensure the condition Mh° = Gr. Also the control
volume is provided with a superscript n + i and the convective and diffusive operator
are computed explicitly at time level n. At time level n the symmetry properties of the
convective and diffusive operator are ensured by the previous time step. This ensures
that the computation can not break down due to increase of kinetic energy. The velocity
at the new time level now becomes

U' = u + ôt (r')1 (_C(um, U)U + - G1p'' + (Fer)). (2.28)


2.3 Temporal discretisation 41

2.3.1 Poisson equation for the pressure


The pressure pfl must be determined in such a way, that equation (2.26) holds. This
can be achieved by solving a Poisson equation for the pressure as will be shown below.
First rename

= u + 0t (Çfl+1)1 (_C(un u)uTh + Du' + (Ff)')),


then equation (2.28) can be rewritten as
u = - at (Qfl+l)lG7-4-lpn+l (2.29)
Combining this equation with the continuity equation (2.26) gives
= (M,)1(ü - at (Ç')'G1p'1) = (M)'1u1,
resulting in the Poisson equation for the pressure
Ç1n+l
= -i-- ((Myu + (M)'1u1)
or, since Mh° = G

_(M)1(M)1Tpn+1 = ((M)+'u + (M)'u'). (2.30)


at
This equation is solved using a successive overrelaxation (SOR) iteration method, i.e.
a Gauss-Seidel method with overrelaxation. The relaxation parameter is automatically
adjusted to optimize the convergence [3]. A more sophisticated method is not necessary
for applications in this study, since this will not significantly increase the performance
[981. Once the pressure is computed from this Poisson equation, the velocity field can be
updated by (2.29).

Without a free surface, no boundary conditions for the pressure are required and the
pressure is completely determined by the Poisson equation (2.30). With a free surface,
the pressure boundary condition (2.22) is applied. This condition is added to the pressure
Poisson equation, so in every S cell the pressure p5 is substituted by 17p5 + (1 -
During the SOR-iteration process the value of ps will be updated every iteration according
to this substitution.

2.3.2 Stability
Convection
Using explicit time integration, like the forward Euler scheme, results in a CFL-restriction
like at for the time step, where h is the size of the uncut cell.
A more precise investigation in the convective stability criterium can be done based
on matrix analysis. The stability of integrating the convective term

= -1Ch(u, Ub)U
42 Chapter 2. Computational method

can be derived from the eigenvalues of j'Ch(u, 'ab). A global estimate for uncut cells
where 11 = 0(h2), Ch = 0(uh) and the eigenvalues of 1Ch = 0(f), results in a time
step limit Elt < . From this the Courant-Friedrichs-Lewy number ,
or CFL number,
can be derived. Stability requires CFL < 1. Using cut cells, may have entries that
become arbitrary small, therefore may have entries that become arbitrary big. One
could expect high eigenvalues of 71Ch when 2j becomes small, resulting in a very re-
strictive time step. However, it can be proven that in the case of non-moving boundaries,
the eigenvalues of 'Ch are bounded in terms of the spacings of the Cartesian grid, i.e.
the small cut cells do not lead to a sharpening of the stability restriction that holds for
uncut cells and the forward Euler method is stable if the CFL condition is satisfied for
cells as large as an uncut cell [97,99]. To give an idea of the proof, consider the momen-
tum control volume for UC in figure 2.26.

Figure 2.26: Example of a small momentum control volume to illustrate the stability of
the time integration.

The convection matrix Ch(u, 'ab) has a skew-symmetric form, consisting of mass fluxes.
The row corresponding to uc is evaluated as

( _md rril O m' m' ) (2.31)

which are the coefficients of


/ US \
U'»

uj
From section 2.2 and keeping in mind the control volume of figure 2.26, the mass fluxes
.3 Temporal discretisation 43

m7 can be written as:

md = 0,

m= h(Auc + Auw),
mT = hy(Aue + Auc),

= (hvne + hv).

Applying the Gerschgorin circle theorem to every row (2.31), one can estimate an upper
bound for every eigenvalue of Ch. All diagonal entries of Ch are equal to 0, since the
matrix is skew symmetric. Further, according to Gerschgorin, an eigenvalue of Ch lies in
the circle with center O and radius r = mdl + mtj + lmnl + Imul. Since Ah and
Ah have order of magnitude m1 and mT are O(u). Discrete mass conservation
guarantees md+mt_mT_mu = 0, consequently mL = Q(Lu). So the relatively small m1
and mT together with mass conservation keep mu also small and the eigenvalue belonging
to this row lies in the circle with center O and radius r = O(.Lu). The corresponding
eigenvalue of 1l71Ch is then estimated by °(t)7 the same order as the eigenvalues for an
uncut cell. Of course there are more geometrical cases to deal with, but it can be proven
that in all cases the eigenvalues of lj'Ch have the order of magnitude as for an uncut
cell, so this will not lead to a more restrictive time step [97].

In the case when the boundary is moving, the stability analysis becomes somewhat
different. Consider the same control volume as in figure 2.26, but now the boundary is
moving perpendicular to the wall with velocity y6.

-, un
Vfl11) vne

uw uc -
I I

Figure 2.27: Small control volume in the vicinity of a moving boundary.


44 Chapter 2. Computational method

The mass fluxes mt now become

md = (hvb+hvb),
m= h(Auc+Auw),
mr = h(Aue+Auc),
mU =

The difference with the former situation, where vb = O, is that md O, moreover


m' = O(hv&). Consequently, the relatively small mt and m' (O(.Lu)) do not keep
ma small by discrete mass conservation, therefore also m' = O(hvb). The eigenvalues
of 7'Ch can be estimated by Gerschgorin again and will be in the circle with center O
and radius of order l71hvò. Ij does have its influence on the eigenvalues now and they
will grow when Ç21 becomes smaller. The nice behaviour of keeping the same time-step
restriction as for the uncut cells is now spoiled by applying a boundary velocity Vb. With-
out decreasing at dramatically or applying other further steps, stability is not guaranteed
anymore: The velocity uc will blow up when becomes too small.

To ensure stability with moving boundaries, without the need of a very small time
step, some additional measures have to be taken. Without diffusion and external forces
the velocity at the new time step U' can be written as:

u1 = u + at (fl+1)_1 (_Cun + I(_G1Pn+1)). (2.32)

To cancel the effect of the that will blow up the eigenvalues of 1)'C, a
formulation based on a weighted average of the velocity u and the boundary velocity ub
could be applied. The weights can be taken equal to the fluid fraction of the uncut cell
(c1/(hh)) and the solid fraction (1 - 1/(hh)), resulting in
ç2n+l / / i
uM1 = _..L__ (u' + at
hh\ (Cu' + p(G1p'1)J) +
/ çn+l\
Iíi..xyJ
j

\
f
Ib n+1

This is a very rough but effective way to ensure stability and keeping the same time-step
restrictions as for uncut cells, namely the CFL condition for cells as large as uncut cells.
Moreover, when lf is small, the velocity u is forced to approach the boundary velocity u,
which is another safe property. A disadvantage of this approach is that some geometrical
information in cut cells may be lost. The velocity in cut cells may be adjusted too
much to the boundary velocity and high fluid velocities along the boundary can not be
simulated very well because of this 'smearing' of the boundary interface in the cut cells.
Applying this to a non-moving boundary case will also modify the velocities in cut cells
and eventually the fluid velocities will be forced to approach O in these cut cells. Since
2.9 Temporal discretisation

for non-moving boundaries this kind of modifying velocities is not necessary for stability
there must be a way in between that leaves the non-moving case as it is and applies a
small modification in the moving boundary case to ensure stability without smearing the
interface too much. These are the reasons why the above approach is not applied.
A better estimate of the eigenvalues of Ch(u, Ub) depending on the velocity u& is
required to achieve such a solution. As we have seen the eigenvalues are bounded by the
theorem of Gerschgorin and they are O(hub), where h are the spacings of the Cartesian
grid. Remember that the entries of Ch only consist of mass fluxes through the cell faces,
and that the mass fluxes satisfy the continuity equation (2.10). When the boundary
velocity is normal to the wall, the term l(ub flb) has its maximum value in absolute
sense, and with order of magnitude O(hu6). This has been shown in the example before
and will result in eigenvalues of Ch of indeed O(hub). However, the situation where
the boundary velocity is tangential to the wall, the value of l(ub flb) is equal to zero,
and therefore following the same reasoning as before, the other mass fluxes are relatively
small, and according to the last example they are O(u), which does not give rise to
additional measures to keep the time integration stable as has been stated before. So this
observation motivates a distinction in stabilising the time integration when the boundary
is moving normal to itself (maximal stabilisation required) or tangential to itself (no
stabilisation required) with a continuous (e.g. linear) transition. The determination of
the degree of stabilisation can be derived from the difference of the momentum control
volumes AJ at the new and the old time step. When the boundary is moving tangential
to itself the control volumes will stay the same in magnitude at the old and new time
level. When the boundary is moving normal to itself !M.rI will be maximal in absolute
sense. This induces a stabilising factor dependent on .AÇ. The following stabilisation
can be applied for example:

u1 Q'(1 + (um + a (+1)_1 (_Cum + (_G1pm+1))) +


çn+l(çn+l +
(I If D') n+ i
Ub (2.33)
Indeed, when = O for a certain control volume, no stabilisation is required and
the stabilising factor equals 1l( + 2jI)' = i which means that no stabilisation is
applied. Moreover, in the non-moving boundary case, /M» is always equal to zero, so
this stabilisation has no influence on the non-moving boundary case, which was stated
as one of the criteria. When is small and the boundary is moving normal to itself, so
IAf I » 1, (2.33) can be approximated by
n+i = 'IiIu + dt (_Cun + (_G1pm+1)) +
(I - çn+lJç»I)
The eigenvalues of dtI/MuíL'Cu now determine the stability. Since the boundary is
moving normal to itself, the value of JIlj approximates the flux ubhdt. The eigenvalues
of Cu are O(ubh) as explained before, which results in an estimate for the eigenvalues
of dtI1líI'Cu equal to O(dt/('ubhdt)ubh) = 0(1). So the eigenvalues are bounded for
small Hence a stabilising factor has been found that leaves the non-moving boundary
46 Chapter 2. Computational method

case as it is, that is stable for small 1l and that keeps iEhe smearing of the interface small.
Of course there will be other stabilising factors or methods that will also work well, but
in practice this stabilisation works satisfactory. Another stabilising factor that would
probably work is the factor 'max(l, 1ì')-'.

Figure 2.28: Situations where different rates of stabilisation are required. Boundary mov-
ing normal to itself where maximal stabilisation is required (left). Boundary moving
tangential to itself where no stabilisation is required (right).

Diffusion
Another restriction for the time step follows from the diffusion. Considering a simple
one-dimensional convection-diffusion equation with diffusion coefficient u, the diffusive
limit for the time step yields
2
at
2v
In uncut cells, where k = h, this limit will dominate the CFL-limit when u > 1í (see
also section 2.2.4). Clearly, cut cells where inkl becomes small are troublesome. Implicit
time integration could help here, but is more difficult to handle. Moreover, for the type of
flow in this thesis the effect of viscosity plays a very minor role and an accurate treatment
of the diffusion is of less importance. Therefore, artificial diffusion is added to the central
discretisation described earlier in this thesis, to guarantee a monotonic behaviour. Artifi-
cial diffusion is introduced with a magnitude of t1art = lull nk /2, which corresponds with
an upwind discretisation of convection. If 1/art dominates the real viscosity the diffusive
limit turns into
inkl2 Inkl2 i.!:t h
(2.34)
= kH72ki = II - lui
From (2.34) it follows that, using artificial diffusion corresponding with an upwind dis-
cretisation, the diffusive time step limit dominates the CFL limit h/lu I in cut cells because
of (2.34), but is in equilibrium with the CFL limit in uncut cells (here nk = h). If the
boundary is now treated as a 'staircase' boundary, i.e. the value of F is either i or O
in each cell and the velocity uk in equation (2.16) is now moved from the boundary to
the corresponding staggered grid location, the distance 72k becomes equal to h, then it
2.4 Behaviour of the pressure 47

is guaranteed that the diffusive limit is in equilibrium with the CFL limit. Since the
accuracy of the diffusive discretisation was already violated by the upwind discretisation,
this diffusive 'staircase' approach is acceptable and does not restrict the time step to be
smaller than the CFL limit. As already stated before, the diffusive terms do not restrict
the time step with respect to the stability of the time integration since they are discre-
tised in a 'staircase' manner When diffusive terms become important and require a more
accurate approach they can also produce some 'headaches' [97].

2.4 Behaviour of the pressure


Since the flow is incompressible, the density is assumed to be constant, thus the computed
pressure is not the thermodynamic pressure as it follows from an equation of state p =
p(p), but it is determined to an additional constant by zero divergence at the new time
step. This divergence-free constraint makes the pressure very sensitive to irregularities in
the flow as will be shown in the next section. This sensitivity will express itself mainly in
a spiky pressure signal in time, while the velocity shows a more smooth behaviour. These
spikes that last one time step, are mostly unphysical, however in some situations a spiky
behaviour may also be expected in reality, but no illusions are made to be able to capture
the physical spikes: The spiky behaviour in the simulations can be explained purely by
numerical irregularities. In a previous study of green-water loading, i.e. the simulation
of the behaviour of water that flows on the deck of a ship in high waves, these spikes
were undeniably present. An example of this spiky behaviour is shown below, where the
pressure at a certain point on a triangular structure is computed. To illustrate a green
Pressure
60

50-

40
'u
u.
4
30-
Lo

S
a
20-

10-

4 6 10
time (u)

Figure 2.29: Computed pressure at triangular structure.

water simulation and the complex behaviour of the flow that it inherits, a snapshot of
the simulation, including the triangular structure, is shown.

The phenomenon of the pressure spikes appeared to be a problem that can not be
attributed to one single reason. The main cause of the spikes can be found as a conse-
48 Chapter 2. Computational method

Figure 2.30: Snapshot of a green water simulation

quence of labels, free surface or geometry, that change from one time step to the next (in
combination with the divergence free constraint), but more reasons are present. Below an
overview of the origins of the spikes will be given and the possible remedies to eliminate
them.

2.4.1 Velocity boundary conditions for the free surface


The spiky pressure signals that are observed in the green-water simulations appear to be
more present with structures on the deck that cut diagonally through the Cartesian grid
than structures that are aligned with the Cartesian grid. The next figures show force
signals for a rectangular structure standing vertically on the deck, a triangular structure
and a cylindrical structure. Since the force is computed as the integral of the pressure
over the structure, the force shows similar spikes as the pressure. From these figures it is
FX

Figure 2.31: Computed horizontal loads on a rectangular structure (left), a triangular


structure (center) and a cylindrical structure respectively (right).

clear that the triangular structure suffers the most from a spiky force signal. A somewhat
lesser spiky signal but still obviously present is computed for the cylindrical structure.
The rectangular structure clearly shows a smoother signal, especially in the first half of
the simulation, during and directly after the impact of the water at the structure. The
2.4 Behaviour of the pressure 49

second bump, due to a hydrostatic pressure that results from the water that creeps up
the structure and then falls down, shows some spikes though. These results suggest that
the origin of the spikes must be found in the treatment of the cut cells that are present
in the green-water simulation for the triangular and cylindrical structure.

A simple two-dimensional test case has been designed to be able to observe the be-
haviour in cut cells in detail. A channel flow is simulated, with a sloping edge as drawn
in figure 2.32. Along the sloping edge, cut cells are created and the pressure at a point
along this edge is monitored. A preliminary test of this channel flow has been done
without a free surface and this test showed a smooth pressure signal in time at the slop-
ing edge. Thus, it is likely that the presence of a free surface is the cause of the spikes.
Therefore the test has been carried out with a free surface, flowing along the sloping edge.

-
Figure 2.32: Channel with a sloping edge to investigate the behaviour in cut cells.

The first test, with the same free-surface treatment as in the green-water simulations
shows a quite spiky pressure signal (figure 2.33), especially immediately after the free
surface passes the monitored point. After that, at more or less regular time intervals some
lower spikes are present with one exception at t 3.0 s. Since the pressure signal without

3.5

time (s

Figure 2.33: Pressure signal for a free-surface flow in a channel at a certain point at the
sloping edge of the channel.

a free surface is found to be smooth, the reason of the spiky signal is likely to be found in
the boundary conditions for the free surface, perhaps in combination with label changes
near the free surface. When having a more detailed look at the cut cells near the sloping
50 Chapter 2. Computational method

edge and the free surface (figure 2.34), the effect of the velocity boundary conditions
can be investigated. The bold circumlined S cell has one neighbouring E cell, thus one

FJE
boundary velocity is required at this S-E cell face. In the original method that was used

F!JE
FE
FB B
Figure 2.34: Detailed look at the cut cells and the free surface labels near the sloping edge
of the channel of the test case.

for the simulation of green water, the velocity boundary conditions for S-E combinations
were determined by the following strategy (for simplification only the two-dimensional
situations are considered). An S cell has one, two, three or four E neighbours. For every
S-E combination a boundary velocity must be computed. The basis of the computation of
the boundary velocities is the continuity equation, thus conservation of mass is demanded
in S cells. For an S cell with one E neighbour, all other velocities are known and the
S-E velocity can simply be computed by the continuity equation. When the S cell has
two E neighbours not opposite to each other, conservation of mass is applied in each
Cartesian direction. When the S cell has two E neighbours opposite to each other, the
S-E velocities are set equal to zero. With three E neighbours a combination of the former
two is applied and with four E neighbours all S-E velocities are set equal to zero.
Originally, these boundary conditions did not account for the presence of cut cells,
i.e. the velocity boundary conditions were computed as if there was no boundary cutting
through the cell, using edge apertures equal to one or zero. Consider the bold circumlined
S cell in figure 2.34, then the S-E velocity was computed from the other three momentum
velocities UIL, y'5 and ys and cell sizes h and h as

SE = u'° + j(v - (2.35)

Consequently, in these cut cells, conservation of mass is not exactly demanded. This
shortcoming can have an effect on the pressure when the S cell changes into an F cell.
After this change the velocity field in F cells will be made exactly divergence free taken
into account the boundary that cuts through the cell. This is done by the pressure
Poisson equation, where the pressure is adjusted to create a divergence free velocity field.
If the velocity field in that cell was not divergence free at the old time step, which is the
case in the concerned S cell, the pressure has to 'work' to achieve conservation of mass
in the newly created F cell. This 'work' will manifest itself in a spike in the pressure
2.4 Behaviour of the pressure 51

signal. The regularity of the spiky signal in figure 2.33 is caused by the free surface that
passes the cell edges further away in the uniform grid. This causes a high pressure spike
in that specific grid cell, but is also noticed in all other cells, since every pressure pulse
is traveling with infinite velocity through the domain. When cut cells are accounted for
in the computation of the velocity boundary conditions, the pressure signal should look
much smoother intuitively, but things are not that easy as will be seen later. However,
the boundary conditions can be adjusted to account for the cut cells resulting in the
following discretisation for the bold circumlined S cell in figure 2.34.

uSE = +
h(Av'Av8) (2.36)

The resulting pressure signal is shown in figure 2.35. In comparison with figure 2.33, the

e
o-

e
5',
(n
e
o.

