Vous êtes sur la page 1sur 60

Applied Computational Fluid Dynamics

with Sediment Transport in a Sharply


Curved Meandering Channel

- Diploma Thesis / Diplomarbeit -


at / am

Institute for Hydromechanics


University of Karlsruhe (TH)
in close collaboration with / in Zusammenarbeit mit

Department of Hydraulic and Environmental Engineering


Norwegian University of Science and Technology (NTNU)

submitted by / vorgelegt von

Jörn Wildhagen∗
Matrikelnummer: 995044

September 29, 2004

Supervision : Dipl.-Ing. Nils Rüther (NTNU)


: Dipl.-Ing. Martin Detert (TH KA)

University of Karlsruhe (TH)


Institute for Hydromechanics
Kaiserstraße 12
D–76131 Karlsruhe

wildhagen@ifh.uni-karlsruhe.de
joern.wildhagen@web.de
Statement / Erklärung

This thesis has been submitted to Institute for Hydromechanics, University of Karls-
ruhe (TH), in fulfillment of the requirements for achieving academic degree. I hereby
testify that the work is original. Other used sources are well identified through
references within this text and summerized in the last chapter. I agree, that this
thesis is exhibited at a library and can be duplicated by photocopies.

Diese Diplomarbeit wurde dem Institut für Hydromechanik an der Universität Karls-
ruhe (TH) zur Erreichung des akademischen Grades des Diplomingenieurs vorgelegt.
Ich versichere, dass ich diese Diplomarbeit selbständig verfasst, noch nicht ander-
weitig für andere Prüfungszwecke vorgelegt, keine anderen als die angegebenen Quel-
len und Hilfsmittel benutzt sowie wörtliche und sinngemäße Zitate als solche gekenn-
zeichnet habe. Ich erkläre mich damit einverstanden, dass meine Diplomarbeit in
eine Bibliothek eingestellt sowie kopiert wird.

Karlsruhe, September 29, 2004

Jörn Wildhagen
Abstract
This thesis contains results from a numerical simulation of sediment transport in a
sharply curved meandering channel. These results were performed by the numer-
ical model SSIIM. The numerical results are tested against the measurement data
taken from a physical experiment with steady flow conditions, gained after a dura-
tion time of ∆t = 4 h. The results are analyzed in order to determine the capabilities
and limitations of the SSIIM model to reproduce the physical processes of sediment
transport.
The comparison of the numerical simulation by SSIIM and the measurement data
show that the overall trends in deposition and formation of point bars agree on
the main points. Concerning erosion, the numerical model predicts the same spot
of erosion zone as observed in the experiment, but SSIIMs quantitative calculation
of erosion is highly overpredicted in comparison to the experimental results. The
numerical model results have been proved to be independent of the time step ∂t, but
show some dependency on the grid size. Nevertheless, the results correspond with
observations of the conducted experiment.
Deviations between model and experimental results are discussed as well as uncer-
tainties in numerical modelling.