2.5 3

time (s)

Figure 2.35: Pressure signal after accounting for mass conservation in cut cells in the
computation for the velocity boundary conditions for the free surface.

spikes just after the free surface has passed the monitored point are gone, also between
t = 2 s and t = 3 s less spikes are observed. However some spikes are still present, with
a few exceptional high spikes.
It turns out applying exact mass conservation in S cells has some positive (a more
smooth signal iii general) but also some negative effects on the pressure (the exceptional
high spikes). The reason for this has to be found in small cut cells near the free surface.
Since the time-step limit is based on the CFL condition for an uncut cell, it is possible
that the free surface crosses a small empty cell completely within one time step. The
consequence is that the free-surface label of this cell changes from E to F in one time
step (figure 2.36). In E cells no conditions are satisfied with respect to mass conservation,
thus the same problem as before with the non-mass conservative S cells arises, which will
result in an unphysical pressure spike. Using the fact that the free surface is displaced
before the new velocity and pressure field is computed, this problem can be resolved to a
high degree. After the new free-surface labels are computed, the E-F transitions can be
detected. If an E-F transition has occurred, the newly created F cell is not divergence
free in general. The pressure Poisson equation forces the F cell to be divergence free in
the next time step, thus a pressure spike is created as explained before. The idea is to
52 Chapter 2. Computational method

S E E E F E E

F S E E F S E E
F E F F S

B F F B

Figure 2.36: Situation where a small E cell at time step n (left) becomes an F celi at
time step n+1 (right).

make the newly created F cell divergence free before the next velocity and pressure field
is computed. This can be done by adjusting the F-S velocities that are created in such a
way, that mass conservation holds in the 'new' F cell. Tri figure 2.36 the bold circumlined
F-S combination is adjusted. If the velocity boundary conditions routine is called after
this action, the S cells will be made divergence free automatically by computing the
S-E velocities as before. Note that only F-S velocities can be adjusted, and not F-F
velocities, since then the other F cell, in the example the F cell to the left of the new
F cell does not conserve mass anymore. This can not be repaired because there are no
velocity boundary conditions left in this cell.
In figure 2.37 the pressure signal is shown after detecting E - F changes and re-
pairing mass conservation by adjusting F-S velocities. A significant improvement is seen

2.5 3
time (s)

Figure 2.37: Pressure signal after repairing mass conservation in situations where E cells
become F cells in one time step.

compared with figure 2.35. Almost all spikes appear to be removed, only one small spike
is still present at t 1.5. In this kind of flow, where the free surface moves along a wall,
an F-S velocity can almost always be found to repair mass conservation. However, situa-
2.4 Behaviour of the pressure 53

tions exist where no F-S velocities are present, e.g. when a free surfaces smashes against
a wall, or when a hole in the fluid collapses (figure 2.38 and 2.39, respectively). When
this is the case, a pressure spike can not be prevented, but this may sometimes not be as
unphysically as it seems to be. A physical collapsing air bubble for instance, which can be
represented by figure 2.39, may also show an impact pressure. Also air entrapment may
occur when a free surface smashes against a wall, resulting in a physically spiky pressure
signal (e.g. [69,70,88, 102, 104]). So it is likely that when such spikes are observed in the
simulations, physics also shows a spiky behaviour. However, the qualitative behaviour of
these spikes can only be investigated when entrapped air flow is included in the model [75].

FSEA FF
-0 FpJg
B

Figure 2.38: Situation ofa free surface smashing against a wall, where the bold czrcumlined
E cell (left) has become an F cell at the new time step (right). No F-S velocity is present
to repair mass conservation, thus a pressure spike can not be prevented.

F S F F F F

sills -0 FcF
F F F F F

Figure 2.39: Situation of a collapsing bubble, mass conservation does not hold in E cell
(left), resulting in a pressure spike at new time step since mass conservation can not be
repaired in the new label configuration (right).

Since in the test channel flow, the free surface is moving quite smoothly up the sloping
wall, it is unlikely that the small spike that is left in figure 2.37 is due to a situation as
described above. A more plausible explanation of this spike can be found when considering
a situation as in figure 2.36. Adjusting the F-S velocity will almost remove the spike,
but the mass conservative boundary conditions also inherit a disadvantage. Investigating
the left situation in figure 2.36 and in particular the S-E combination at the bottom,
the S-E velocity is the only boundary condition that is required in the concerned S cell.
54 Chapter . Computational method

This boundary condition is computed such, that mass conservation is satisfied in the S
cell. The edge aperture of the S-E cell face is now important. Since a division through
this aperture is performed (compare 2.36), small values of this aperture can make the
S-E boundary condition explode (figure 2.40). This unphysically high boundary velocity

Figure 2.40: Small S-E cell-face aperture, resulting in an umphysically high velocity boun-
dary condition to satisfy mass conservation.

appears in the momentum equations and can thus produce an irregularity. The reparation
of mass conservation in the new F cell can not prevent a small spike due to this small
aperture. This situation occurs in the test channel flow, looking at figure 2.35 and 2.37,
the high peak due to the E-F situation at t 1.6s is replaced by the small spike due to
a small S-E cell-face aperture.
Applying a limiter to the S-E boundary condition could be an option in order to
control its magnitude. Mass conservation is not satisfied anymore then, thus still some
irregularities can be expected. Furthermore, the choice of the limiter is not unique.
Different situations may also need different limiters to optimize the pressure signal. For
the test situation, in figure 2.41 the pressure signals are shown for the simulation using
three different limiters. The limiter LA is chosen in a way that conservation of mass
still holds when the S-E cell-face aperture A > LA, if A < LA the boundary condition
is computed to conserve mass as before as if A = LA. As was expected, still some

time(s) time (s) time (s)

Figure 2.41: Pressure signals for test channel flow using three different limiters for the
computation of the velocity boundary conditions. LA = 0.05 (left), LA = 0.08 (center),
LA = 0.1 (right).

irregularities are present at t 1.6 but they are a bit smaller than without a limiter
(compare figure 2.37). For LA = 0.1 another spike appears at t 1.5. Another option to
prevent the S-E boundary condition from exploding, is to limit the apertures globally.
This means that the geometry apertures are adjusted in a way that the value of the
2.4 Behaviour of the pressure 55

apertures has a lower limit These apertures are then used throughout the complete
simulation. The geometry in the cells that have been limited will become a little different.
The geometry in figure 2.36 becomes a little indented, like in figure 2.42. Here the
indentation is somewhat exaggerated. The indentation does not have a big influence; only

s\E E E

F\S E E

F\E,,IÍ
VEB
Figure 2.42: Indented geometry, due to limiting of the apertures (somewhat exaggerated).

in the neighbourhoud of the indentation some small pressure differences are observed. The
simulation done with this 'new' geometry gives a result for the pressure signal as shown
in figure 2.43. As can be seen, the influence of the indented geometry on the pressure

Figure 2.43: Pressure signal for simulation with a global limiter on the apertures of the
geometry.

signal is negligible. Furthermore, no more spikes are observed. Things will become more
difficult when using moving bodies (see 2.4.3), since then the apertures are changing in
time, while without moving bodies the indented geometry is used throughout the whole
simulation. However, the green water simulations were carried out with a fixed geometry,
hence an attempt can be made to 'mow the grass' in the simulation by applying above
measures.
Thus recapitulating, mass conservation is demanded in S cells, E-F transitions are
detected and repaired if possible and a global limiter is applied to the geometry apertures.
56 Chapter 2. Computational method

The results of the pressure signals after applying these measures are given in figure 2.44.
Comparing these results with figure 2.31, the spiky behaviour has completely disappeared.
F0 F3 FO
6000 6000

7000
2000

6000

4000
5000
z z

r30
3000
2000

2000

000
1000

00 2 4 6 10 2 4 6 6 lO
woo o) 5mo (o)

Figure 2.44: Computed horizontal loads on a rectangular structure (left), a triangular


structure (center) and a cylindrical structure respectively (right).

Thus, also in three dimensions with very complex flow behaviour the measures mentioned
above appear to be sufficient for generating an acceptable smooth pressure signal.

Energy considerations of boundary conditions


Unfortunately, though producing nice pressure signals, a troublesome side effect shows
up. It sometimes happens that simulations break down, which can be very inconvenient.
Most of the simulations will pass the test, but sometimes very high velocities are observed
forcing the time step to be very small, which can eventually result in a break down of
the simulation. To understand what is going on, the cells with the small apertures have
to be investigated again. Despite the introduced limiter, it still can happen that unphys-
ically high S-E velocities are produced. Unphysically, since these velocities are assumed
to represent velocities in the 'air', which are boundary conditions for the momentum
equations in the fluid. High 'air' velocities will therefore show up as high fluid velocities
in the momentum equation, and they will put more energy in the fluid, which can have
an accumulating effect. Especially, when an S cell almost contains no fluid, a high S-E
boundary condition, due to a small aperture (compare figure 2.40), even when a limiter
is used, is far from physical. Looking at figure 2.40, the S-E boundary condition will
unphysically 'pull' the fluid to the right. The nice matrix properties generated by the
discrete conservation of mass and the discrete conservation of the other continuous op-
erator properties that guarantees a simulation without a free surface to conserve energy,
do not appear to guarantee a free-surface simulation to be stable. To understand this,
the energy of the complete system including velocities and boundary condition velocities
has to be considered. So the evolution of the complete velocity vector (u, u13)T has to
be investigated, where u represents the internal velocities and u1 the velocity boundary
conditions of the free surface. Since conservation of mass is satisfied in the free-surface
boundary conditions, all the nice matrix properties are preserved and the kinetic energy
induced by the velocity vector (u, u12)T will be constant. The gravity term can put more
energy in the discrete system via the free surface. More important however is that the
2.4 Behaviour of the pressure 57

free-surface boundary velocities u15 are controlled from outside and therefore can change
the energy of the system, even if conservation of mass is satisfied. Suppose for example
that an S cell has two opposite E neighbours. If all other velocities at the cell faces of
the S cell are equal to zero, these two S-E boundary conditions must be chosen equal
to each other to satisfy conservation of mass. Hence they can be chosen arbitrarily high
but equal to each other, putting more energy in the system since they appear in the
momentum equations. High velocities due to small apertures may put energy into the
system in the same manner. Thus the discretisation itself does not induce an energy
change, but the choice of the boundary conditions does. From this point of view, de-
manding mass conservation in S cells can work out completely wrong, though producing
smooth pressure signals. Therefore it may be better to apply the velocity boundary con-
ditions in such a way, that energy conservation is violated as little as possible, instead of
demanding divergence free boundary conditions. This will inevitably result in some ir-
regularities in the pressure signal. It seems to become a trade-off between robustness and
smooth pressure signals. From energy considerations, it would be perfect if the velocity
boundary conditions can be constructed in such a way, that energy is conserved exactly.
However, this is very difficult, since many aspects have to be taken into consideration
near the free surface. Not only the problem of the small apertures has to be taken into
account, also the change of free-surface apertures F5 and the change of free-surface labels
so that boundary velocities become internal velocities and vice versa, not to mention the
boundary conditions for the pressure in S cells and the measures that have to be taken
for the displacement of the free surface. No attempt has been made to solve this problem
completely, but it is clear that the principle of energy conservation should not be ignored.

Selected approach
In this study a rather safe solution is used, based on a constant extrapolation of the
internal velocities. Thus S-E boundary velocities are set equal to an internal velocity,
which choice depends on the number of E neighbours and the direction of the free surface.
First, if the concerned S cell has only one E neighbour, the S-E velocity is set equal to
the opposite F-S velocity (figure 2.45). When an S cell has more than one neighbouring

E E E

s-8s 11

£1,

F F F

Figure 2.45: S cell with one E neighbour. The S-E velocity i is set equal to the opposite
internal F-S velocity 1'.
58 Chapter . Computational method

E cell, there are a few possibilities. In two dimensions if an S cell has four E neighbours it
represents a small drop and then the S-E velocities are set equal to the old velocities that
were present at the cell faces, if applicable in combination with gravity. Some care must
be taken here to make sure that these drops do not blow up the computation. Further
they have no big influence on the flow, so the choice of these boundary conditions does
not matter much. When the S cell has two or three E neighbours, the normal direction
of the free surface is estimated with help from the F8 values. Then the S-E boundary
conditions are determined by setting these velocities equal to internal velocities just below
the waterline (figure 2.46). Another property of these boundary conditions is that they

E E E
11

Efl3E
S-F F S 2'F 3'8

Figure 2.46: Left: S cell with two E neighbours. The S-E velocities i and 2 are set equal
to the internal F-S velocities 1' and 2' respectively. Right: S cell with three E neighbours.
Now the S-E velocities 1, 2 and 3 are set equal to the internal F-S velocities 1', 2' and
3'.

provide a smooth behaviour of the velocities near the free surface, since they are always
related to an internal velocity. For the simulation of steep waves for example, where the
boundary conditions are important to describe the wave accurately, a smooth behaviour
of these boundary conditions appears to be advantageous [59].
With these new boundary conditions, based on a constant extrapolation of internal
velocities, the green-water simulations can be carried out. The results for the force
signals are shown in figure 2.47. The solution for the removal of the spikes due to E-F
transitions is still applied, and it seems that this action is sufficient to remove most of
the spikes. Compared to figure 2.44 the signals are a bit more noisy - remind that mass
conservation is sacrificed in S cells - but still acceptable and also qualitatively the signals
are a bit different. In ongoing research which is not treated in this thesis it turns out that
compared to experimental data sometimes even better results are achieved with these
new boundary conditions. However, the most important property is that these boundary
conditions result in a much more robust method.
Concluding, free-surface boundary conditions are very important and to a high de-
gree determine the flow solution. Therefore, they have to be treated very carefully since
they can even make a simulation break down as well. Sometimes a trade-off has to be
made between robustness and smooth pressure signals. For very complex flow behaviour
like green-water loading, robustness will make life more easy. To eliminate all problems
regarding the velocity boundary conditions, the inclusion of a two-phase model is prob-
2.4 Behaviour of the pressure 59

FX F2 F2

2 4 6
()

Figure 2.47: Computed horizontal loads on a rectangular structure (left), a triangular


structure (center) and a cylindrical structure respectively (right) with the new velocity
boundary conditions for the free surface.

ably required. This is not treated in this study, but another study [8] has shown some
promising results.

2.4.2 Pressure boundary conditions at the free surface


The boundary conditions for the pressure are applied in the S cells, and they are normally
determined by an interpolation with an F cell that is chosen by estimating the normal of
the free surface (figure 2.23). Usually, this results in a smooth behaviour of the pressure.
There are situations however, that a spiky pressure behaviour is observed due to the
pressure boundary conditions. Consider a simulation of a dambreak (figure 2.48) where
the free-surface front is quite thin, only covering one grid cell in vertical direction. The

E E E E E E E E E E

S E E E E E E E

F F F E E E E E E

F F F F S E E E E E

F F F F F S E E E

Figure 2.48: Typical free-surface distribution for a dambreak simulation. The free-surface
front only covers one cell in vertical direction.

pressure signal of the bold circumlined S cell is shown below in figure 2.49. A spiky
pressure signal is observed, which can be explained as follows. In the S cell concerned, a
boundary condition for the pressure has to be applied. Based on the orientation of the
free surface, which is chosen to be either horizontal or vertical, depending on surrounding
60 Chapter 2. Computational method

0.5
04 0.7

Figure 2.49: Pressure signal in a grid cell just above the bottom for a dambreak sirnulatzon,
e.g. the bold circumlined cell in figure 2.48.

F5 values, an F cell is chosen used for interpolation. When the S cell is oniy filled with
a small portion of fluid, the orientation of the free surface will be estimated as vertically.
Then the F cell to the left of this cell will be chosen to interpolate with. However, when
the S cell becomes fuller, and when the cell to the right of this S cell also becomes an
S cell, the orientation switches to horizontal. In this case an F cell is searched below
the S cell to interpolate with. Since there is no F cell in this direction, but a boundary
cell, no interpolation is performed. Usually, the pressure in such an S cell is set to the
atmospheric pressure with eventually a correction for gravity. This value appears to be
quite different from the interpolated value at the time step before and therefore produces
a spike. The spike just after t = 0.55 in figure 2.49 occurs when the direction in the
concerned S cell switches from vertical to horizontal. The smaller spikes thereafter are
caused by the same procedure when the free surface passes the cells to the right of this
cell. An attempt has been made to remove the spike in the particular S cell by always
interpolating with a neighbouring F cell if present, even if in the normal direction of the
free surface no F cell is found. So in this case, for the bold circumlined cell in figure 2.48,
the pressure in the S cell is interpolated with the F cell to the left. The new pressure
signal is shown in figure 2.50. Clearly, the big spike is gone, only the smaller spikes
are still present but this is because the new measure has only been taken ad hoc in one
particular S cell. The sensitivity of the pressure interpolation with respect to pressure
spikes is visible now. In the future, to minimize this kind of spikes, when no F cell is
found in the desired direction, the interpolation is performed with an F cell, if present,
in another direction. The influence of the pressure interpolation can also be seen clearly
in the first part of the pressure signal, when the free surface enters the S cell. Since the
F5 value is very small, the free surface location is expected to be very close to the left
cell face with vertical orientation. To achieve a pressure Po = O at the free surface by
interpolation with the adjacent F cell, the pressure in the cell center has to be negative.
That is the reason why the pressure signal starts negative when the free surface enters
the cell and slowly increases when the cell becomes fuller.
2.4 Behaviour of the pressure 61

1.5

C,

0.5
04 0.55 0.7
time (st

Figure 2.50: Pressure signal of the dambreak simulation, where the interpolation in the
particular cell is always done with the adjacent F cell.

2.4.3 Moving bodies with sharp corners


The origin of the spikes that have been treated until now lies in the boundary conditions
for the free surface. Another spiky behaviour is observed when objects are moving through
the grid, especially when the surface of the object is not smooth, i.e. when it possesses
sharp edges. Since the only geometrical information that is provided in a Cartesian grid
cell are the volume apertures and the edge apertures, and the discretisation is based
on a linear reconstruction, the method does not detect a sharp corner in the geometry
within the cell. This lack of geometrical information is expressed in irregularities in the
pressure when the sharp corner crosses a cell face. A typical example of a moving body
with sharp edges is a cube, or in two dimensions a square, traveling through the fluid. A
two-dimensional test case has been designed to investigate the behaviour of the pressure
when a square is moving horizontally through the fluid (figure 2.51). To avoid free-surface
effects, the test case has been performed without a free surface.

Figure 2.51: Test case for investigating numerical spikes due to moving bodies with sharp
edges, like a square.

A typical pressure signal for a certain monitored position in the fluid is shown below
in figure 2.52. To illustrate the origin of such spikes, consider a one-dimensional situation
62 Chapter 2. Computational method

20

15-

10-

1111
Ii

0.5 1.5 2 2.5 3 3.5 4


time (s)

Figure 2.52: Spiky pressure signalfor a certain point in the fluid in the designed test case.

(figure 2.53). An object with a sharp corner is moving through a channel where only
one cell is used in vertical direction. The vertical location of the sharp corner is situated
halfway the vertical cell face. For simplicity, the convective and diffusive contributions
are omitted in this illustration. The velocity of the moving body is assumed to have a
constant value ub = 2, while in the channel the fluid velocity has a constant velocity
u = 1. Because of the divergence constraint, the velocity at cell faces that are covered
half by the fluid and half by the moving body has to be equal to zero. Looking at the
continuity equation in the left situation of figure 2.53,

nc - - = O. (2.37)

When ub = 2 and u' = 1, uW indeed has to equal zero to satisfy (2.37). When the front
of the moving body does not cross a cell face, all cell face velocities keep the same value,
therefore the pressure gradient Ghp is equal to zero according to the simplified version of
equation (2.28)
dt n+1
= h P (2.38)

In figure 2.53, the front of the moving body crosses the central cell face (c) from time step
n to time step n + 1. The fluid velocity at this cell face is then forced by the divergence
constraint to change from u = i to u = O. From (2.38) it follows that a pressure gradient
across cell face c is created:
Ç1)+lp

- dt
According to (2.17), Ghp = Ah(p' - p1), with A = , the difference pr - p1 becomes
) 012+1
f
(
tP
r
P,1\n+1..._- hdt (2.39)
2.4 Behaviour of the pressure 63

+1

=I u=i

Figure 2.53: One-dimensional illustration of a situation where a sharp corner crosses a


cell face from time step n to time step n + 1.

A pressure spike is thus created here, caused by the edge aperture 'jump' from time step
n to n + 1. Note that using a smaller time step is making things worse: When the time
step is halved, the magnitude of the spike is doubled. The product of time step and pT_pl
will be constant, when the same values of p, ì1 and Ah are used. This is illustrated
in figure 2.54, examining one pressure spike for different time steps in the test case of
figure 2.51. These spikes are of course far from physical and are a consequence of the

14
- cit = 0.001
- dt = 0.002
12- - dt = 0.004

10-

t97 0.975 0.98 0.985 0.99 0.995 1 1.005 1.01


time (s)

Figure 2.54: Behaviour of a pressure spike using different time steps.

loss of geometrical information of the sharp corners within a cell. The aperture 'jump'
in combination with the divergence constraint that forces the fluid to move according
to the moving body is the cause of these spikes. The pressure is determined in such
a way that the flow is divergence free, and can be seen as a potential function that
forces this constraint. An analogon with the fictitious domain approach of Glowinski
(e.g. [27, 28, 41J) is recognized here, where Lagrange multipliers are used to match the
fluid velocities with the rigid body velocities. In the test channel case with the square,
it is clear that these spikes are not physical. Consider the computed pressure p(x, y, z),
and the physical pressure pphys(X, y, z). Write the computed pressure field as the sum of
the physical pressure and a potential (x, y, z), thus

p(x, y, z) = Pphya(X, y, z) + q(x, y, z).


64 Chapter 2. Computational method

This partitioning of the computed pressure enables a mathematical filtering of the pres-
sure spikes, assuming that the spike is contained in the potential Ç5 and Pphys is assumed to
be smooth. In time intervals when the computed pressure behaves smoothly and equals
the true physical pressure, the value of Ç5 equals zero. A numerical filtering of these
pressure spikes in this way is not that easy, since the partitioning of the physical and the
computed pressure is not always clear. Moreover, the crossing of the sharp corners with
cell faces has to be detected in a certain way. Since some geometrical information of the
sharp corners is lost at cell level, it is impossible to compute the exact physical pressure,
but an approximation will be possible. Consider again the one-dimensional test case, the
pressure gradient equals zero in time intervals when the sharp corner does not cross a cell
face. Assume then the computed pressure field is equal to zero: p(x) = O, satisfying a zero
pressure gradient. In these time intervals let the physical pressure Pphys be approximated
by this pressure, thus Ç5(x) = O and Pphys(Z) = O. When the edge aperture switches from
i to , the physical pressure should stay equal to zero. Therefore, the spike (2.39) should
be contained in Ç5:

(pr pt)fl+l r I
= (Ph5 - Pphysì
n+1
+( = + 2 hdt
To accomplish this numerically, an attempt can be made to split off the physical pressure
and the spike by a numerical trick. First, the crossing of the sharp corner with a cell
has to be detected. Second, the computation of a temporary velocity and pressure field
is done using the new apertures as usual, but only using old apertures in cells where
the sharp corners are crossing a cell face. The aperture 'jump' is now prevented and the
pressure Ptemp that is computed does not contain the spike as consequence of this jump.
The smooth physical pressure Pphy can be approximated by the so computed pressure
field Ptemp, which is also smooth. After this, the velocity field at the new time step with all
apertures at the new time step has to be made divergence free. This is done by computing
the field as usual, thus the pressure then computed contains the spike Ç5 and the physical
pressure Pphys Note that now two Poisson equations have to be solved in one time step,
one to compute Ptemp and one to compute the sum (not the individual terms) Pphys + Ç5.
This procedure is illustrated in figure 2.55. Applying this procedure to the test case in
figure 2.51, the pressure signal Piemp can be computed and this is shown in figure 2.56.
The spiky behaviour is no longer present now, though a sawtooth shaped irregularity is
observed, but this is caused by another property of the discretisation which is treated in
the next subsection. This irregularity is almost negligible in comparison with the former
spiky behaviour.
In the fictitious domain method of Glowinski, the body is filled with fluid and the
Lagrange multipliers that are distributed over the body act as a body force that forces
the velocities inside the body to equal the rigid body velocity. In the present method, the
rigid body is not filled with fluid, but the rigid body velocities are applied at the boundary
of the body, acting as boundary conditions. The pressure has to enforce the divergence
free constraint hereafter. In analogon with the method of Glowinski the pressure can
be seen as a Lagrange multiplier forcing conservation of mass including the effect of the
moving body. In the present method this pressure is split in a physical pressure p and a
potential Ç5 that 'absorbs' unphysical spikes.
24 Behaviour of the pressure 65

n temp n+1

Figure 2.55: Procedure for splitting of the physical pressure and the spike pressure. Ge-
ometry at time step n with pressure pfl (left). 'Temporary' geometry with smooth pressure
Ptemp Pphys (center). Geometry at time step n + i with spiky pressure p72+l Pphs +

time (s)

Figure 2.56: Pressure signal Ptemp in designed test case after applying above procedure for
smoothing the pressure signal (compare figure 2.52).

2.4.4 Pressure positioning


In the finite volume method, the exact positioning of the pressure is not defined, but
it is assumed that the pressure in a Cartesian cell has a constant value throughout the
complete cell. Since the pressure term is discretised by a boundary integral, namely the
integral of the boundary of the momentum control volume of pm, no pressure gradient is
required. In the absence of cut cells, the finite-volume discretisation of this pressure term
is identical to a finite-difference discretisation where the pressure gradient is multiplied
by the size of the momentum control volume. The pressure gradient in each Cartesian
direction is computed by a pressure difference p' - p' divided by a distance h. This
distance is derived from a cell-centered positioning of the pressure. Considering the
pressure discretisation in x-direction over the control volume V in figure 2.57, the finite-
66 Chapter . Computational method

¿p! V f pr
V
4

Figure 2.57: Momentum control volume V with cell centered pressures p1 and p' to illus-
trate the discretisation of the pressure term.

volume discretisation of the pressure term becomes

dôV = hy(pr - pl).


fP'X
0V

Using the cell-centered positioning of the pressure induces a distance while the size of
the momentum control volume V is equal to hh5, then the finite-difference discretisation
of the pressure term reads

f?V
x
= hh' - 1
= h,(pT -
V

On the other hand, writing the discretisation of the pressure term in finite-difference form
and demanding this has to be equal to the finite-volume formulation, an interpretation
of the pressure discretisation based on a certain positioning of the pressure in a cell
can be derived. In the absence of cut cells this turns out to be the above cell-centered
positioning. This can also be done for cut cells, where it turns out that the pressure will in
general not be positioned in the center of the cell. The general form of the finite-volume
discretisation in x-direction of the pressure term over a momentum control volume V
with edge aperture A and h and as in figure 2.57 and preserving the symmetry
property V = V., is equal to (see (2.17))

fPr döV = A h (pr - (2.40)

The finite-difference discretisation is equal to


pr p1 (Fhh +Fhhy)(pT l)
I av
fOx h 2h
(2.41)
V

The interpretation of (2.40), based on (2.41), is obtaind by setting (2.40) equal to (2.41)
and calculating h. Hence,
h= (2.42)
2Ah
2. Behaviour of the pressure 67

Therefore, for cut cells, the positioning of the pressures pr and p1 regarding the momen-
tum control volumes, can be seen with a distance h between each other. This distance h
is the consequence of the discretisation of the pressure term, which is determined by the
discretisation of the divergence and preserving the symmetry property V = V., and
the choice of the momentum control volume.

The effect on the pressure due to the above discretisation choices will be investigated
by a simple one-dimensional test case (figure 2.58). The front of a moving body is
accelerated through a channel with 0.1 rn/s2. All cell dimensions are chosen equal to

Figure 2.58: One-dimensional testcase to investigate the effect of the discretisation on the
pressure. In this case the moving body front is parallel to the cell faces.
i m, so h = h5 = i m, thus ff = i m2. Due to this acceleration, a pressure gradient
will arise in the fluid. The divergence constraint fixes all velocities, and they will all be
equal to each other, so no convective and diffusive contributions are present. When the
density is scaled to one, the following holds:

= Vp = 0.1.
Since
Vp= pr - pt
the pressure difference in a fluid cell in the absence of the moving body where ii = 1 will be
equal to 1. Near the moving body, the value of h depends on the edge aperture A and the
control volume . The pressure difference pT Pt will be computed so that Vp = 0.1.
For a moving body in figure 2.58, the edge aperture stays equal to 1, while the control
volume decreases from i to 0.5. The value of h then varies also according to (2.42) from
i to 0.5. The pressure p1 can therefore always be thought of situated in the center of the
control volume part in the left cell. The pressure signal in this cell is shown in figure 2.59.
As can be seen in the pressure signal, when the moving-body front has not reached the
considered cell, the pressure ptis equal to 1. The pressure pT will be equal to 0.9 because
of Vp = 0.1. When the front is moving through the cell, the pressure decreases linearly
to 0.95, where h = 0.5, which is correct when the pressure is situated in the center of the
left part of the control volume (since then Vp (pr - pt)/h = (0.9 - 0.95)/0.5 = 0.1)
After the front has crossed the cell face, the cell has become a boundary cell where
the pressure is set equal to zero.
Sofar, nothing strange is going on. Things become different when the front of the
moving body is not parallel with the cell faces, but is cutting diagonally through the
68 Chapter 2. Computational method

1.2

0.8
s.
0.6
s
0.4

0.2

00 5 10 15 20 25
lime(s)

Figure 2.59: Pressure signal p1 in one-dimensional test case in figure 2.58.

grid, as in figure 2.60. The pressure signal for pt that is observed is shown in figure 2.61.

Figure 2.60: One-dimensional test case to investigate the effect of the discretisation on the
pressure. In this case the moving body front is cutting diagonally through the grid.

The pressure signal is still equal to i when the diagonal front has not reached the left
cell. When it enters the cell, the pressure slightly decreases, but then increases to show
a peak whereafter it becomes a boundary cell with pressure p = 0. The reason of this
spike becomes clear when the evolution of h is considered. When the front enters cell
I, the momentum control volume decreases in magnitude, while the aperture A stays
equal to i for a while, therefore h decreases according to (2.42). As in the previous case,
the consequence is that the pressure decreases. When the front crosses cell face c, A
decreases more rapidly than the momentum control volume decreases, thus h will increase
with a maximum when A admits its smallest value. The pressure follows this evolution
of h and thus increases with a maximum where h admits its maximum. In these cases the
edge aperture can become very small, say A = e, while the momentum control volume
keeps order of magnitude O(hh5). Then the distance h can become uuphysically big:

The position of the pressure pt is now definitely not situated in the cell center but some-
where far to the left of it, consequently the pressure pt becomes unphysically high. That
is the reason of the pressure spike in figure 2.61. To obtain a more physical pressure, one
could try to interpolate pt and p? to the center of the cell, but this can become compli-
cated when two- or three-dimensional situations of these kind are interfering. Typically,
2.4 Behaviour of the pressure 69

2.5

0.1.5

0.5

00 10 15 20 25
timo (s)

Figure 2.61: Pressure signal p1 in one-dimensional test case in figure 2.60.

these spikes are present when the geometry is cutting diagonally through the grid which
results in situations where very small cells with small edge apertures lie next to cells of
magnitude of the order of the uncut cell sizes like in figure 2.62. The pressure in the
small cell gets an unphysically high value, due to its virtual location according to h, that
can be interpreted by the discretisation of the pressure term.
Since this spiky behaviour depends on the magnitude of the momentum control volumes

Figure 2.62: Small cell lying next to a cell of normal size, resulting in an unphysically
high pressure in the small cell.

and the edge apertures, not oniy the pressure signal in time is spiky due to the moving
geometries, but also the pressure signal in space is spiky even when the geometry is not
moving. An example of such a pressure field is shown in figure 2.63, for a channel flow
as in figure 2.32, where small edge apertures are present due to the sloping edge. These
kind of spikes are very difficult to remove, since they are a consequence of the discretisa-
tion choices. They can be removed by other choices of the control volumes of apertures,
but then important information of the geometry is likely to be lost. An interpolation
procedure to find more physical pressures in the cell centers could be possible, but is not
easy in more dimensions. For the moment, no measures are taken, since it appears not
to influence the flow behaviour too much, and for the computation of forces on such a
wall the contribution of these high pressures can be neglected because they only work on
a very small area.
70 Chapter 2. Computational method

25

Figure 2.63: Pressure field in space for a channel flow as in figure 2.32.

2.4.5 Other possible reasons for spikes


The most important reasons and remedies for pressure spikes have been discussed above,
but still a spiky behaviour is observed sometimes. Especially a moving body in combi-
nation with a free surface is very sensitive for pressure variations. When a free surface
smashes against a moving body a spiky pressure behaviour is commonly computed. First,
most of the E-F situations that cause spikes, explained in subsection 2.4.1, can not be
repaired. Secondly, when physics is violent, e.g. when the body is moving in almost
opposite direction of the fluid motion, and if the fluid moves rapidly along the body,
momentum fluid velocities change in boundary conditions within one time step which are
totally different from each other. These boundary velocities do appear in the discreti-
sation of other momentum velocities, creating discontinuities in the momentum velocity
field, thus creating discontinuities in the pressure. Apparently, it is very difficult to ban
all pressure spikes. An interface smearing over a few cells may help, as is done in fictitious
domain methods and immersed boundary methods for example. In the fictitious domain
method of Glowinski [27,28], the moving body is filled with fluid and Lagrange multipli-
ers are distributed over the boundary of the body to force the rigid body velocities in the
body. The complete moving body is now included in the computational domain, so also a
pressure field is computed inside the body. The influence of the boundary is transmitted
to the fluid through the Lagrange multipliers. In immersed boundary methods, origi-
nally by Peskin [71], the material properties on both sides of the interface are distributed
smoothly over a transition zone of a few grid cells. Interpolation of the velocity is per-
formed to acquire velocities at interfacial points and the interface force is spread to the
nearby grid points by a discrete Delta function. In such methods, the detailed informa-
tion of the boundary is thus not discretised as a sharp entity, but as a somewhat diffused
interface. Therefore, they are not very well suited for convection-dominated flows.
.5 Coupled motion between moving bodies and fluid 71

Interface smearing is not an option in the present method, since in the kind of ap-
plications the method is meant to simulate, high velocities along the boundaries play an
important role, thus the flow equations must be discretised near the boundary based on
a sharp interface. The cut-cell method of Udaykumar et al. [94] and Ye et al. [107] treats
the boundary as a true sharp interface, but differs from the method in this thesis, since
it uses a non-staggered approach, including an interpolation procedure for the variables
and reshaping of cells near the boundary. Furthermore, no free surface is included in this
model. With respect to the method in this thesis, it seems that the price that is paid to
keep a sharp interface including sharp corners and a sharp free surface, is a less smooth
pressure signal in time.

2.5 Coupled motion between moving bodies and fluid


Until now, it has been assumed that the motion of the moving bodies was prescribed,
thus the fluid itself did not have any effect on the motion of the body. In practice, of
course, the motion of the body will be affected by the fluid motion. In this section a first
attempt is made to couple the motion of the fluid, including the free surface, with the
motion of the moving bodies.

2.5.1 Basic idea


In general, the forces exerted on a moving body in a viscous fluid exist of pressure forces
and viscous drag forces. The viscous drag forces are neglected in this study, since they
do not play an important role. Theoretically, the acceleration A of a moving body with
inertia matrix M can be computed by Newton's second law, thus
A
'1body - Al E'
IVI (2.43)

Here the acceleration A is the acceleration vector (, , , w2,, ) and Fb0d is the
vector (Fi, F, F, M, M, Me).

2.5.2 Numerical stability


In the numerical method, the complete pressure force Fb0d5 on an immersed, or partly
immersed body, can be found by integrating the pressure that is computed by the method,
over the complete boundary Fb of the body:

Fb0d =
frb
The so computed force can be coupled back to the moving body according to (2.43).
With a time step dt, the velocity vector Vb0d of the moving body can be updated:
Tffl
vbody - + (Lti IVI U body.

From this, the new position of the body can be calculated.


72 Chapter 2. Computational method

However, the above coupling may be numerically unstable. Consider a one-dimensional


example without rotation. Only motion in x-direction is allowed and the body is com-
pletely submersed. The force that is computed at time level n is written as FdY. The
acceleration A at the new time level n + i in x-direction can be computed as

A1 dy
(2.44)

This discretisation, with the force term treated explicitely, is not guaranteed to be stable.
To explain this, consider a simple hydrodynamical model from added mass potential
theory. The force FdY can then be written as
FdY = Ma A.
Here Ma is referred to as the added mass of the moving body, which is dependent on the
shape of the body. The following system for A is created:

A1 - M
Atm (2.45)

The factor MaIM can be seen as the growth factor, this means that initial errors will
grow with factor MaIM every time step. Therefore, the system (2.45) is stable, provided
that the growth factor MaIM < 1. In other words, it is required that the mass of the
moving body is larger than the added mass of the body. For example, a two-dimensional
sphere with radius r has added mass Ma = 7Tpfiuidr2 The mass of a two-dimensional
sphere with density Pbody is equal to M = 7rpbodyr2. The stability demand M > Ma then
turns into Pbody > Pfluid, meaning that the sphere has to be heavier than the same sphere
filled with liquid. To illustrate the sharpness of this limit, assuming that the flow behaves
approximately as in potential flow when the fluid and the body is initially at rest, three
simulations have been done with a two-dimensional sphere with MaIM = 2, MaIM i
and MaIM = 1/2. A small perturbation is added to the simulation and the effect in the
force for the different masses of the moving body is shown in figure 2.64. The predicted
M,M -1
20

40 10

20
-05-

-20

-40

0.002 0,004 0,006 0.006 0.002 0.004 0.006 0.008 0,002 0.004 0.006 0.008 0.01
0,110 0)

Figure 2.64: Stability behaviour of simulations of a sphere with different masses in a fluid
with a small initial perturbation.

growth factor MaIM can be recognized in the evolution of the initial perturbation in the
results.
The stability limit M > Ma is a severe restriction, since only heavy objects can be
treated. E.g. the two-dimensional cylinder in the example must be heavier than the
2.5 Coupled motion between moving bodies and fluid 73

surrounding fluid, in order to give stable results. For the kind of moving bodies with
its application in ship hydrodynamics, in order to be floating bodies, the density of the
bodies is in general much smaller than the surrounding fluid. The explicit treatment of
the force in (2.44) is now limiting the possibilities for solving the coupled problem.

A first attempt to solve this problem, Consists of treating a part of the computed force
implicitly. The idea is that the force is approximated by the linear added mass potential
theory. Write equation (2.44) as

MA'21 =

Treat the added mass part implicitely, by adding the term MaA'21 to the left hand side
of the equation and MaA'2. This gives

(M + Ma)A'2' = FdY + MaA'2. (2.46)

A part of the fluid, namely the added mass of the moving body, is treated simultaneously
with the moving body. This idea is also worked out in Gerrits [23], where the motion of a
container carrying liquid is coupled with the liquid inside the container. A 'frozen' state
of the liquid which is the position of the liquid with respect to the inertial reference frame,
is treated simultaneously with the container. Oscillations around this state correspond to
the liquid dynamics with respect to a moving reference frame. In (2.46) the added mass
of the body is 'frozen' with the moving body, while oscillations around this state appear
on the right-hand side of the equation. The term MaA'2 will approximate Fd, so this
right-hand side will be relatively small. The new system can be written as

FdY Ma
M+Ma M+Ma A'2
The growth factor M/(M + Ma) is guaranteed smaller than one. When MaA'2 is a rela-
tively good approximation of FdV, the system will be stable. For completely submersed
bodies this attempt works quite well. However, when the body is not completely sub-
mersed and in water impact problems, the added mass theory for submersed bodies is
not useful anymore for approximating the force on the body.