Kurzfassung
In dieser Diplomarbeit werden nummerische Simulationen zum Sedimenttransport
in einem eng mäandrierenden Kanal mit dem Simulationsprogramm SSIIM durch-
geführt. Ein Experiment mit stationärem Abfluss wird nachgebildet. Die Ergeb-
nisse der nummerischen Simulation werden mit experimentellen Daten nach einer
Durchführungsdauer von ∆t = 4 h verglichen. Die Ergebnisse werden analysiert, um
die Fähigkeiten und Grenzen des nummerischen Modells SSIIM im Hinblick auf die
Reproduzierbarkeit der physikalischen Prozesse des Sedimenttransportes aufzeigen zu
können.
Der Vergleich zwischen den Simulationsergebnissen von SSIIM und den Messergeb-
nissen zeigt, dass die Ablagerung von Sedimenten und die Formation von Anhäu-
fungspunkten weitgehend übereinstimmen. Das nummerische Modell ermittelt den
gleichen Erosionsbereich wie im Experiment beobachtet. Allerdings wird die quan-
titative Berechnung der Erosion von SSIIM überbewertet, wie ein Vergleich mit den
vorliegenden Messdaten zeigt. Die nummerischen Ergebnisse zeigen, dass sie un-
abhängig von der Wahl des Zeitschrittes ∂t sind, aber eine leichte Abhängigkeit in
der Rechengittergröße besteht. Die Ergebnisse liefern insgesamt eine zufrieden stel-
lende Übereinstimmung mit weiteren Beobachtungen aus dem Experiment.
Abweichungen zwischen den Modell- und Experimentergebnissen werden diskutiert
sowie Unsicherheiten bei nummerischen Berechnungen aufgezeigt.
Nomenclature
Symbol Unit Explanation
A m2 flow cross–sectional area
a m reference level
B m width
C kg/m3 concentration
ca ppm reference concentration
cµ ,c1 ,c2 ,σk ,σ − k– turbulence model constant
D, d m particle diameter
D∗ − dimensionless particle diameter
DH m hydraulic diameter
Et m/s2 turbulent diffusion coefficient
Em m/s2 molecular diffusion coefficient
Fr − Froude–number
F r∗ − particle Froude–number
g m/s2 acceleration due to gravity
h m flow depth
Ie − energy slope
Io − channel slope
k m2/s2 turbulent kinetic energy
K 1/day or 1/s reaction coefficient
ks m equivalent sand roughness
1
kSt m /3/s Strickler–coefficient
L m length scale of turbulence
p N/m2 pressure
Pw m wetted perimeter
Q m3/s discharge
qb m3/ms bed–load transport rate
Re − Reynolds–number
Re∗ − particle Reynolds–number
s − specific density
T − transport stage parameter
t s time
u∗ m/s bed shear velocity
u0∗ m/s bed shear velocity related to grains
V̂ m/s velocity scale of the turbulent motion
x, y, z m cartesian coordinates
Z − suspension parameter
u, v, w m/s velocity components in x,y,z–direction
v

Greek Symbols
Symbol Unit Explanation
α − transverse bed slope
β − ratio of sediment and fluid diffusion coefficient
β0 − averaging factor
δij − Kronecker delta, 1 for i = j, else 0
 m2/s3 dissipation rate
κ − von-Karman-constant, κ = 0.4
µ N s/m2 dynamic viscosity
µt N s/m2 eddy–viscosity
ν m2/s kinematic viscosity
φ − angle of repose of sediment particles
%,ρ kg/m3 density
σt − turbulent Schmidt number
σg − geometric standard deviation of particle grain size
τ N/m2 shear stress
τ̃ − τb/τ
h

θ − stream–wise/longitudinal bed slope

Subscripts
Symbol Explanation
’ fluctuation quantity
time averaged quantity
b quantities associated with bed/bottom
c center
crit critical
h quantities associated with horizontal bed
i,j,k component running index
m mean
r radial
s sediment
sur surface
w water
84, 50, 16 sediment grain size d; defined as the size for which 84%, 50% or 16%
weight of the material is finer
vi

Abbreviations
Symbol Explanation
cf. confer
CFD Computational Fluid Dynamics
e.g. exempli gratia (lat.) = for instance
eqn. equation
et al. et alii (lat.) = et al.
fig. figure
i.e. id est (lat.) = that is to say
no. number
resp. respectively
SSIIM Sediment Simulation In Intakes with Multiblock option
SIMPLE Semi–Implicit Method for Pressure–Linked Equations
Contents
Statement ii

Abstract iii

Nomenclature iv

List of Figures ix

List of Tables ix

Acknowledgement x

1 Introduction 1

2 Fundamentals 2
2.1 Governing Equations for Water Flow Calculation . . . . . . . . . . . 2
2.2 Assumptions and Approximations . . . . . . . . . . . . . . . . . . . . 3
2.3 Governing Equations in Simplified Form . . . . . . . . . . . . . . . . 4
2.4 Turbulence Modelling . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.4.1 Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.4.2 Reynold Stress Terms . . . . . . . . . . . . . . . . . . . . . . . 5
2.4.3 Boussinesq’s Approximation and Eddy–Viscosity . . . . . . . . 5
2.4.4 Eddy–diffusivity Concept . . . . . . . . . . . . . . . . . . . . . 6
2.4.5 The k– Turbulence Model . . . . . . . . . . . . . . . . . . . . 6

3 Sediment Transport Mechanism 8


3.1 Shields Parameter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
3.2 Incipient Motion of Sediment Particles on Generalized Sloping Fluvial
Beds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
3.3 Transport Modes and Particle Motion Modes . . . . . . . . . . . . . . 10
3.4 Computation of Transport Quantities . . . . . . . . . . . . . . . . . . 11
3.4.1 Bed–Load Transport . . . . . . . . . . . . . . . . . . . . . . . 11
3.4.2 Suspended Load . . . . . . . . . . . . . . . . . . . . . . . . . . 13

4 Flow in Curved Open Channels 15


4.1 Secondary Flows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
4.2 Formation of Secondary Flows . . . . . . . . . . . . . . . . . . . . . . 15
4.3 Distribution of Longitudinal Velocity . . . . . . . . . . . . . . . . . . 17
4.4 Distribution of Vertical Velocity . . . . . . . . . . . . . . . . . . . . . 17
4.5 Effects on Sediment Transport . . . . . . . . . . . . . . . . . . . . . . 17

5 Numerical Model SSIIM 19


Contents viii

6 Description of Physical Model 20


6.1 Geometry and Flow Parameters . . . . . . . . . . . . . . . . . . . . . 20
6.2 Sediment Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

7 Numerical Simulation 24
7.1 Represented Domain . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
7.2 Presentation of Water Flow Results . . . . . . . . . . . . . . . . . . . 24
7.2.1 Basic Flow Configuration . . . . . . . . . . . . . . . . . . . . . 25
7.2.2 Secondary Flow . . . . . . . . . . . . . . . . . . . . . . . . . . 25
7.2.3 Distribution of Longitudinal Velocity . . . . . . . . . . . . . . 26
7.2.4 Distribution of Vertical Velocity . . . . . . . . . . . . . . . . . 28
7.2.5 Stream Separation and Formation of Recirculation Zone . . . 28
7.2.6 Concluding Remarks . . . . . . . . . . . . . . . . . . . . . . . 32
7.3 Presentation of Sediment Transport Results . . . . . . . . . . . . . . 32
7.3.1 Basic Sediment Configuration . . . . . . . . . . . . . . . . . . 33
7.3.2 Default Case . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
7.3.3 Base Case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
7.3.4 Parameter Variation . . . . . . . . . . . . . . . . . . . . . . . 35
7.4 Uncertainties in Numerical Modelling . . . . . . . . . . . . . . . . . . 39

8 Conclusions and Outlook 41

A Listing C++ Subroutine 42

B Listing of Sediment Calculation Input File 44

C Photos 46

Bibliography 48
List of Figures
1 Definition sketch of angles θ and α . . . . . . . . . . . . . . . . . . . 10
2 Variation of τ̃ with α according to Deys empirical equation . . . . . . 11
3 Dimensionless particle motion diagram . . . . . . . . . . . . . . . . . 12
4 Vertical distribution of sediment concentration according to Rouse [20] 14
5 Definition sketch of transverse inclination of surface and transverse
circulation at bend in channel . . . . . . . . . . . . . . . . . . . . . . 16
6 Definition sketch of channel shape seen from above . . . . . . . . . . 21
7 Definition sketch of channel shape seen from side . . . . . . . . . . . 21
8 Particle size distribution . . . . . . . . . . . . . . . . . . . . . . . . . 23
9 Definition sketch of mesh, top view . . . . . . . . . . . . . . . . . . . 24
10 Definition sketch of cross–section, seen looking upstream . . . . . . . 25
11 Velocity vector plot at cross-section no. F243 . . . . . . . . . . . . . . 26
12 Position sketch of cross–sections . . . . . . . . . . . . . . . . . . . . . 27
13 Longitudinal velocity distribution . . . . . . . . . . . . . . . . . . . . 29
14 Horizontal velocity distribution (begin) . . . . . . . . . . . . . . . . . 30
14 Horizontal velocity distribution (end) . . . . . . . . . . . . . . . . . . 31
15 Sediment transport results conducting default case . . . . . . . . . . . 34
16 Comparison of bed elevation changes between base case and experiment 36
17 Bed elevation changes with regard to different bed roughness heights 37
18 Sediment calculation with refined mesh and ∂t = 60s . . . . . . . . . 38
19 Sediment calculations with distorted mesh at different time steps . . . 39
20 Channel before experimental run . . . . . . . . . . . . . . . . . . . . 46
21 Photo of channel bend showing one point bar . . . . . . . . . . . . . 47
22 Photo of channel bend showing separated point bars . . . . . . . . . . 47