The above attempts to treat only part of the force simultaneously with the moving
body have failed, due to lack of robustness and difficulties that arise when the body
is piercing through a free surface. Moreover, since the problem is time dependent, it
is difficult to keep a time-accurate method in this way. The simultaneous treatment as
mentioned above, is more suitable for calculating steady flows rather than unsteady flows.
To ensure stability and time accuracy of the coupling, an iterative method is used to solve
the problem fully implicit for the time being. This necessary subiteration will slow down
the computation since a number of Poisson equations have to be solved in each time step
as will be explained below. Introduce subiteration index k, then the following iteration
is performed:
Fk
=
M
74 Chapter 2. Computational method

When the acceleration is rewritten as (un+l - u')/dt, then this becomes

(u''Y1 = utm + dt Ç7. (2.47)

The force Fc is computed by integrating the pressure over the boundary of the moving
body, therefore every subiteration k a new pressure field has to be computed correspond-
ing to the boundary velocities (tm+1)k, by solving the Poisson equation for the pressure.
Since the iteration (2.47) can diverge, as explained before, a relaxation parameter w is
introduced:
(un+ = (ul)c +w((u)* - (ufl+1)k),
where
Fc
u + dt-.
Consequently, one can write
Fk
= w(u' + dt-7) + (1 - w)(u'1. (2.48)

Here w depends on the stability of the system, which is dependent on F and M. For
problems with moving heavy bodies, the coupling will be stable without relaxation,
while lighter bodies need underrelaxation. Heavy underrelaxation will make the iter-
ation slower, while every iteration also a Poisson equation for the pressure has to be
solved. Impact phenomena where F becomes large also need underrelaxation. To find an
optimum between stability of the iteration and convergence speed, the value of w must
be as large as possible. The following procedure for finding an optimal value of w is used
in the method:
Start loop with w = 1
Perform one iteration (2.48)
Compute error as difference between the last two iterates
Check if the error diverges, if this is the case then w = w/2, reset all variables u, F
and k and go to 2.
If the error is small enough then stop, otherwise go to 2.
This procedure is bisecting the value of w until a stable value of w is found. Once this w is
found, the iteration stops when it has sufficiently converged. The amount of subiterations
in the procedure is dependent on the density of the moving body and the behaviour of
the flow. Heavy bodies show less stability problems, thus less underrelaxation is required
resulting in faster convergence. Only a few (1-3) Poisson equations have to be solved
then. Lighter bodies need stabilisation, thus underrelaxation and slower convergence.
Also during impact phenomena where high impact forces are present, convergence is
slower. Typically, the amount of subiterations lies between i and 20, dependent on the
mass of the body and the fluid forces acting on the body. This method, though it slows
down the computation due to the extra Poisson equations that have to be solved and
sometimes it shows difficulties in finding an optimal w, works reasonably well. However,
in the future, a somewhat more robust and faster method is preferred.
Chapter 3
Results
In this chapter results of simulations are discussed. First some validation has been done
in two- and three dimensions without a free surface. Then some results of applications
the method has been designed for, the motion of moving bodies piercing through a free
surface, are presented. Thereto, the method has been tested against other methods and
against experiments, like the oscillation of a rectangular cylinder piercing through the
water surface and a cylindrical section moving through the fluid. Some applications are
discussed both with motion of the body prescribed a priori and motion following from
the fluid behaviour (interactive motion). The water impact phenomenon of the wedge
entry is widely discussed. A comparison is made with experimental results of a falling
life-boat, and the simulation of some floating bodies moving to their equilibrium states.
Further, some simulation results of the sideways launching of a ship are presented.

3.1 Validation tests without free surface


3.1.1 Free stream consistency tests
One of the basic properties that should hold for any discretisation method is free stream
consistency. Therefore, also in our applications of flow around moving bodies this prop-
erty should be checked. When all velocities in the fluid are the same and the moving body
is carried along with the flow with the same velocity, the complete velocity field must
stay the same when evolving in time. This property can be validated by a channel flow
with an inflow velocity equal to the outflow velocity and a free-slip boundary condition at
the boundaries (figure 3.1). Further, an object is put in the flow with the same velocity.
First, a two-dimensional simulation of this channel flow has been performed with a square
as moving body as in figure 3.1. All velocities are set equal to one (u = u i rn/s). The
channel size is 10 by 2 m, the dimensions of the square are 1 by 1 rn. The velocity field
in the neighbourhoud of the square is shown in figure 3.2 after 5 seconds, thus after the
square has moved 5 meter. Note that also the velocities in the square have been drawn
here. From this picture the free-stream-consistency property seems to be satisfied for the
square, but to be sure, the velocity at some positions in the fluid are monitored. The
vertical line x = 6 m has been monitored, and the horizontal velocity u and the vertical
velocity w is shown in figure 3.3. The horizontal velocity u is equal to i rn/s and the
vertical velocity w is equal to zero during the complete simulation of 5 seconds as can be
76 Chapter 9. Results

free slip

inflow outflow

free slip

Figure 3.1: Free stream consistency test channel with inflow, outflow, free-slip boundaries
and a moving body.
t = 5.0 (s)

1.5

(o

0.5

65 7 7.5 8 8.5
xaxis

Figure 3.2: Velocity field of the free-stream-consistency test with a square.

seen in figure 3.3. Note that the monitored line (x = 6 rn) is passed by the moving square
during the simulation, without noticing any disturbances. This is not trivial, because the
geometry of the square, including the sharp corners, induces cut cells that vary in time.
When the discretisation near this cut cells is not done properly, disturbances are easily
created. Concluding, the free-stream-consistency test has been passed successfully for the
test with the square. Showing that the method is still free-stream consistent for arbitrary
geometries, a test has been done with a very complex geometry. A star-shaped geometry
moves with a velocity of 1 rn/s through the channel (figure 3.4). The results for this test
again do not show any disturbances in the velocity field and can thus be represented by
similar figures as in figure 3.3. Regarding these results, the method can be concluded to
be free stream consistent.

3.1.2 Added mass coefficients


An appropriate validation of the moving body simulation is the comparison with hydro-
dynamic mass coefficients as they appear in potential flow. When a submerged body is
accelerated through the fluid, it experiences a resistance force F,. from the fluid which
3.1 Validation tests without free surface 77

x=6 (m) x=6 (m)

1.01 0.01

1.005- 0.005

0.995-

0.99
2
IIII -0.005

-0.01
2
1.5

z )m) 0.5
o o time (s)

Figure 3.3: Horizontal (left) and vertical (right) velocity evolution in time of the line

*
x = 6 (m) in the free-stream-consistency test. No disturbances in the velocity fields are
observed.
free slip

inflow outflow

free slip

Figure 3.4: Free stream consistency test with a star-shaped geometry.

can be written as
Fr=m*. (3.1)

The value of m5 is called the hydrodynamic or added mass coefficient and is dependent
on the shape of the submerged body. In the applications in this study, viscous effects
play a minor role and the flow can be compared with potential flow where the fluid is
assumed to be ideal. Therefore the added mass coefficients as they are computed by the
resistance force F,. and the acceleration of the body according to (3.1) can be compared
with the known added mass coefficients for some different shapes of submerged bodies.

Two-dimensional sphere
The added mass coefficient for a two-dimensional sphere with radius r in a fluid that
behaves as in a potential flow is equal to

m* = pr2 (3.2)

with p the density of the fluid. Thus, the added mass is exactly equal to the mass of
the sphere filled with fluid. To validate this, the force on the two-dimensional sphere
has been computed by a simulation where the sphere is accelerated by = i rn/s2.
78 Chapter 3. Results

Then the added mass is computed from this force. A simulation has been done with
three different grids with a simulation time of one second. The radius of the sphere that
has been used is 10 m, and the flow domain is relatively large, namely 200 m by 200
m, to minimize the unwanted effects of the fixed walls. Theoretically, the added mass
in a potential flow according to (3.2) is equal to m* i0ir 314000 kg. The added
mass as computed from the simulation is shown in figure 3.5. The tendency is that at
350
- -. 160x160
345 - 240x240
- 320x320
340

.,335
330

¡325
C
6320
0a,
g 315
C
310

305

300o 0.2 0.4 0.6 0.8


time (s)

Figure 3.5: Added masses for a two-dimensional sphere with r = 10 m for three different
grids.

the start of the simulation the added mass is between 310000 kg and 315000 kg, which is
about equal to the added mass in a potential flow. During the simulation the added mass
slowly increases. This can be explained by the fact that at the start of the simulation,
there is not yet a viscous wake behind the sphere, thus the flow behaviour is very much
comparable with a potential flow. Therefore, the added mass that has been computed
should be equal to the added mass for a potential flow which is satisfied quite well. Later
on, a viscous wake has developed behind the sphere, resulting in an increasing resistance
force and consequently an increasing added mass. Especially the fact that the added mass
at the start of the simulation agrees with potential flow is a satisfactory result. Some
variations are visible in the signals, which are caused by numerical effects. The force at
the sphere is computed by integrating the pressure. In each cell the pressure is assumed
to be constant and no extrapolation has been performed towards the boundary of the
sphere. This causes variations in the signal when geometry labels change and when small
cells are producing pressure spikes as explained in section 2.4.4.

Two-dimensional square
The added mass in a potential flow for a two-dimensional square with edge size of a meter
is equal to
m* = 1.189pa2.
As a test case, a square has been used with a = 20 in, consequently m* 475000 kg. The
square is accelerated with lm/s2 and the added mass is computed from the simulation for
three different grids. The results are shown in figure 3.6. The added mass again slowly
3.1 Validation tests without free surface 79

Figure 3.6: Added masses for a two-dimensional square of 20 by 20 in for three different
grids.

increases due to the wake that has developed behind the square, but at the simulation
start the added mass agrees with the added mass for a flow behaving as a potential flow
( 475000 kg).

Three-dimensional sphere
A final test case to verify the added mass coefficients has been done with a three-
dimensional sphere. The added mass for a sphere in three dimensions for a potential
flow is equal to
m = irpr 23
which is equal to half the mass of the sphere filled with fluid. In the test case r = 10 m,
thus nf 2100000 kg. In figure 3.7 the added mass has been plotted as computed by
the method. The so computed added mass approximates m* quite well, again slowly
2500

2450

2400-

'235O -

2300-
¡2250-

2050

2000
0.2 0.4 0.6 0.8
time (s)

Figure 3.7: Added mass for a three-dimensional sphere with radius r = 10 in.
80 Chapter 9. Results

increasing because of viscous effects and some more high frequency variations are visible
due to grid effects explained before. Concluding the added mass validation, it may be
said that the resistance force is predicted correctly when the flow can be approximated
by a potential flow. When viscous effects play a role, the resistance force becomes higher,
which is also a correct behaviour.

3.2 Oscillation of a floating body


A first test has been done with an oscillating rectangular body in a fluid, where the body
is partly below and above the water level. The results are compared with the results found
by Yeung and Ananthakrishnan, who used a finite-difference method based on boundary-
fitted coordinates [109]. An illustration of the test definitions is found in figure 3.8. The

Figure 3.8: Test definitions of oscillating rectangular body.

equilibrium draft of the body is denoted by d, the beam by B, the amplitude of the
oscillation by a and the frequency by a. The oscillation is described by z(t) = asin(at).
According to the paper of Yeung, the variables are nondimensionalized with respect to
B, p and a. Time is scaled by 1/a, length by B, velocity by aB, pressure by pB2a2 and
force by pB3a2. Further, the following definitions are used:

R= aB2 V

B
- o.
g
A total number of six simulations has been done, all with R = i0 and d/B = 0.5.
Further, a/B = 0.1 and 0.2 and Fc. = 1.0, 1.5 and 2.0. To avoid reflections from the
left and the right boundary, a damping zone is used there. For every simulation the
hydrodynamical heave force has been calculated. The results are shown in figure 3.9 and
compared with the results of Yeung and Ananthakrishnan in [109] and linear potential
results in [108].
3.2 Oscillation of a floating body 81

a/B = 0.1, F = 2.0 a/B = 0.2, F = 2.0

- pensent method
- Young at et.
- t).Q.2sintt)
- - linear potentIal resulte

5 ito 15 20 0 5 15 20
ito

a/B = 0.1, F = 1.5 a/B o 0.2, F o 1.5

- present method - present method


- Young et el. - Young et al.
0.8- - - Ytl)O.2ain(t) 0.8 - YItl=O.2or(t)
- - linoer potentiel resulte - liseet poterrtlol resulte
0.6- 0.6

0.4- 0.4

0.2
re

-0.2
-0.4
-0.6

-0.8

io 15 20 _io 5 10 15 20

a/B=0. 1, F=i .0 a/B = 0.2, F = 1.0

0.6
- pensent method
- Young et al.
- y(tl0.2elfl«)
- present method
- Yousg et al.
ylt)=0.2slr(tl
- - 1100er potentral resulte - - Ir000r potentiel resulto
0.6 0.6-

0.4 0.4-

0.2 0.2

,//
Oi/
"\ "\
'N:..', 'T'I -

_10 5 10 15 20 5 10 15 20

Figure 3.9: Normalized dimensionless hydrodynamic heave forces for the oscillating cylin-
der with frequencies F,. = 2.0, 1.5 and 1.0 and amplitudes a/B = 0.1 and a/B = 0.2 (grid
dimensions: 201x 101).
82 Chapter 3. Results

The line y(t) = 0.2 sin(t) has been included in the results just to indicate the body
position. Comparing the results with Yeung's results, it can be seen that for all amplitudes
and frequencies the results are comparable. The deviations of the results with respect to
the results of the linear, potential-flow theory are also present and are a measure of the
nonlinear and viscous effects. Other details for different amplitudes and frequencies also
agree reasonably well with Yeung's results. A noteworthy detail is the presence of a rather
spiky behaviour during the second quarter of the periodic motion at large amplitude of
oscillation which is also seen in the results of Yeung, although less violent. During
this interval the eddies formed in the wake are convected and diffused, and the vortices
at the sides near the sharp edges are developing. The reason of this spiky behaviour
given by Yeung is that the generation of the vortices at the edges are accompanied by
drastic pressure changes. The fixed Cartesian grid method with the changing cut cells
and changing labels is quite sensitive to pressure changes (see chapter 2) and may be
more sensitive than a boundary fitted method like the method of Yeung. A velocity-
vector plot is shown in figure 3.10, to get an impression of the flow behaviour around
the edges. Here, the generation of the vortices near the sharp edges can be seen clearly
t=1.OT t=1.25T

t=1 .51 t=1 .75T

Figure 3.10: Velocity-vector plot for the oscillating cylinder with F = 2.0 and a/B = 0.2
at different moments in time. T is the period of oscillation.

during the upwards motion. These wake eddies undergo diffusion while they are convected
downwards. Thereafter, vortices in the shear layer along the sides near the sharp edges
are created. Finally, these vortices are translated outward.
3.3 Motion of a cylinder moving through a free surface 83

Concluding, the results of the oscillating body show that, despite artificial diffusion
is certainly present, the method shows similar results for these kind of flows including
nonlinear and viscous effects as the results found by the method of Yeung [109].

3.3 Motion of a cylinder moving through a free sur-


face
Moyo and Greenhow [63] studied the unsteady free motion of cylinders rising and sinking
in a fluid by a potential flow model of Brevig et al. [5]. The cylinders were initially
at rest starting from a prescribed depth. Different densities of the cylinders were used.
This method suffers from numerical break down when topological changes of the free
surface are present. This happens when upwards moving cylinders are coming out of the
fluid with only a thin layer of fluid on their upper surface, and when downwards moving
cylinders are moving through the free surface, which causes two in-flowing layers of fluid
that meet above the cylinders. Simulations have been done with a cylinder with radius
i m rising from depth D = 5 m and density Pcyi = °6Pwater and Pcyl = °9Pwater and
compared with the results of Moyo and Greenhow. Further, simulations have been done
with a cylinder of density p = 12Pwate,- and Pcyl 2Pwater sinking from D = 1.2 m.
Finally, a simulation has been done with a cylinder of density pj = '2Pwater sinking
from D = 0.0 m, thus sinking through the free surface. All these simulations have been
done with grid dimensions 81x81.
In Greenhow and Lin [29], photographs of a cylinder impacting on a free surface
are available. A numerical simulation has been done to simulate this water impact and
free-surface deformations are shown to compare with the experiments.

3.3.1 Cylinder rising to a free surface


First, a cylinder of Pcyl = °.6Pwater is initially situated at D = 5 m. The cylinder will
start rising to the free surface due to the upwards directed net force. In figure 3.11, the
1.4

1.2-
- - - added mass model
- Moyo and Greentrow
- present results
0.8-
>
0.6-

10
T

Figure 3.11: Vertical velocity of the gravitational center of the cylinder with density Pcyl =
O.6Pwater A comparison between the results of Moyo and Greenhow [63], an added mass
model and the present method.
84 Chapter 3. Results

non-dimensional velocity of the gravitational center of the cylinder is shown, together


with the results of Moyo and Greenhow [63] and an added mass model also explained in
[63], which gives an analytic solution. Here time is non-dimensionalised by T =
where r is the radius of the cylinder. The velocity is non-dimensionalised by V v/v/F.
Looking at figure 3.11, the difference between the two methods of Moyo and Greenhow
and the present method is evident. The added mass model and the potential flow model
show a linear increase of the velocity, while a non-linear increase of the velocity is present
in the actual simulation. The increase at the start of the simulation is the same in all
methods. The potential flow model however, does not notice any drag from the cylinder,
while this drag force is certainly present (d'Alembert's Paradox). The present model
calculates a drag force, forcing the cylinder to accelerate less and finally reach a constant
velocity. When the cylinder comes out of the free surface and reaches its maximum height,
the cylinder moves downwards quickly. This behaviour can be seen in all three methods.
The difference between the potential flow model and the added mass model is caused by
the very simple treatment of the free surface in the added mass model.
To get some more confidence in the drag force that has been calculated by the model,
a simulation has done with a cylinder moving horizontally through the domain with a
constant speed U i rn/s. The drag force can be calculated by Bernoulli theoretically:

FD = CDpU2A, (3.3)

where CD is the drag coefficient of a cylinder. This CD value is about 1.1 - 1.4 (e.g.
[107]) for laminar flow. Further A is the frontal area of the cylinder. The computed
force on the cylinder is shown in figure 3.12, the saw-tooth signal is due to the fact that
pressures at the cell centers are used to compute the force and not the exact location of
the boundary. From equation (3.3) and the experimental values of CD it follows that the

O8

0.6-
0.4-
0.2-

time (s)

Figure 3.12: Drag force on a horizontally moving cylinder with U = i rn/s.

drag force must lie between 1.1 and 1.4 kN for water. In the beginning of the simulation
some start-up effects are present, but after a while the value of the drag force lies between
1 and 1.3 kN. This result gives some confidence in the correct computation of the drag
forces in the case of the rising cylinder. Some more confidence in this behaviour is created
when looking at the vertical force in the simulation of the rising cylinder and the vertical
velocity (figure 3.13). The vertical force shows a more, or less linear behaviour, while the
3.3 Motion of a cylinder moving through a free surface 85

lo

Figure 3.13: Vertical dimensionless force on the rising cylinder.

velocity has a more or less parabolic profile in the first 6 seconds of the simulation (figure
3.11) . This is also correct, since from (3.3), F U2.

An impression of the free-surface deformation for the simulation with the potential
flow model described in [63] and the present simulation is given in figure 3.14.

2 2

-i
-2 -2 -2

-3
-2 0 2 - -2 0 2 -2 o 2
X X X

Figure 3.14: Free-surface deformations for rising cylinder with density pj =


computed with the potential flow method of Moyo (upper row) and with the present method
(lower row).

Since the different time evolution of both methods, due to the drag force that is absent
in the potential flow method, no time information is given in the figures, but comparable
free-surface situations. This is also the reason that the maximum height of the cylin-
der is lower in the present simulation, corresponding with the last snapshots in figure
3.14. In both methods, a strong repulsion effect from the free surface is observed when
86 Chapter 3. Results

the cylinder is near the free surface. The rapid free-surface deformations must require a
significant energy transfer, thus the cylinder must do work on the fluid that results in a
repulsion force. After this, the potential flow calculation breaks down in the troughs on
each side of the cylinder, while the present method still continues.

Moyo and Greenhow also performed simulations with other densities of the cylinder.
For comparison, one other simulation has been done with a rising cylinder, with initial
depth D = 5 m, but now with density of the cylinder Pcyl = °9Pwater The motion of the
cylinder will be much calmer in this case and the maximum height of the cylinder will
be lower. The vertical velocity evolution of the cylinder in time is shown in figure 3.15.
0.6

0.5

0.4

>0.3

0.2

0.t - - - added oreos model


- Moyo and Greonhow
present resulto

00 '0 15 20
T

Figure 3.15: Vertical velocity evolution of the gravitational center of the cylinder with
density Pcyt = 09Pwater A comparison between the results of Moyo and Greenhow [63],
an added mass model and the present method.

The same behaviour as for the previous simulation is observed. The drag force which is
present in the simulation causes the cylinder to accelerate less than in the potential flow
method, while the acceleration at the start of the simulation is the same. An impression
of the free-surface evolution is shown in figure 3.16 for the method of Moyo and Greenhow
and the present method. The early stages of the simulations are not shown since the free
surface is almost flat there.
Again, the time evolution is different, so these situations are not comparable in time.
However, the free-surface deformations are quite comparable for these situations. The
potential flow method shows a little higher maximum height of the cylinder, due to the
absence of drag. The repulsion effect from the free surface is less than in the case of the
cylinder in the former simulation, since the motion is much calmer and the deformations
of the free surface are not that big. The potential flow calculations still break down after
the last snapshot in figure 3.16. Moyo and Greenhow are therefore unable to determine
exactly when the top of the cylinder emerges from the fluid.

3.3.2 Cylinder sinking from a free surface


Next, a simulation has been done with a cylinder of density pi = l2Pwater, that starts
sinking from depth D = 1.2 m. The acceleration of the cylinder for the method of Moyo
and Greenhow and the present method is shown in figure 3.17. Again, the accelerations
3.3 Motion of a cylinder moving through a free surface 87

-3
-2 0 2

Figure 3.16: Free-surface deformations for rising cylinder with density Pcyt = °.9Pwater
computed with the potential flow method of Moyo (upper row) and the present method
(lower row).

are comparable in the beginning of the simulation, but because of the presence of a drag
force in the present method, later the acceleration is tess than in the potential flow method
of Moyo and Greenhow. A comparison of the free-surface evolution of both methods is
shown in figure 3.18. As can be seen, the same behaviour of the free surface is observed
in both methods. A shallow trough is formed, since the fluid is drawn down with the
cylinder, whereafter outgoing waves are created.
Another simulation has been done with the cylinder at the same initial depth, but
now the density of the cylinder Pcyl = 20Pwater, causing a more rapid motion of the
cylinder. Figure 3.18 shows the free-surface behaviour for Moyo and Greenhow and the
present method. Since the motion of the cylinder is faster now, a small jet is created

MeyeandGreenhow
- present meted

-OO5

Figure 3.17: Vertical acceleration of the gravitational center of the cylinder with density
Peyl = '2Pwater A comparison between the results of Moyo and Greenhow [63] and the
present method.
88 Chapter 3. Results

3. -3.
-1 0 1 2 0 1 2
X

2 2 2

o o o

-2
-3
-4
-2 o 2 -2 o 2 -2 o 2
X

Figure 3.18: Free-surface deformations for sinking cylinder with density Pcyl = '2Pwater
(upper two rows) and density Pcyl = 20Pwater (lower two rows) computed with the potential
flow method of Moyo (first and third row) and the present method (second and fourth row).
3.3 Motion of a cylinder moving through a free surface 89

above the cylinder. This causes a numerical breakdown in the potential flow method
when it starts to fall. Until this happens, the free-surface behaviour is quite comparable
in both methods.

3.3.3 Cylinder sinking through a free surface


The engulfment of a cylinder with density p = '2PWateT is studied in this subsection. In
the potential flow model of Moyo and Greenhow {63], the cylinder is initially positioned
at depth D 0.0, thus half submerged. This is done to avoid spray jets formation that
arise when the cylinder initially enters the fluid, since the numerical computation will
then break down. Moyo and Greenhow combined a method for surface piercing bodies
for the first part of the simulation, with a submerged body method for the second part of
the simulation. The two in-flowing layers above the cylinder meet at the vertical line of
symmetry above the cylinder, which is called engulfment, and this is the moment where
their surface-piercing-body method fails. The intersection points are replaced by one
intersection point and the calculation is continued with the submerged body method.
The same simulation has been done with the present method, with initial depth of the
cylinder D = 0.0. The vertical displacement of the center of mass and the vertical
velocity are shown in figure 3.19 for both methods. The vertical displacement and

Figure 3.19: Vertical displacement (left) and vertical velocity (right) of the engulfing
cylinder for the method of Moyo and Greenhow and the present method.

the velocity show a similar starting behaviour, but after that the method of Moyo and
Greenhow varies from the present method, likely because of the drag force that has not
been accounted for in Moyo's method. The hydrodynamic component of the force versus
the vertical distance is shown in figure 3.20 for both methods.
In general the hydrodynamic force that has been calculated in the present method is
higher than in the method of Moyo and Greenhow, which is partly caused by the drag
force. The force at the end may be influenced by bottom effects, since the bottom in
the present simulation was at 7 meter depth. However, both results show a decreasing
force from depth 0.2m, followed by a large hydrodynamic force, just after complete
submergence of the cylinder, which occurs approximately at a vertical displacement of
1.5 rn. This force exceeds 0.2, meaning that the total force is upwards. In [63] this large
upwards force is explained by the consideration that the cylinder must transfer energy to
90 Chapter 3. Results

0.3

0.25

0.2
Q

0.15

0.1
=
0.05

1 2 3 4
vertical displacement

Figure 3.20: Computed hydrodynarnic force of the engulfing cylinder for the method of
Moyo and Greenhow and the present method.

the fluid to support the resulting free-surface deformations. These free-surface deforma-
tions of Moyo's method are shown in figure 3.21. only the free surface is shown from just
after engulfment since no pictures were available before. Some free-surface deformations,

Figure 3.21: Free-surface deformations for engulfing cylinder with density Pcyl = i.2Pwaier
computed with the potential flow method of Moyo.

including the period before complete submergence, calculated with the present method
are shown in figure 3.22. Comparing the free-surface deformations computed by Moyo
and Greenhow and the present method, the same behaviour can be observed. Above the
cylinder a trough is formed, which closes symmetrically. A cavity is formed, which then
rapidly 'flips through' to form an upwards moving jet. This thin jet can not completely
be captured by the present grid, but the main interest here was to calculate the overall
free-surface deformations and the cylinder dynamics. In the potential method the high-
velocity thin jet is removed and the calculation is continued after that. Otherwise this jet
3.3 Motion of a cylinder moving through a free surface 91

N-1

-2

-3 -3
-4
-2 o 2 -2 o 2
X X

2 2 2

o o o

N_1 N-I N_1

-2 -2 -2

-3 -3 -3

-4 -4 -4
-2 o 2 -2 o 2 -2 o 2
X X X

2 2

o o

N-1 '41

-2 -2

-3 -3

-4 -4
-2 o 2 -2 o 2 -2 2
X X

Figure 3.22: Free-surface deformations for engulfing cylinder with density Pcyl = 12Pwater
computed with the present method.

would start to fall after it reaches its maximum height and the potential-flow calculations
would break down.

3.3.4 High speed water entry of a cylinder


To conclude the simulations of cylinders moving through a free surface, a simulation is
done with a cylinder impacting on the water surface. The diameter of the cylinder is
chosen equal to 10 cm and the vertical velocity is set constant to 1 rn/s. To capture
the thin jets that will be formed during impact, a high grid resolution is chosen (420 x
390). In figure 3.23 photos of such an experiment by Greenhow and Lin [29] are shown
together with the actual simulation to compare free surface profiles.
92 Chapter 3 esu1

Figure 3.23: Free-surface deformations for a, clinder impacting on. the water surf e,
experimental results (left) and simulation results (right).
3.4 The wedge entry phenomenon 93

3.3.5 Concluding remarks


Considering the various simulations of the moving cylinder, it may be concluded that the
free-surface deformations, computed by the method and the method of Moyo and Green-
how are similar. The present method shows more robustness, since it can handle large
deforming free-surfaces, while Moyo's method breaks down when a topological change
of the free surface occurs. Especially, when a cylinder comes out of the fluid, the latter
method will stop the calculations before the cylinder moves down again. Moreover, the
potential method does not account for any drag force, while the present method does,
with realistic drag force coefficients. This results in a more realistic motion of the cylin-
der. During high-speed water entry of a cylinder, free-surface profiles are realistic, the
jets that are formed can be captured when the grid resolution is taken high enough.

3.4 The wedge entry phenomenon


The water entry of wedge-shaped objects is a phenomenon that has been investigated
in the past theoretically, experimentally and numerically. These kind of water impact
problems were first studied by von Karman [40] and Wagner [101], later studied again by
Comte and Armand [11, 12]. More theoretical water impact studies involve the work of
Korobkin, e.g. [45-48, 77].
This water-entry phenomenon is an important aspect of this thesis, since the wedge-
shaped objects may represent ship-shaped sections. A number of simulations have been
done, both with prescribed motions and interactive motion, to compare results with
theory and experiment.

3.4.1 Model tests Marin


First, a model test carried out by MARIN, has been simulated by the method and the
results are compared. The test case exists of a wedge that is falling vertically in the
water, thus piercing through the water level (figure 3.24). The prescribed motion of the
wedge in the simulation has been chosen the same as the motion in the experiment. The

Figure 3.24: Schematic overview of the wedge model test.


94 Chapter 3. Results

7
r
Figure 3.25: Snapshots with intervals of 0.1 seconds of the simulation of downwards mov-
ing wedge hitting the water surface, case i

flare angle of the wedge is equal to 30 degrees, the height of the wedge is 17.5 cm and
the width is 10.1 cm. Two cases have been studied:

The case where the bottom of the wedge is initially at 20 cm above the water level,
whereafter it is entirely submerged and not lifted out of the water
The case where the bottom of the wedge starts at the water level, is not entirely
submerged and then lifted out of the water

More details about the experiments can be found in [15] together with the experimental
results. For case i four snapshots with intervals of 0.1 seconds of the simulation are
shown in figure 3.25, to illustrate the behaviour of the flow. A pressure signal has been
computed at the sloping edge of the wedge and this signal has been compared with the
experimental pressure signal. For case i both signals are shown in figure 3.26.
The global behaviour of the pressure signals is similar, however the computation still
shows a spiky behaviour. The experimental domain was much larger than the computa-
tional domain with respect to the right edge. No outflow boundary condition was used,
but a solid wall boundary condition, therefore a small shift in the hydrostatic pressure is
observed between the experimental pressure and the computed pressure during the entry
of the wedge. Overall it may be concluded that the agreement between the measurement
and the calculation is good. The pressure signals for case 2 are shown in figure 3.27.
The velocity of the wedge is smaller in case 2, thus the reaction of the water is less
violent. The pressure first increases and then again decreases since the wedge is lifted out
of the water finally. Again the agreement between simulation and experiment is good.

3.4.2 Impact forces on sections of a planing craft


Second, the water entry of a planing craft has been studied. According to the added mass
theory by Payne [67], the total lift on a planing craft exists of a hydrostatic component,
a dynamic lift component and a stern suction component. The effect of a planing craft
is similar to a hull section being pushed vertically into the water at a constant velocity,
as explained in [91]. Therefore, the water entry problem of a wedge-shaped body is an
important test case for the theory of planing. In [91] a CFD analysis has been carried out
with FluentV4, and compared with results from Zhao and Faltinsen [112] and Payne [68].
Also experimental work has been carried out for validation. All tests and simulations
3.4 The wedge entry phenomenon 95

50x50 100x100
5 5

4-5. 4.5-
4. 4.
3.5, 3.5.
3.
62.5

6 2-
1.5- 1:

0.5- - cxkt4aIed 0.5


- moasurod
1.2 1.6 1.6 2
tine (o)
1504150 200x200

4.5-
4.
3.5-

3.
a.2.5.

2-
1.5

0.5- - CaituSted 0.5


- measured
01 1.2 1.4 1.6 1.8 2 1.2 1.4 1.6 1.8
timo (s) timo (s)

Figure 3.26: Pressure signal at the sloping edge of the wedge for four different grids,
simulation and experiment, case 1.

have been performed with constant velocity of the hull section. The experimental set-up
was constructed in such a way that the flow can be considered two-dimensional. More
information about the experimental set-up can be found in [91].

For chine dry conditions, Tveitnes et al. [91] compared force coefficients based on
immersion depth with the boundary element method of Zhao and Faltinsen [112] and a
semi-analytical method by Payne [68]. Payne combined the theory of Von Karman [40]
and Wagner [101] which resulted in a formula for the force as a function of the immersion
depth z, the vertical velocity w and the deadrise angle fi:

F'= _ (3.4)
tan fi
From this, the force coefficient CF,0 which is defined as

F'
CF0 = (3.5)
pw z
can be derived.

Simulations have been done with a grid of 241 x 121 for three different deadrise angles
fi = 10e, 30° and 450W The maximum beam of the wedge is equal to 1 in and the vertical
96 Chapter 3. Results

1.6
calculated
1.4 - measured

1.2

0.8

0.6

0.4

0.2

_0.2o 0.5 1 1.5 2 25


time (s)

Figure 3.27: Pressure signal at the sloping edge of the wedge, simulation and experiment,
case 2 (grid: 150x150).

velocity is equal to 1.5 rn/s. An impression of the flow for /3 = 45° is given in the
snapshots in figure 3.28.
In figure 3.29 the forces have been plotted against the non-dimensional immersion
depths z/z,, where z is the immersion depth when the chines reach the undisturbed
water level, e.g. for /3 = 45°, z = 0.5 in. The hydrostatic contribution to the forces has
been subtracted to obtain only the hydrodynamic force. In dry chine conditions, the force
increases more or less linearly until the peak force is reached. After this peak force the
chines are wetted, where the semi-analytical method of Payne [68] does not hold anymore.
The linear behaviour of the simulation agrees quite well with the theory of Payne in all
three cases. The sensitivity of the pressure is again present in the simulation results. At
z/z,. 0.6 the peak forces occur. From this linear behaviour the force coefficients CF,
can be derived. This can be computed by estimating the slope of this 'linear' part of the
force signal and divide it by pw2. Tveitnes et al. [91] compared the force coefficients of
his CFD results of Fluent with the theory of Payne and the boundary element method
of Zhao and Faltinsen [112]. This is shown in figure 3.30.

Analysing the results in figure 3.29, the estimated force coefficients for the different
/3 become:

/3 = 45° : CF, 3.7


(3 = 30° : CF, 15
fi = 10° : CF, 220

These values agree well with the force coefficients in figure 3.30.
Tveitnes [91] compared the experimental results with the CFD-results of Fluent,
both for chines dry and chines wetted conditions (figure 3.31). These forces are non-
dimensionalised based on maximum beam. In certain regions the experiments give unre-
liable results, due to the design of the actual test rig, especially for low deadrise angles
3.. The wedge entry phenomenon 97

t = 0.0 (s) 0.1 (s) t 0.2 (s)

1.5

0.5 0.5
.12

-0,5 -0.5 -0.5-

-1 -1

-0.5

h - -1 2 fi -1 0 2

t = 0.3 (s) t 0,4 (S) t 0.5(s)

.5. 1.5 1.5

0.5'

-0.5-
s
0.5-

-0.5
o-'" 0.5-

-0.5

-1 -1

-1,5

2 fi - fi o
S (nl)

Figure 3.28: Snapshots with intervals of 0.1 seconds of the simulation of planing craft
with deadrise angle ¡3 450 and constant vertical velocity w = 1.5 rn/s.

and after z/z, = 0.6. However, like the computed results in figure 3.29, the peak force
occurs at values of z/z, of about 0.6, both for the Fluent results as the experimental
results. The force signals produced by Fluent are globally similar to the force signals of
the method in this thesis. Overall it may be concluded that the results of the present
simulations agree with the CFD-results and theory in [91].

3.4.3 Water entry of different-shaped two-dimensional sections


Then, the water entry of a symmetrical wedge with deadrise angle 30° and a typical bow
fiare section (figure 3.35) has been simulated. Slamming loads have been computed at the
different sections in fully interactive motion, thus no prescribed motion has been used.
The results are compared with the method described by Muzaferija et al. [64], who used
a finite volume method with a free-surface capturing model, and with the method of Zhao
et al. [113], a potential flow code. Also experimental results from drop tests of ship cross-
sections by MARINTEK are available for comparison. The big difference between the
simulations in this thesis and the simulations by Muzaferija and Zhao is that they used
the measured vertical velocity of the drop tests as input velocity for their simulations,
instead of a fully coupled motion as in this thesis. Moreover, Muzaferija used a wedge
fixed in space with prescribed inlet boundary velocities at the lower boundary and outlet
pressure boundary conditions at the upper boundary of the domain, instead of a wedge
98 Chapter 3. Results

50

45
- simulation
- Payne
15
- aimulation
- Payne
40

35
10

z
8
o
u-
5

0.5 15 0.5 15
z/z z/z

- simulation
- Payne

Figure 3.29: Computed forces for deadrise


angles /3 = 10°, /3 = 300 and 45° respectively
compared with the semi-analytical method
of Payne. The peak forces occur when the
chines get wetted, which is the case for val-
ues z/z 0.6, compare the snapshot at
t=O.3 s in figure 3.28.
0.5 1 15
z/z

moving through a fixed grid.

30° wedge
First, the slamming force on the wedge with deadrise angle 30° is discussed. The width
of the wedge is equal to 0.5 m, the length is 1 in and the total weight is 241 kg. The
length of the measuring section in the experiment is equal to 0.20 m. The velocity of
the wedge when the keel hits the water surface is approximately 6.15 rn/s. The complete
time history of the drop velocity that has been measured in the drop tests can be found
in figure 3.34. This is used as prescribed velocity for the wedge in the simulations of
Muzaferija [64] and Zhao [1131. The forces computed by Zhao et al. [113] and Muzaferija
et al. [64] are shown in figure 3.32. In all signals, only the hydrodynamic component
of the force has been plot. The computed forces are significantly higher than the
experimental results. This is caused by three-dimensional effects, which were observed
by both authors. Zhao et al. [113] successfully corrected their tw&-dimensional potential
solution to account for three-dimensional effects, while Muzaferija et al. [64] performed
three-dimensional simulations with different spacings between the wedge and the lateral
boundary to investigate three-dimensional effects and to estimate the unknown spacing
between the wedge and the tank wall in the experiments (figure 3.33).
3.4 The wedge entry phenomenon 99

F
Force if ib
10000
ICFD

1000;

100

o -
10

lH O 10 20 40 50
b

Figure 3.30: Force coefficients according to Fluent, Zhao and Faltinsen [1L2] and Payne's
theory for the planing craft with different deadrise angles.

b.W -
23

20 -
0.0

0.0
00 03 1.0 10

Figure 3.31: Non-dimensional forces for planing craft with different deadrise angles, Flu-
ent results compared with experimental data [91].

The results of the vertical velocity of the wedge and the vertical force on the wedge in
the present two-dimensional simulations is shown in figure 3.34. The grid dimensions are
201 x 151 cells. The results also overpredict the experimental forces, which is due to the
three-dimensional effect. The vertical velocity also decreases more rapidly than in the
experiment due to these somewhat higher forces. Further, compared to the results of Zhao
and Muzaferija, the shape of the signals are similar, the maximum force of 6000 N is
obtained at t 0.015 s. Then the force decreases to approximately 3000 N at t = 0.025 s.

ship section
Second, a simulation of a drop test with a flared ship section (figure 3.35) has been
performed. The width of the section was 0.32 m, the height 0.203 m, the length 1 m and
the total weight was equal to 261 kg. The length of the measuring section was 0.10 m.
The vertical velocity of the ship section that has been measured from the drop tests
has been used again as input velocity for the simulations of Zhao and Muzaferija. The
computed forces on the ship section by Zhao et al. [113] and Muzaferija et al. [64] are
shown in figure 3.36.
loo Chapter 3. Results

D Q.1O OCt O.25


7TM jj

Figure 3.32: Left: computed forces by Zhao et al. [1131. The solid line is the experimental
result, the dashed line is the numerical result and the circles mean a correction for three-
dimensional effects. Right: computed forces by Muzaferija et al. [64] for different grids.

The fully interactive two-dimensional simulation with the present method again used
a grid of 201x151 cells. The computed vertical force and computed velocity together with
the measured experimental velocity are shown in figure 3.37. Comparing these results
with the results of Muzaferija and Zhao, the computed drop velocity of the ship section
is quite comparable with the experimental drop velocity. The force also shows a similar
behaviour, however the force is a little overestimated. Some snapshots of the flow are
shown in figure 3.38.

3.4.4 Impact forces on a falling life-boat


A comparison has been made with an experimental investigation in impact forces and
pressures that are exerted on a falling life-boat dropped from different heights. Forces
at the life-boat have been calculated by measuring accelerations of the boat. A two-
dimensional section of the lifeboat has been used in the simulation (figure 3.39). The
weight of the two-dimensional section is equal to 575 kg/rn. Experiments have been done
with a wedge-shaped bottom and a flat bottom. Drop heights of 0.75 rn, 1.5 m, 2.25 in
and 3 m have been investigated for the life-boat with the wedge-shaped bottom. For the
life-boat with the flat bottom, only drop heights of 0.75 m and 1.5 in have been used,
because bigger heights were expected to damage the boat or the measuring equipment,
because the forces and pressures of the lower heights already showed very high values.
Experimental and simulation results for the wedge-shaped bottom are shown in figure
3.40, results for the life-boat with the flat bottom are shown in figure 3.41. Considering
the results for the wedge-shaped bottom, it may be concluded that the order of magnitude
is reasonably similar for the simulation results and the experimental results except for
the force in the case of 0.75 m drop height and the pressure in the case of 1.5 in drop
height. However, the experimental peak pressure in the last case is somewhat excessive,
compared to the experimental peak pressures of the other drop heights. Further, in the
last stage of the signals, the experimental signals have a slightly oscillating pattern, which
3.4 The wedge entry phenomenon 101

Figure 3.33: Forces computed by Muzaferija et al. [64] to investigate the three-dimensional
effect and the effect of changing the distance between wedge and lateral boundary.

-4 8000

7000-
-4,5
6000-

-5 5000-

4000
s
3000-
- simulation
- experiment 2000-
-6
1000

_6.5o 0.005 0.01 0.015 0.02 0.025 0.005 0.01 0.015 0.02 0.025
time (s

Figure 3.34: Left: computed vertical velocity of the wedge together with the measured
experimental velocity. Right: computed vertical force on the wedge.

is not the case for the simulations. The effect of the sprayrails (the small horizontal pieces
at the sloping edge of the life-boat, see figure 3.39) is remarkable. In the pressure signals,
a secondary peak pressure is visible in both the experiments and the simulations. This is
the effect of the first sprayrails. The forces of the simulation show a high peak force at
the same time of the peak pressure, which is not visible in the experiments. The effect
of the second sprayrails is clearly visible in the force signals of both the simulation and
experimental results. In the pressure signals the effect of these second sprayrails is hardly
noticed. This is probably because the distance between the second sprayrails location
and the location of the calculated pressure is relatively big. The pressure peak is most
noticed in the neighbourhoud of the sprayrails and is damped in space.
[n the fiat-bottomed life-boat, the experimental forces show a highly oscillating be-
haviour in the first part of the signal, due to the flat bottom. High impact forces are
already measured in the experiments with drop heights of 0.75 m and 1.5 rn. To avoid
102 Chapter 3. Results

Figure 3.35: Two-dimensional ship section, used in the drop tests and the simulations.

u',
OD O rO OCQ (

Figure 3.36: Left: computed forces by Zhao et al. [1131. The solid line is the experimental
result and the dashed line is the numerical result. Right: computed forces by Muzaferija
et al. [64] for different grids.

damage to the boat and measuring equipment, no more experiments with bigger drop
heights have been carried out. The oscillating behaviour might be an effect of structural
elasticity. In the simulations, this highly oscillating behaviour is not observed. The over-
all force in the case of drop height 0.75 in is higher than in the experiment, but this is
consistent with the wedge-bottomed life-boat. Also consistent with the wedge-bottomed
life-boat is that the computed pressure in the case of drop height 1.5 m is lower than
the experimental value. Impact pressures are very sensitive to numerical noise, since the
flat bottom will make velocities change instantly in time when the bottom hits the water
surface. No compressibility and structural elasticity modeling are included in the model
to simulate these effects.
Concluding, except for the above mentioned differences, the comparison between sim-
ulations and experiments is acceptable. Both, numerical and experimental noise is ob-
served.

3.5 Equilibrium state of various floating objects


A nice way to verify the method's ability of predicting the interactive motion of vari-
ous free-floating bodies, is to verify if they reach their correct equilibrium state. This
equilibrium state depends on the shape, mass and center of mass of the moving body.
3.5 Equilibrium state of various floating objects 103

1.8 600
- exporiment
- oifll4IatiOS
-2 500-

-2.2 400-

300

-2.6

-2.8

0.02 0.03 0.05 0.07 Oo 0.03 0.04 0.05 0.07 0,08


0.01 0.04 0.06 0.08 0.01 0.02 0,06
Sx,o (s) titIlO (s)

Figure 3.37: Left: computed vertical velocity of the ship section together with the measured
velocity in the experiment. Right: computed vertical force on the ship section.

V
t=0.Q(s) t0.01 (s) 1=0.02 (s) t 0.03 (s)

1=0.04(5) t0.05)s) 0.06(5) t 0.07(s)

Figure 3.38: Snapshots with intervals of 0.01 seconds of the simulation of a vertical drop
of a ship section.

3.5.1 Floating circular cylinder


First, a simple simulation with a two-dimensional cylinder with radius i m has been done
(the same cylinder as in section 3.3). The density of the cylinder has been taken equal to
half the density of water. The cylinder's initial depth is equal to D = 0.5 m. To verify if
Archimedes' law is satisfied, the cylinder has to reach the final equilibrium state, where
the lower half of the cylinder is submerged. The final vertical force must be equal to

Ff nat
i
= irg
Some snapshots of the simulation are shown in figure 3.42.
The size of the domain, used in the simulation, was equal to 8 m by 5 m, while the
water depth was 3 m. Initially, the free-surface line was positioned at z = 0. The final
position of the free-surface line and the position of the center of mass of the cylinder can
be calculated, when the lower half of the cylinder is submerged. The final position of the
104 Chapter 3. Results

pressure measurement wedge-Shapeo bottom

Figure 3.39: Geometry of a two-dimensional section of the falling life-boat used in the
experiments, the geometry of the flat-bottomed life-boat is obtained when the triangular
area at the bottom is removed. Also the position of the pressure measurement is marked.

free-surface line will be 0.12 m.


The temporal evolution of the vertical position of the center of mass and the vertical
force are shown in figure 3.43. It can be seen that Fiinai 15.4 kN is approximated
quite accurately by the method, as well as the final position of the center of mass of the
cylinder (-0.12 m).

3.5.2 Floating rectangular cylinder


Next, a two-dimensional, homogeneous square or a square-shaped cylinder, with density
equal to half the density of water is placed with the center of gravity at the free-surface
line as in figure 3.44. To understand the dynamics of the cylinder, the ship hydrodynam-
ics theory of ship stability is a useful tool. The forces that act on the cylinder have to be
balanced. The forces in vertical direction balance when Archimedes' law is satisfied, so
the cylinder is in equilibrium when the buoyancy force is equal to the gravity force. In the
case of figure 3.44 and with density of the cylinder half the density of water, the vertical
forces on the cylinder are balanced. The question is however, whether the rotational
forces are balanced. The center of gravity G is positioned at the center of the cylinder,
while the center of buoyancy B is positioned at the center of the submerged part of the
cylinder (figure 3.45: left). Since the direction of the gravity force is downwards and the
buoyancy force is upwards, there is no external momentum to force a rotation. Hence the
cylinder is in equilibrium as drawn in figure 3.44. However, this equilibrium is unstable.
A small rotational disturbance (figure 3.45: middle) will change the submerged volume
fraction. The triangular areas are the same, so the submerged volume to the right of the
vertical line I through G has the same magnitude as the submerged volume to the left
of that line. The center of buoyancy BL of the left submerged volume is situated closer
to the I than the center of buoyancy BR, resulting in an external moment and thus a
rotation in anticlockwise direction. A small disturbance will force the cylinder to move
from the equilibrium state, and it will find a stable equilibrium, i.e. when the cylinder is
rotated 45 degrees (figure 3.45: right). In ship hydrodynamics, the stable and unstable
equilibrium is determined by defining the metacenter M, which is defined as the limit
of the intersection point with the vertical line through B and the concerned equilibrium
3.5 Equilibrium state of various floating objects 105

0.75 er 0.75 In
80
- slmulatJxn - simulation
70-
- experiment - experiment

60-

20-

10-

812 0.14 0.16 .18 0.2 0.22 0.24 0.26 0.28 0.2 04
Orne (s) titee (o)
1.5 er
200 1.5er
- simulatJon
180- - expenment
160-

140-

o12O
100.

280
60

40

20

808 0.1 0.12 0.14 0.16 0.18 0.2


0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2 0.22 0.24
flete (e) 5mo (e)
2.25 et
250 25er
100
- expetitnent - exporteront
90
200
80

70
150

6100

30
50
20

0
0.06 0.08 0.1 0.12 0.14
.05 0.08 0.1 0.12 0.14 0.16 0.18 02
tiere (s) linee (s)

250
- simulation
- experiment
200

50

5.06 0.08 0.1 0.12 0.14 .08 0.08 0.1 0.12 0.14 0.16 0.18 02
titeo (s) time(o)

Figure 3.40: Experimental and simulation results for the falling life-boat with wedge-
shaped bottom, dropped from 0.75 m, 1.5 m, 2.25 m and 3 m. Pressures are shown left
and forces are shown right.
106 Chapter 3. Results

0.75 rn 0.75 m
70
- simulation
- experiment
60-

50-

f40.
30

20-

10-

0.12 0.14 0.08 0.18 0.2 0.22 0.24 0.26 0.28


Orne(s)
1.5 m
200
250
- simulation 180
- expsrlmonI
180
200-
140

oI5
lOO
280
100-
60
40
50-
20

0.18 02 0.1 0.2 03


.08 0.08 0.1 0.12 0.14 0.16
Orne (s) tinte(s)

Figure 3.41: Experimental and simulation results for the falling life-boat with flat bottom,
dropped from 0.75 m and 1.5 rn. Pressures are shown left and forces are shown right.
3.5 Equilibrium state of various floating objects 107

t = 0.0 (s)
1.5 t = 0.6 (s)
1.5

0.5-
0.5-
o

E
-1-

-2-
-2 ...................
-2.5

-3 -2 -1 O 3 -2 -1 0 3
x(m) s (m)
t = 1.2 (s) t= 1.8 (s)

-2 -1 0 I 2 3 -2 -1 2
s )m)
t= 2,4 (s)
1.5 1.5

0.5

-1-

-2
-2.5- -2.5

-2 -1 O 2 3 -2 -1 0 3
s (m( X (rs)

Figure 3.42: Snapshots of the simulation of a floating cylinder with density half the density
of the fluid. The last snapshot shows the final equilibrium state.
108 Chapter 3. Results

20

19

18

17

16
a,
15

14

13

12

110 5 10 15 20
time (s)

Figure 3.43: Temporal evolution of the vertical position of the center of mass (left) and
the vertical force (right) of the simulation of the floating cylinder. The dashed lines denote
the equilibrium states.

Figure 3.44: Initial position of two-dimensional square with density Psquare = °5Pwater

axis, when small rotational disturbances approximate zero. When M lies above G, the
equilibrium is stable, when M lies below G, the equilibrium is unstable. This metacenter
M can be calculated by (e.g. see [89]) MC = MB - GB, where MB = I/V5. MB is the
distance between the metacenter and the center of buoyancy, GB is the distance between
the center of gravity and the center of buoyancy, I = 1b3/12 is the transverse moment
of inertia of the waterplane area of the body (1 = length of the section in transverse
direction, taken equal to 1 m for simplicity, since it is a two-dimensional situation, and
b = waterline beam). V8 is the submerged volume of the body. If MG> 0, the equilib-
rium is stable, if MG < O the equilibrium is unstable. For a squared section 10 cm by
10cm with submerged volume 10cm by 5cm, MB = I/V8 = (1*0.1/12)/O.005 = 1/60.
GB = 1/40, thus MG = 1/60 - 1/40 = 1/120 < 0, hence the equilibrium is unstable.

To verify that the cylinder with initial phase in figure 3.44 is in unstable equilibrium,
one can do an experiment with a piece of wood for example in a kitchen sink. The
3.5 Equilibrium state of various floating objects 109

Figure 3.45: Unstable equilibrium for rectangular cylinder (left), small rotational dis-
turbance to the unstable equilibrium causing a rotational motion away from the unstable
equilibrium (middle) and the stable equilibrium (right).
70

60

50

40

030

20

10

o 4 6 10
time (s)

Figure 3.46: Time history of the angle of rotation O of the rectangular cylinder with
density half the density of water. The dashed line denotes the equilibrium state.

cylindrical piece must be square-shaped and must be long enough to simulate a two-
dimensional situation. The density of the wood has to be approximately equal to half
the density of water (e.g. pine wood). When the cylinder is initially put as in figure 3.44,
after release it will rotate 45 degrees to find its stable equilibrium.

A numerical simulation has been done with a two-dimensional square of 10 cm by


10 cm with as initial disturbance a small rotational velocity is given. The time history
of the total anglè of rotation O is shown in figure 3.46.
As expected, the cylinder rotates away from its unstable equilibrium and finds a new
equilibrium 45 degrees rotated, after a damped oscillation around that equilibrium.

Another simulation has been done, but now the density of the cylinder has been taken
equal to 0.25 times the density of water. The initial position of the cylinder is still the
110 Chapter 3. Results

same. Therefore, the cylinder will start rising, and will reach a stable equilibrium state,
with one quarter of the cylinder submerged. The equilibrium where the cylinder has not
rotated, is unstable since MG = MB - GB = I/V3 - 3/4 1 * 0.1/(12 * 0.1 * 0.025) -
3/80 = 1/30 - 3/80 1/240 < 0. The angle of rotation is shown in figure 3.47. The

Figure 3.47: Time history of the angle of rotation O of the rectangular cylinder with
density 0.25 times the density of water. The dashed line denotes the equilibrium state.

cylinder starts to rotate and shoots over the rotation of 45 degrees, then it oscillates
around its equilibrium state of 63 degrees of rotation. That means that the cylinder is
floating with one vertex at the free surface and the other vertex below the free surface,
in such a way that the cylinder is submerged for one quarter (figure 3.48). Theoretically,
this state will be at angle O = 90 arctan0.5 = 63.4°.
Yet, one other two-dimensional simulation has been done with a rectangular cylinder.

Figure 3.48: Theoretically stable equilibrium state for rectangular cylinder with density
0.25 times the density of water.

Now, the cylinder has size 10 cm by 5 cm and has density equal to half the density of
water. Initially, the cylinder is half submerged, with the long edge vertically directed
(figure 3.49). This equilibrium state is unstable and therefore will start to rotate to
reach its stable equilibrium state which is with the long edge in horizontal direction,
thus rotated 90 degrees. This can be seen again by computing the metacenter M via
8.6 Sideways launching of a ship 111

20

-20

-40

a
° -60

-80
to
-100

-120
0 1 2 3
time (s)

Figure 3.49: Initial position of rectangular cylinder (left). Time history of the angle of
rotation of the rectangular cylinder 10 cm by 5 cm with density 0.5 times the density
of water (right).

MG = MB - GB, which is in this case equal to 1/240 > 0. The time history of the angle
of rotation is shown in figure 3.49. For a simulation time of 5 seconds, grid dimensions
85x52, time step 0.001, the computational time was about 1 hour on a pc.

To conclude the verification of the interactive method by predicting the correct equi-
librium states, a three-dimensional simulation has been done with a rectangular cylinder
with a square base and the long edge twice as long as the short edge. The density of the
cylinder is equal to half the density of water. Initially, the cylinder is positioned with its
bottom just above the free surface. Combining the former two-dimensional simulations
of the squared section and the rectangular section, the cylinder will eventually float along
the long edge and the squared section rotated 45 degrees, so one vertex downwards. To
get an impression of the motion, some snapshots of the simulation are shown in figure
3.50. Indeed, the cylinder moves to its theoretical equilibrium state as becomes clear
from the snapshots. Actually, it is very illustrative to carry out such an experiment in
your kitchen sink with a piece of pine wood or another material with density of about
half the density of water.
Overall, summarizing the simulations in this section, it may be concluded that the
method can predict the correct equilibrium states of variously shaped floating bodies.

3.6 Sideways launching of a ship


An interesting problem in shipbuilding is the launching of a ship into the water. Sideways
launching of a ship has been done for more than two hundred years and this type of
launching is based on a lot of experience. For new ships, or unusual types of ships, it is
sometimes considered to be too risky to launch a ship sideways. Then the ship is launched
forward or is lifted in the water, which is much more expensive. A picture of a sideways
launch of a ship is shown in figure 3.51.
A pilot study to simulate the sideways launching of a ship with the method in this
112 Chapter 3. Results

Figure 3.50: Snapshots of the simulation of a three-dzmensional floating rectangular cylin-


der with density half the density of the fluid.
3.6 Sideways launching of a ship 113

Figure 3.51: Side launch of 78 meter Standard Coaster Hull. This picture is taken from
http://www.georgeprioreng.com/hull/hull_images/.

thesis has been carried out by De Jong [35]. Two-dimensional validations with exper-
iments have been done and some fully three-dimensional simulations have been carried
out. For full details, see [351.
Some results of the two-dimensional validation study are shown in figure 3.52. The
experiments consisted of a rectangular block of length 1.76 m, width 0.47 m and height
0.228 m, sliding into a rectangular basin of length 8 m and width 1.5 rn with adjustable
waterdepth. Different parameters have been varied, such as drop height, waterdepth, ship
weight, slope etc. The experiments were carried out at Delft University of Technology
[33]. In figure 3.52, a selection from [35] has been made of three simulations, namely the
standard simulation, then the slope of the sleighs has been doubled, and finally the center
of gravity of the block has been moved upwards. The simulations have been done with
grid dimensions of 120x60, which means that the cell size in vertical direction is equal
to 1 cm. Simulation results and experimental results show a globally similar behaviour,
though in general the displacements are somewhat less for the simulations than the ex-
periments. This may partly be due to a three-dimensional effect in the experiments,
since fluid is able to flow around the body. In the simulation the resistance force may be
overestimated, resulting in smaller displacements. Another reason may be the effect of
numerical spikes in the simulation. A force spike may have only a little effect on the body
motion, but more spikes can cumulatively cause a more significant effect. More results
can be found in [35].
Chapter 3. Results

06

E
5

E s
a

06
experiment experiment
simulation simulation

0.2
E
E E
E

L a0.1
- experiment
simulation

05
experiment - experiment
- - - - simulation - - - - simuletion

0.2
E E
E E

s
a - experiment
- - - - simulation

01
0.5 i tinre )s)15 2.5 0 0.5 i time ))i .5 2 25 0 0.5 i time(s)1.5 2 2.5

Figure 3.52: Some results of the two-dimensional validation for sideways launching of
a ship. A comparison with experiments. Dashed lines represent simulations, solid tines
represent experiments. Standard simulation (upper row), doubling of the sleigh slope
(middle row), center of gravity moved upwards (lower row).
Chapter 4
Summary and conclusions
In this thesis, a numerical method has been presented for simulating fluid flow with a
free surface, including moving rigid bodies. The study has been focused on designing a
simulation tool for green-water loading on ships and offshore structures and other impact
phenomena. Key aspect has been the ability of the method to simulate large deforming
free surfaces in combination with arbitrary moving complex bodies in two and in three
dimensions. The method has been implemented in a computer program called ComFlo.

Moving boundaries
Two types of moving boundaries are involved in the flow, namely the free surface and
the boundary of the moving rigid body. In chapter 2, the discretisation of these different
boundaries is discussed.
The free surface is described by a volume-of-fluid method. The VOF-method is a
so-called Eulerian method, this means that the interface is captured from Eulerian data
on a fixed computational grid. In this way the computational grid does not need to
be adjusted to the free surface every time step, as in Lagrangian methods. Because in
the applications in this study large deformations of the free surface and even topological
changes are possible, it would be very complicated to adjust the computational grid to
this free surface. Furthermore, the method used in this thesis is able to maintain a
relatively sharp interface by combining the VOF method with a local level-set function.
A combined Lagrangian-Eulerian method is used for the boundary of a moving rigid
body. The computational grid is still fixed and the interface of the rigid body is tracked
explicitely, so a sharp interface is maintained. Since arbitrary complex moving bodies
can be used, it is a very time-consuming task to use boundary fitted grids, while a
fixed Cartesian grid is trivial to generate. However, the Cartesian grid can be cut by
the boundary in many ways. This results in the creation of so-called cut cells near the
boundary of the rigid body, which also depend on time. In diffuse interface methods, like
immersed boundary methods and fictitious domain methods, the information near the
interface is smeared over a few cells. In these methods it is not important to know exact
information about cut cells near the boundary. However, in convection dominated flow
with high velocities along the boundaries, it is important to discretise the equations based
on the detailed information near the boundary, resulting in a sharp interface method. In
two dimensions it is relatively easy to compute the cut cells from a polygon that is
116 Chapter 4. Summary and conclusions

defined by connecting a set of marker points. In three dimensions the reconstruction


of the cut cells becomes more complicated, but a relatively simple approach based on
marker particles is found to be sufficient for calculating cut-cell data.

Flow discretisation
Also in chapter 2, the incompressible governing Navier-Stokes equations including the
free surface and the motion of a rigid body are discretised.
The flow discretisation is based on a finite-volume method with staggered variables.
Since the velocities are defined at the cell faces of the Cartesian cells, mass conservation
can be satisfied up to machine precision. The pressure is defined in the cell centers,
thus the pressure gradients will be defined also at the cell faces. This property makes
it possible to discretise the continuity equation and the pressure gradient in such a way,
that the underlying analytical property that the divergence operator is equal to minus the
transpose of the gradient operator, is preserved. Further, the skew symmetry property
of convection is preserved in the discretisation and diffusion is discretised such that its
negative definite character is preserved, thus dissipating energy. From these discrete
properties, it can be derived that the energy in the discrete system can not increase when
no external forces are present, making the method very robust. The inclusion of moving
bodies may have influence on the energy via pressure variations, but this is completely
controlled by the velocity of the moving body.
Near solid boundaries, like domain boundaries and moving body boundaries, cut cells
are created. The interface between the fluid and the solid is kept sharp by discretising the
governing equations using detailed information about cut cell volumes and cut areas. No
remeshing and cell merging is performed, thus arbitrary small cut cells can be present.
For non-moving boundaries, it can be proven that the discretisation of convection near cut
cells does not sharpen the time-step restriction that follows from the CFL-condition for
the uncut cells. The presence of a moving boundary does have an influence on the time-
step restriction for convection in small cells, but this can be controlled by a stabilisation
factor that is added to the momentum equations. This factor depends on the difference
of the momentum control volumes between two time steps, thus dependent on whether
a boundary is moving normal or tangential to itself. Accurate discretisation of diffusion
shows stability problems in small cells. However, diffusion is not very important in the
applications in this thesis, and it can be discretised in a 'staircase' manner, as if the cells
were uncut, to avoid stability problems. Further, some subtleties have been discussed
in the discretisation of the continuity equation and mass fluxes that are present in the
discretisation of convection, by considering two alternatives.
For the free surface, a labeling system has been introduced to distinguish the fluid
from the air. Tangential and normal stress boundary conditions are applied at the free
surface. A pressure interpolation procedure has been performed for surface cells, to ac-
count more accurately for the free-surface location in a cell. Velocity boundary conditions
at the free surface have been investigated, and turn out to have a crucial effect on the
pressure. Pressure spikes have been observed widely in previous free-surface simulations.
Conservation of mass is a critical issue in this investigation. However, the energy of the
system can be influenced by velocity boundary conditions and demanding conservation
of mass at the free surface is not always the best option with respect to energy conser-
117

vation. Sometimes computations even break down due to energy increase via velocity
boundary conditions. A rather safe solution for applying boundary conditions has been
found, based on a constant extrapolation of related internal velocities, still producing
a reasonable smooth pressure signal. It has been concluded that sometimes a trade-off
between smooth pressure signals and robustness has to be made. The introduction of a
two-phase flow model is probably the best way to overcome this problems, and recently
a start has been made to implement such a model in the method.
Pressure spikes are not only produced by the free surface treatment. Also a moving
body may be the cause of a spiky pressure signal. For moving bodies with sharp corners,
spikes will be present in the pressure. This is a consequence of loss of geometrical in-
formation within a cell. These spikes can be overcome by a partitioning of the physical
pressure and a pressure potential containing the unphysical part of the computed pres-
sure (as in the fictitious domain method). Further, the discretisation of the pressure as
is described in this thesis, gives rise to irregularities in the neighbourhoud of cut cells.
The exact positioning of the pressure is not defined, and appears to be dependent on
the cell face areas and momentum control volumes. Also, rapidly moving free surfaces
along a moving body are very sensitive with respect to irregularities in the pressure. An
attempt has been made for the coupled motion of a moving body and the surrounding
fluid. Crucial is the stability of the numerical coupling, which depends on the mass of
the body when the force on the body is treated explicitly. A more expensive implicit
coupling has been introduced to overcome these stability problems.

Applications
The method has been especially designed for studies of floating ship-type and offshore
structures, interacting in waves with or without impact phenomena such as green-water
loading or bow slamming. In chapter 3, the method is validated and comparisons are
made with other methods and experimental results. Simulations with prescribed motion
and with interactive motion are performed. An extensive study of the water entry of
wedge-shaped geometries is done, investigating the ability of the method to predict this
phenomenon. It is concluded that the method is able to predict this quite well, though
unphysical irregularities in the computed pressure are still present. The method has also
been applied in a study regarding the sideways launching of ships showing promising
results. Less impacting phenomena have also been investigated, such as heave motions
of rectangular objects, cylinders rising to and sinking from a free surface, and variously
shaped objects interacting with a fluid converging to their equilibrium states. All these
simulations have been found to be successful. In the future, green-water loading on ships
will be further investigated with the present method. Next to the offshore applications,
the method can also be applied in many industrial applications where free liquid surfaces
play a role.

Concluding remarks
Regarding the objectives that have been mentioned in section 1.1, the following can be
concluded:
118 Chapter 4. Summary and conclusions

Arbitrary moving bodies have been succesfully implemented in ComFlo in two and
in three dimensions, avoiding remeshiug or cell merging. A beginning is made with
interactive motion, showing quite promising results.
The free-surface treatment has been improved with respect to spiky pressure signals
and robustness.
Impact phenomena have been studied extensively with ComFlo, showing realistic
results, though force and pressure signals still suffer from a spiky behaviour.
Bibliography
R. Almgren. Variational algorithms and pattern formation in dendritic solidifica-
tion. J. Comp. Phys., 106:337-354, 1993.

N. Ashgriz and J. Y. Poo. FLAIR:flux line segment model for advection and inter-
face reconstruction. J. Comp. Phys., 93:449-468, 1991.

E. F. F. Botta and M. H. M. Ellenbroek. A modified SOR method for the pois-


son equation in unsteady free-surface flow calculations. Journal of Computational
Physics, 60:119-134, 1985.

J. U. Brackbill, D. B. Kothe and C. Zemacli. A continuum method for modelling


surface tension. J. Comp. Phys, 100:335-354, 1992.

P. Brevig, M. Greenhow and T. Vinje. Extreme wave forces on submerged cylinders.


In Proceedings of the Second BHRA International Symposium on Wave and Tidal
Energy, pages 143-166, Cambridge: British Hydraulics research Association, 1981.

B. Buchner. Green water on ship-type offshore structures. PhD thesis, Delft Uni-
versity of Technology, 2002.

B. Buchner and J. L. Cozijn. An investigation into the numerical simulation of


green water. In Proc. mt. Conference on the Behaviour of Offshore Structures.
BOSS'97, pages 113-125, Delft, 1997. Elsevier Science, Oxford, Vol. 2.

M. ten Caat. Numerical simulation of incompressible two-phase flow. Master's


thesis, University of Groningen, september 2002.

J. W. Cahn and J. E. Hilliard. Free energy of a nonuniform system, I. Interfacial


free energy. J. C'hem. Phys., 28:258-267, 1958.

S. Chen, D. B. Johnson and P. E. Raad. Velocity boundary conditions for the


simulation fo free surface fluid flow. Journal of Computational Physics, 116:262-
276, 1995.

R. Comte. Two-dimensional water-solid impact. J. Offshore Mech. 4 Arctic Engng,


111:109-114, 1989.

R. Comte and J-L. Armand. Hydrodynamic impact analysis of a cylinder. J.


Offshore Mech. & Arctic Engng, 109:237-243, 1987.
120 Bibliography

P. Colella and P. R. Woodward. The piecewise parabolic method (PPM) for gas-
dynamical simulations. J. Comp. Phys., 54:174-201, 1984.
E. F. G. van Daalen. Numerical and theoretical studies of water waves and floating
bodies. PhD thesis, University of Twente, 1993.
E. F. G. van Daalen. Time domain simulations of bow flare oscillations with a two-
dimensional boundary element method. Technical Report 14580-1-0E, MARIN,
November 1998.

E. F. G. van Daalen, J. Gerrits, G. E. Loots and A. E. P. Veldman Free surface


anti-roll tank simulations with a volume of fluid based navier-stokes solver. In
T. Miloh and G. Zilman, editors, Proc. 15th hit. Workshop on Water Waves and
Floating Bodies, Caesarea, Israel, February 27 - March 1, 2000.
E. F. G. van Daalen, K. M. T. Kleefsman, J. Gerrits, H. R. Luth and A. E. P. Veld-
man. Anti-roll tank simulations with a volume-of-fluid based Navier-Stokes solver.
In Twenty- Third Symposium on Naval Hydrodynamics, pages 457-473, National
Academy Press, Washington D.C., 2001.
[18J G. Ersdal and A. Kvitrud. Green water on Norwegian production ships. In Proc.
the 10th ISOPE Conf., Seattle, 2000.
0. M. Faltinsen. Sea loads on ships and offshore structures. Cambridge University,
1990.

L. J. Fauci and C. S. Peskin. A computational model of aquatic animal locomotion.


J. Comp. Phys., 77:85-108, 1988.
G. Fekken, A. E. P. Veldman and B. Buchner. Simulation of green-water loading
using the Navier-Stokes equations. In J. Piquet, editor, Proceedings 7th Interna-
tional Conference on Numerical Ship Hydrodynamics, pages 6.3-1-6.3-12, Nantes,
France, July 19-22, 1999.
Foley, van Dam, Feiner and Hughes. Computer graphics, principles and practice.
Addison Wesly, 1990.

J. Gerrits. Dynamics of liquid-filled spacecraft. PhD thesis, University of Groningen,


The Netherlands, 2001.
J. Gerrits and A. E. P. Veldman. Numerical simulation of coupled liquid-solid
dynamics In G. Bugeda E. Onate and B. Suarez, editors, Proc. ECCOMAS 2000,
Barcelona, Spain, 2000. paper 575.
J. Gerrits and A. E. P. Veldman. Transient dynamics of containers partially filled
with liquid. Moving Boundaries VI, pages 63-72, 2001. In: B. Saner and C. A.
Brebbia (eds.).
R. A. Gingold and J. J. Monaghan. Smoothed Particle Hydrodynamics: theory and
application to non-sperical stars. Mon. Not. R. Astron. Soc., 181:375-389, 1977.
Bibliography 121

R. Glowinski, T. W. Pan, J. Pénaux. A Lagrange multiplier/fictitious domain


method for the numerical simulation of incompressible viscous flow around moving
rigid bodies: (I) case where the rigid body motions are known a priori. C. R. A cad.
Sci. Paris t. 324, Série 1, Mathematical problems in mechanics, 1997.
R. Glowinski, T. W. Pan, T. I. Hesla. D. D. Joseph and J. Pénaux. A fictitious
domain approach to the direct numerical simulation of incompressible viscous flow
past moving rigid bodies: application to particulate flow. Journal of Computational
physics, 169:363-426, 2001.

M. Greenhow and W.-M. Lin. Non linear free surface effects: experiments and the-
ory. Technical Report 83-19, Department of Ocean Engineering, MIT, Cambridge,
Mass.

S. Guignard et al. Solitary wave breaking on sloping beaches: 2-d two phase flow
numerical simulation by si-vof method. Eur. J. Mech. B - Fluids, 20:57-74, 2001.
X. He, C. Fuentes, W. Shyy, Y. Lian and B. Carrol. Computation of transitional
flows around an airfoil with a movable flap. AIAA Fluids 2000 and Exhibit, Paper
No. AIAA-2000-2240, June, 19-22, Denver, Colorado 2000.

C. R. Hirt, and B. D. Nichols. Volume of Fluid (VOF) method for the dynamics of
free boundaries. Journal of Computational Physics, 39:201-225, 1981.
C. M. van Hooren. Dwarsscheeps te water laten. Technical report, Delft, University
of Technology, Laboratorium voor scheepsbouwkunde, 1971. Rapport no. 297.
R. A. Johnson and D. M. Belk. Multigrid approach to overset grid communication.
AIAA J., 33:2305-2308, 1995.
P. de Jong. Numerical simulation of sideways launching of a ship. Master's thesis,
University of Groningen, the Netherlands, 2003. In preparation.
D. Juric and G. 1yggvason. A front tracking method for dendritic solidification.
J. Comp. Phys., 123:127-148, 1996.
H.-C. Kan, H. S. Udaykumar, W. Shyy and R. Tran-Son-Tay. Hydrodynamics
of a compound drop with application to leukocyte modeling. Physics of Fluids,
10:760-774, 1998.

H.-C. Kan, H. S. Udaykumar, W. Shyy and R. Tran-Son-Tay. Numerical analysis of


the deformation of an adherent drop under shear flow. J. Biornech. Eng., 121:160-
169, 1999.

11.-C. Kan, H. S. Udaykumar, W. Shyy, P. Vigneron and R. Trari-Son-Tay. Effects


of nucleus on leukocyte recovery. Annals of Biomedical Engineering, 27:648-655,
1999.

T. von Karman. The impact of sea planes floats during landing. NACA TN, 321,
1929.
122 Bibliography

K. Khadra, P. Angot, S. Parneix and J. P. Caltagirone. Fictitious domain ap-


proach for numerical modelling of Navier-Stokes equations. International journal
for numerical methods in fluids, 34:651-684, 2000.
K. M. T. Kleefsman and A. E. P. Veldmari. Numerical simulation of wave loading
on a SPAR platform. In Proceedings of IWWWFBO3, International Workshop on
Water Waves and Floating bodies, Le Croisic, France, April, 2003.
[43J K. M. T. Kleefsman, G. Fekken, A. E. P. Veidman, B. Buchner, T H J Bunnik and
B. Iwanowski. Prediction of green water and wave loading using a Navier-Stokes
based simulation tool. In A. Naess et al., editor, Proceedings 2lth ASME Offshore
Mechanics and Arctic Engineering, 2002. ISBN 0-7918-3611-8, American Society
of Mechanical Engineers.
R. Kobayashi. Modelling and numerical simulations of dendritic crystal growth.
Physica D., 63:410-423, 1993.
A. A. Korobkin. Formulation of penetration problem as a variational inequality.
Din. Sploshnoi Sredy, 58:73-79, 1982.

A. A. Korobkin. Asymptotic theory of liquid-solid impact. Phil. Trans. Roy. Soc.


Lon. A, 355:507-522, 1997.

A. A. Korobkin. Shallow water impact problems. J. Eng. Math., 35:233-250, 1999.


A. A. Korobkin and V. V. Pukhnachov. Initial stage of water impact. Ann. Rev.
Fluid Mech., 20:159-185, 1988.

D. B. Kothe and R. C. Msjolsness. RIPPLE: a new model for incompressible flows


with free surfaces. AIAA J., 30(11):2692-2700, 1992.
R. J. LeVeque and Z. Li. The immersed boundary method for elliptic equations with
discontinuous coefficients and singular sources. SIAM J. Numer. Anal., 31(4):399-
417, 1994.

P. Y. Liang. Numerical method for calculation of surface tension flows in arbitrary


grids. AIAA J., 29:161-167, 1991.
H. Liu and K. Kawachi. A numerical study of insect flight. J. Comp. Phys.,
146:124-156, 1998.

H. Liu and K. Kawachi. A numerical study of undulatory swimming. J. Comp.


Phys., 155:223-247, 1999.

G. E. Loots. Fluid-structure interaction in hemodynamics. PhD thesis, University


of Groningen, The Netherlands, 2003.
E. Loots, B. Hillen and A. E. P. Veldman. The role of hemodynamics in the
development of the outflow tract of the heart. Journal of Engineering Mathematics,
45:91-104, 2003.
Bibliography 123

E. Loots, B. Hillen, H. Hoogstraten, A. Veidman. Fluid-structure interaction in the


basilar artery. In R. Herbin and D. Kroener, editors, Finite Volumes for Complex
Applications III, pages 607-614. 2002. ISBN 1-9039-9634-1, Hermes Fenton Science.
L. B. Lucy. A numerical approach to the testing of fusion process. Astronomical
J., 88:1013-1024, 1977.
J. E. Melton. Automated Three-Dimensional Cartesian Grid Generation and Eu-
lerFlow Solutions for Arbitrary Geometries. PhD thesis, Univ. CA. Davis CA,
1996.

G. Meskers. Realistic inflow conditions for numerical simulation of green water


loading. Master's thesis, Delft University of Technology, 2002.
J. J. Monaghan. Simulating free surface flows with sph. J. Comp. Phys., 110:399-
406, 1994.

J. J. Monaghan and A. Kos. Solitary waves on a Cretan beach. J. Waterway, Port,


Coastal, and Ocean Engng, ASCE,125,3, 1999.
W. D. M. Morris, J. Millar and B. Buchner. Green water susceptibility of North
sea FPSO/FSUs. In Proc. 15th conference on floating production systems (FPS),
London, 2000.
S. Moyo and M. Greenhow. Free motion of a cylinder moving below and through a
free surface. Applied Ocean Research, 22:31-44, 2000.
S. Muzaferija, M. Peric, P. Sames and T. Schellin. A two-fluid navier-stokes solver
to simulate water entry. In Proceedings of the Twenty-Second Symposium on Naval
Hydrodynamics, pages 638-649, 2000.
S. Osher and J. A. Sethian. Fronts propagating with curvature dependent speed:
Algorithms based on Hamilton-Jacobi formulations. J. Comp. Phys., 79:12-49,
1988.

N. A. Patankar, P. Singh, D. D. Joseph, R. Glowinski and T.-W. Pan. A new


formulation of the distributed Lagrange multiplier/fictitious domain method for
particulate flows. mt. J. Multiphase flow, 26:1509-1524, 2000.
P. R. Payne. Design of High Speed Boats, volume 1:Planing. Fishergate Inc.,
Annapolis, Maryland, 1988.
P. R. Payne. Recent developments in "added-mass" planing theory. Ocean Engng,
21, 1994.

D. H. Peregrine and L. Thais. The effect of entrained air in violent water impacts.
J. Fluid Mech., 325:377-397, 1996.

D. H. Peregrine and M. E. Topliss. The pressure field due to steep water waves
incident on a vertical wall. In Proc. 21th Internat. Conf on Coastal Engng. ASCE
2, pages 1496-1510, 1994.
124 Bibliography

C. S. Peskin. Numerical analysis of blood flow in the heart. J. Comp. Phys.,


25:220-252, 1977.

W. J. Rider and D. B. Kothe. Reconstructing volume tracking. J. Comp. Phys.,


141:112-152, 1998.

B. D. Rogers, R. A. D. Dairymple, O. Knio and M. Gesteira. Smoothed Particle


Hydrodynamics for naval hydrodynamics. In Proc. 18th International Workshop on
Water Waves 4 Floating Bodies (IWWWFB), Le Croisic, France, April, 6-9, 2003.

J. E. Romate. The numerical simulation of nonlinear gravity waves in three di-


mensions using a higher order panel method. PhD thesis, University of Twente,
1989.

Z. E. Sabeur, J. E. Cohen, J. R. Stephens and A. E. P. Veidman. Investigation on


the free surface flow oscillatory impact pressures with the volume of fluid method.
In K. W. Morton and M. J. Baines, editors, Numerical methods for fluid dynamics,
volume VI. Oxford Science Publications, pages 193-198.

R. Scardovelli and S.Zaleski. Direct numerical simulation of free-surface and inter-


facial flow. Annu. Rev. Fluid Mech., 31:567-603, 1999.

Y.-M. Scolan and A. A. Korobkin. Three-dimensional theory of water impact. part


1. inverse wagner problem. J. Fluid Mech., 440:293-326, 2001.

J. A. Sethian. Numerical algorithms for propagating interfaces: Hamilton-jacobi


equations and conservation laws. J. Duff. Geom., 31:131-161, 1990.

J. A. Sethian and J. Strain. Crystal growth and dendritic solidification. J. Comp.


Phys., 98:231-253, 1992.

W. Shyy. Computational modeling for fluid flow and interfacial transport. Elsevier
Science Publishers B.V., Amsterdam, The Netherlands, 1994.

W. Shyy, H. S. Udaykumar, M. M. Rao and R. W. Smith. Computational fluid


dynamics with moving boundaries. Taylor & Francis, Washington, D.C., 1996.

W. Shyy, M. Francois, H. S. Udaykumar, N. N'dri and R. Tran-Son-Tay. Moving


boundaries in micro-scale biofluid dynamics. Applied Mechanics Reviews, pages
405-454, 2001.

J. B. Smith. Shape instabilities in pattern formation in solidification: A new method


for numerical solution of the moving boundary problem. J. Comp. Phys., 39:112-
127, 1981.

J. L. Steger. Thoughts on the Chimera grid method of simulation of three-


dimensional viscous flow. In Proc. Computational Fluid Dynamics Symposium on
Aeropropulsion, pages 1-10, 1991. NASA CF-3078
Bibliography 125

M .Sussman and E. Fatemi. An efficient, interface preserving level set re-distancing


algorithm and its application to interfacial incompressible fluid flow. SIAM J. Sci.
Comput., 20(4) :1165-1191, 1999.

M. Sussman and E. G. Puckett. A coupled level set and volume-of-fluid method for
computing 3d and axisymmetric incompressible two-phase flows. J. Comp. Phys.,
162:301-337, 2000.

M. Sussman, P. Smereka and S. Osher. A level set approach for computing solutions
to incompressible two-phase flow. J. Comp. Phys., 114:146-159, 1994.

M. E. Topliss, M. J. Cooker and D. H. Peregrine. Pressure oscillations during wave


impact on vertical walls. In Proc. 23rd Internat. Conf. Coastal Engng. ASCE 2,
pages 1639-1650, 1992.

B. Trenhaile. Understanding stability. Hawaii Ocean Industry and Shipping news,


April/May 1998.

M. P. Tulin and M. Landrini. Breaking waves in the ocean and around ships. In
Proc. 23rd ONR Symposium on Naval Hydrodynamics, Val de Reuil, France,, 2000.

T. Tveitnes, K. S. Varyani and A. C. Fairlie-Clarke. CFD analysis and experimental


work on water impact forces on transverse sections of planing craft. MARNET-CFD
First Workshop - Barcelona, 1999.

0. Ubbink and R. Issa. A method for capturing sharp fluid interfaces on arbitrary
meshes. J. Comp. Phys., 153:26-50, 1999.

H. S. Udaykumar, H.-C. Kan, W. Shyy and R. Tran-Son-Tay. Multiphase dynamics


in arbitrary geometries on fixed Cartesian grids. J. Comp. Phys., 137:366-405, 1997.

H. S. Udaykumar, R. Mittal, P. Rampunggoon and A. Khanna. A sharp inter-


face Cartesian grid method for simulating flows with complex moving boundaries.
Journal of computational physics, 174:345-380, 2001.

s. o. Unverdi and G. Tryggvason. A front tracking method for viscous, incompress-


ible, multi-fluid flows. J. Comp. Phys., 100:25-37, 1992.

V. Venkatakrishnan Perspectives on unstructured grid flow solvers. AIAA J.,


34:533-547, 1996.

R. W. C. P. Verstappen and A. E. P. Veldman. A symmetry-preserving Cartesian


grid method for turbulent flow. In preparation.

R. W. C. P. Verstappen and A. E. P. Veldman. Data-parallel solution of the in-


compressible Navier-Stokes equations. In P. Wesseling, editor, High performance
computing in fluid dynamics, pages 237-260, Kluwer Academic Publishers, Dor-
drecht, The Netherlands, 1996.
126 Bibliography

[99] R. W. C. P. Verstappen and A. E. P. Veidman. Numerical computation of a viscous


flow arouñd a circular cylinder on a Cartesian grid. In Proceedings ECCOMASS
2000, Barcelona, Spain, 11-14 september 2000.
[loo] V. R. Voller and C. Prakash. A fixed grid numerical modeling mehodology
for convection-diffusion mushy region phase-change problems. mt. J. Heat Mass
Transf., 30:1709-1719, 1987.
H. Wagner. Über stoß- und gleitvorgänge an der oberfläche von flüssigkeiten (phe-
nomena associated with impacts and sliding on liquid surfaces). Zeitschrift Für
Angewandte Mathematik Und Mechanik, 12:193-215, 1932.
M. J. Walkden, D. J. Wood, T. Bruce and D. H. Peregrine. Impulsive seaward
loads induced by wave overtopping on caisson breakwaters. Coastal Engineering,
42:257-276, 2001.
A. A. Wheeler, B. T. Murray and R. J. Schaefer. Computations of dendrites using
a phase field model. Physica D., 66:243-262, 1993.
D. J. Wood, D. H. Peregrine and T. B. Bruce. Study of wave impact against a wall
with pressure-impulse theory: part 1, trapped air. Journal of Waterways, Port
Coastal and Ocean Engineering, ASCE, 126:182-190, 2000.
F. Xiao. A computational model for suspended large rigid bodies in 3d unsteady
viscous flows. J. Comp. Phys., 155:348-379, 1999.
G. Yang, D. M. Causon and D. M. Ingram. Calculation of compressible flows about
complex moving geometries using a 3d Cartesian cut cell method. mt. Journal for
numerical methods in fluids, 33:1121-1151, 2000.
T. Ye, R. Mittal, H. S. Udaykumar and W. Shyy. An accurate Cartesian grid
method for viscous incompressible flows with complex immersed boundaries. Jour-
nal of computational physics, 156:209-240, 1999.
R. W. Yeung. A hybrid integral-equation method for the time-harmonic free-surface
flows. In Proceedings, First International Conference on Numerical Ship Hydrody-
namics, Gaithersburg, Maryland, pages 581-608, 1975.
W. Yeung and P. Ananthakrishnan. Oscillation of a floating body in a viscous fluid.
Journal of Engineering mathematics, 26:211-230, 1992.
D. L. Youngs Time-dependent multimaterial flow with large fluid distortion. In
K. W. Morton and M. J. Baines, editors, Numerical methods for fluid dynamics,
Aceademic Press, New York, pages 273-285.
D. L. Youngs. An interface tracking method for a 3D Eulerian hydrodynamics
code. Technical report, AWRE/44/92/35, Atomic Weapons Research Establish-
ment, April 1987.
R. Zhao and O. Faltinsen. Water entry of two-dimensional bodies. J. Fluid Mech.,
246, 1992.
Bibliography 127

[113] R. Zhao, O. Faltinsen and J. Aarsnes. Water entry of arbitrary two-dimensional


sections with and without flow separation. In Proceedings of the Twenty-First
Symposium on Naval Hydrodynamics, pages 408-423, 1997.
S amenvatt ing

Motivatie
Als schepen op zee in een zware storm verzeild raken, kunnen de golven zo hoog worden
dat er water op het dek van bet schip stroomt. Dit wordt bet 'groen water' probleem
genoernd, naar het phytoplankton dat bet water groen kleurt. Dit groen water kan
serieuze schade veroorzaken aan constructies die op het dek zijn geplaatst. Ook offshore
platformen kunnen schade ondervinden van groen water.
Ret groen water probleem was de belangrijkste motivatie voor het onderzoek dat in dit
proefschrift wordt bescbreven. 0m meer inzicht te krijgen in bet gedrag van groen water
op schepen en offshore constructies, is bet doen van experimenten nog steeds een van de
belangrijkste middelen. Modelproeven zijn uitgevoerd door het MARIN (Maritime Re-
search Institute Netherlands) in grote waterbassins waarbij specifleke golven door middel
van computergestuurde golfopwekkers kunnen worden gecreëerd. Ret realiseren van dit
soort experimenten is echter erg duur en tijdrovend. Een andere methode om het gedrag
van stromingen te bestuderen is de numerieke simulatie. De bewegingsvergelijkingen die
de stroming beschrijven, namelijk de Navier-Stokes vergelijkingen die al meer dan 150
jaar bekend zijn, worden met behuip van de computer benaderend opgelost. Theoretisch
zijn deze vergelijkingen namelijk slechts voor enkele zeer eenvoudige gevallen op te lossen.
In 1998 is de mogelijkheid onderzocht 0m het gedrag van groen water te voorspellen met
het computerprogramma ComFlo, ontwikkeld aan de Rijksuniversiteit Groningen. De
resultaten van dit onderzoek waren positief, echter een beperkende factor was dat in
de simulatie het schip niet kon bewegen. De realiteit is natuurlijk anders, vandaar het
onderwerp van dit proefschrift, waarbij ComFlo is uitgebreid met de mogelijkheid orn
een willekeurig bewegend object door de vloeistof te laten bewegen. Een belangrijk doel
hiervan is orn in de toekomst een schip te simuleren dat beweegt onder invloed van de
golven inclusief effecten zoals groen water. Alle numerieke aspecten die hierbij naar voren
komen, zoals de discretisatie van het vrije vloeistof oppervlak en de bewegende objecten,
de methode voor het oplossen van de Navier-Stokes vergelijkingen en de problemen die
zieh hierbij voordoen, alsmede validatie van de methode door middel van vergelijking met
theorie, experimenten en andere methoden worden beschreven in dit proefschrift.

Het rekenrooster en discretisatie


In hoofdstuk 2 van dit proefschrift wordt de methode uitgelegd voor de numerieke si-
mulatie. Er wordt begonnen met het formuleren van de onsamendrukbare Navier-Stokes
vergelijkingen. Deze vergelijkingen zijn gebaseerd op een aantal behoudswetten, llamelijk
behoud van massa, impuls en energie. Vervolgens wordt bet rekendomein opgedeeld in
130 Samenvatting

een rooster van cellen, en worden voor elke cel de behoudswetten opgesteld. Het stelsel
van vergelijkingen dat hierdoor wordt gevormd, wordt door de computer opgelost. Hoe
meer cellen gebruikt worden, hoe nauwkeuriger de oplossing wordt, maar ook hoe langer
de berekening duurt. Hierin kan een optimum worden gevonden. In ons geval gebruiken
we rechthoekige cellen, een zogenaarnd Cartesisch rooster. Voordeel van zo'n rooster is
dat de boekhouding ervan vrij eenvoudig is, omdat het regelmatig en overzichtelijk is en
alle cellen op dezelfde manier geörienteerd zijn. Nadeel ten opzichte van een zogenaarnd
boundary-fitted rooster (een rooster waarbij de randen van het rekengebied aansluiten
bij het rooster) is dat de randen van het bewegend object op verschillende manieren
door het rooster snijden. Hiermee moet rekening gehouden worden met het discreti-
seren van de behoudswetten. Voordeel is wel weer dat het rooster niet steeds hoeft te
worden aangepast aan het bewegende object. De discretisatie in de doorsneden cellen,
moet nauwkeurig gebeuren orn de rand van het object 'scherp' te houden. Cellen kunnen
willekeurig klein worden, waardoor nurnerieke stabiliteitsproblemen kunnen ontstaan De
meeste methoden gaan dit soort problemen uit de weg door de rand over een paar cellen
uit te smeren, maar deze methoden zijn meer geschikt voor diffusie-gedomineerde stro-
mingen. Voor convectie-gedomineerde stromingen, waarbij hoge snelheden langs wanden
kunnen voorkomen, moet de rand echter zo scherp mogelijk worden behandeld. Ret blijkt
dat, in vergelijking met een stilstaand object, een bewegend object in de stroming tot
stabiliteitsproblemen leidt in de buurt van kleine cellen.

Bewegende randen
Twee verschillende soorten bewegende randen zijn aanwezig in het model, namelijk het
vrije oppervlak (het scheidingsvlak tussen bet water en de lucht) en de rand van het
bewegende object. Deze randen hebben een verschillend karakter, omdat het vrije opper-
vlak vrij kan bewegen en vervorrnbaar is, terwijl de positie van de rand van het bewegend
object volledig wordt vastgelegd door translatie en rotatie. Ze worden dan ook op een
verschillende manier behandeld in de methode. Het vrije oppervlak wordt beschreven
door een zogenaamde Volume Of Fluid (VOF) methode. In elke cel wordt de hoeveelheid
vloeistof bijgehouden door de zogenaamde VOF-functie, die een waarde tussen O (lege
cel) en i (volle cel) heeft. Tijdens elke rekenstap wordt gekeken hoeveel vloeistof er in
elke cel instroomt of uitstroomt. Zo kan de positie van het vrije opperviak worden bijge-
houden en kan de methode willekeurig vervormbare oppervlakken aan. De randen van het
bewegende object kunnen door het onvervormbare karakter worden bijgehouden door het
massamiddelpunt bij te houden inclusief rotatiehoeken. In principe kunnen de doorsnij-
dingen van de rand met de rechthoekige roostercellen exact worden uitgerekend. In twee
dirnensies gaat dit vrij gemakkelijk door het object te beschrijven met een aantal pun-
ten die met elkaar worden verbonden en zodoende een polygon vormen. Vervolgens kan
door middel van een simpel 'clip' algoritrne het gedeelte object en het gedeelte vloeistof
in elke cel worden bepaald. In drie dirnensies kan in principe jets soortgelijks worden
gedaan, maar wordt de procedure een stuk ingewikkelder. Daarom is voor het 3D geval
een eenvoudigere procedure gekozen, gebaseerd op het met merkers vullen van het object.
De nauwkeurigheid van de geometriebeschrijving kan worden aangepast door het aantal
merkers te vergroten. De merkers worden dan expliciet verplaatst volgens de beweging
van het object.
Samenvatting 131

Gevoeligheid van de druk


Een probleem dat zich voordoet, is dat de druk die door de methode wordt berekend
erg gevoelig is. Omdat het vrije opperviak en het object door het vaste rooster ver-
plaatsen, verandert het discrete stelsel voortdurend. Kleine onregelmatigheden kunnen
relatief grote gevolgen hebben voor de druk. Dit uit zich voornamelijk in een spijkerig
druksignaal in de tijd: 'gras'. Het blijkt dat randvoorwaarden bij het vrij opperviak een
grote invloed kunnen hebben op de druk. Massabehoud speelt hierin een essentiële rol.
Als massabehoud wordt geëist aan het vrij opperviak kan een redelijk glad druksignaal
worden gegarandeerd. Er ontstaan soms echter problemen in de doorsneden cellen bij
het vrije opperviak. Hier moet nauwkeurig mee orn worden gegaan. Massabehoud eisen
blijkt niet altijd de beste keuze, er kan hierdoor namelijk energie worden toegevoegd aan
het systeem, waardoor de simulatie kan exploderen. Beter is orn de randvoorwaarden te
relateren aan inwendige sneiheden, ten koste van een jets minder glad druksignaal. Het
blijkt dat soms een keuze moet worden gemaakt tussen robuustheid en gladde druksig-
nalen. Er blijken nog meer redenen te zijn voor 'gras' in de druk, die te maken hebben
met scherpe hoeken in de geometrie, of door de keuze van de discretisatie, of door een
combinatie van een vrij oppervlak langs een bewegend object. Sommige ervan kunnen
worden opgelost, sommige (nog) niet, zoals beschreven in hoofdstuk 2.

Interactie
In eerste instantie zijn de numerieke aspecten van bewegende objecten onderzocht bij
voorgeschreven beweging. De realiteit is echter dat een object krachten ondervindt van
de vloeistof en beweegt onder invloed van die krachten. Deze interactie tussen de vloeistof
en het object kan numeriek worden opgelost. Echter, expliciet oplossen van deze interactie
levert numeriek een stabiliteitsprobleem op, waarbij de massa van het object een grote rol
speelt. Als de massa te klein wordt gekozen, kan de oplossing numeriek instabiel worden.
Impliciet oplossen van de interactie kan dit stabiliteitsprobleem verhelpen, maar dit is
veel duurder met betrekking tot rekentijd.

Resultaten
In hoofdstuk 3 wordt de methode gevalideerd en een vergelijking gemaakt met andere
methoden en experimentele resultaten. Simulaties met voorgeschreven beweging en met
interactieve beweging zijn uitgevoerd. 0m te beginnen zijn 'eenvoudige' fenomenen on-
derzocht waarvan gedetailleerd validatiemateriaal beschikbaar is, zoals op- en neerwaartse
bewegingen van rechthoekige objecten in water, en zinkende en stijgende cilinders in corn-
binatie met een vrij opperviak. Ook interactieve simulaties met een aantal verschillende
vorrnen en dichtheden van objecten zijn uitgevoerd die naar hun juiste evenwichtstoe-
stand convergeren. Daarna zijn complexere verschijnselen onderzocht, bijvoorbeeld de
intreding van wigvormige objecten in water, orn de mogelijkheid van de methode orn dit
fenomeen te voorspellen te onderzoeken. Er kan worden geconcludeerd dat dit vrij goed
kan worden voorspeld, hoewel er nog steeds onregelmatigheden in de berekende druk
aanwezig zijn. Ook een heftige intreding van een cilindervormig object is gesimuleerd.
In de figuur wordt een vergelijking gemaakt tussen een experiment en simulatie van het
gedrag van het vrije vloeistof opperviak tijdens deze intreding. De methode is ook ge-
132 Samenvatting

bruikt voor een studie naar het zijdelings te water laten van. een schip; wat veelbelovende
resultaten opÏeverde. Toekomstige toepassingen zijn onder andere groen-water simulaties
op séhepen in interactie met golven. De flexibiliteit van de methode. staat echter een
groot aantál andere toepassingen toe, niet alleen offshore maar ook talrijke industriële
toepassingen.

Vergelijking van experiment (boyen) en simulatié (onder): intreding van een cilinder-
vormig object in water.
D ankwo ord

Na de totstandkoming van mijn proefschrift, wil ik graag nog een aantal mensen bedanken.
Allereerst wil ik mijn promotor, Arthur Veidman, bedanken voor het in goede banen leiden
van mijn onderzoek. Zijn ondersteuning, ideeën en ongeevenaarde enthousiasme waren
een enorme stimulans voor mij in de afgelopen jaren.
Het MARIN wil ik bedanken voor het financieren van het onderzoek en het leyeren
van validatiemateriaal. Met name wil ik Bas Buchner bedanken voor zijn enthousiasme
en vertrouwen, het lezen van mijn proefschrift en de suggesties. Verder wil ik Tim Bunnik
en Ed van Daalen bedanken voor de samenwerking.
De beoordelingscommissie, bestaande uit Aad Hermans, Jo Pinkster en Guus Stelling
wil ik bedanken voor het zorgvuldig lezen van mijn proefschrift.
Ook wil ik TM-AIO's Erwin, Jeroen, Marc en Theresa bedanken voor de samenwerking
en de nuttige ideeën voor de ontwikkeling van ComFlo.
Verder wil ik nog Ferry Roelofs, Ralph Postma, Garrelt Alberts, Geert Meskers en
Peter de Jong bedanken voor de samenwerking aan ComFlo. Ook Frans Sas wil ik
bedanken voor de interessante opdrachten en bet nuttige validatiemateriaal.
Voor de leuke tijd in het IWI wil ik in de eerste plaats bedanken mijn kamergenoten
Erwin, en later Arie. Verder wil ik hiervoor alle medewerkers van TM bedanken en alle
AlO's en postdocs uit het IWI die ik heb meegemaakt.
Erwin, Hubert, Arjan, Pauline, Richard, Susan, Richard, Ronald en Jantina wil ik
bedanken voor de leuke tijd naast miju promotiewerkzaamheden. Verder wil ik mijn
familie bedanken en speciaal mijn ouders die mij altijd hebben gesteund. Heit, Mem,
bedankt voor alles. Tot slot wil ik rnijn vriendin Nynke bedanken, die mij in de laatste
anderhalf jaar met heel veel liefde heeft gesteund bij bet voltooien van dit pro efschrift.

You might also like