List of Tables
1 The values of the k– turbulence model [19]. . . . . . . . . . . . . . . 7
2 Computation time for a simulation time of ∆t = 4 h . . . . . . . . . . 37
Acknowledgement
This diploma thesis is the result of collaboration between the author, presently at
the Department of Hydraulic and Environmental Engineering in the Norwegian Uni-
versity of Science and Technology (NTNU) and the author’s home department, the
Institute for Hydromechanics at the University of Karlsruhe (TH). The study has
been carried out at the Department of Hydraulic and Environmental Engineering,
NTNU in summer 2004.

First of all I would like to thank both my supervisors at NTNU, Prof. Nils Rei-
dar B. Olsen and Nils Rüther for their advice and support. They have always been
available and took time for discussion. Their help guided me through this work and
led me to a deeper understanding of sediment engineering and computational fluid
dynamics, which I appreciated very much. It has been a pleasure to work with them.

I would also like to mention with thanks the help of Ian Guymer and Richard
Dutton from the Department of Civil and Structural Engineering at the University
of Sheffield. They provided me with all requested data from the physical experiment
in order to carry out reasonable simulations and gave me advice during my work.

Furthermore, I would like to thank all the people at my working place and all the
master’s thesis students I met at the ’brakka’ during lunch time. I really enjoyed
their warm welcome.

I also want to express my gratitude to all the people who encouraged and supported
my request to go abroad during my final thesis at a foreign university.

And finally I am deeply indebted to my parents, who provided me with both moral
and financial support during my period of study.
1 Introduction
Natural meandering rivers are very complex in their water flow situation. The stream
is characterized by turbulent, strongly three-dimensional and irregular channel topog-
raphy. Due to spiral motion, which is also known as secondary current or transverse
circulation, the river tends to erode the outer bank, yielding deposits at the inner
bank. These phenomena cause local scouring and local pooling. Important engi-
neering efforts have been undertaken on rivers of all scales to stabilize the banklines.
Therefore quantitative information with respect to erosion and deposition must be
made available for sound river management, e.g. for navigable river systems. Depo-
sition of sediment in rivers may decrease the water depth, making navigation difficult
or even impossible. But also erosion causes problems. For example, the Iffezheim
barrage on the Rhine River near Karlsruhe needs about 170, 000 m3 of sediments
per year added artificially to avoid erosion in the downstream river bed. The costs
amount to 5 million Euro each year [1].
Sediment transport modelling is an important tool for prediction as well as diagno-
sis. Computer models permit the simulation and prediction of environmental impacts
such as river discharge, sediment grain size and its distribution and bed and bank
formation.
In this thesis, the numerical sediment transport model SSIIM will be examined.
The numerical results are tested against the measurement data taken from a physical
experiment. A new algorithm to calculate the incipient motion of sediment particles
on generalized sloping fluvial beds will be introduced as this is indispensable for a
realistic simulation. This algorithm can adequately describe the threshold of sediment
motion on a combined transverse and longitudinal sloping bed Dey [5]. Numerical
parameters will be varied to test the sensitivity to the model results.
The results will be compared to measurements of the experimental run in order to
determine the capabilities and limitations of the numerical model SSIIM. Advantages
and disadvantages of SSIIM with respect to the reproduction of the physical processes
and practical applications will be discussed.
The implementation of the additional sediment transport algorithm with the exist-
ing CFD-code constitutes an improvement with regard to the prediction of sediment
transport problems. SSIIM is able to make more reliable predictions and is therefore
a useful tool for river, environmental and sedimentation engineering.
2 Fundamentals
2.1 Governing Equations for Water Flow Calculation
Conservation of Mass The law of conservation of mass states that mass can neither
be destroyed nor created, but it can only be transformed by physical, chemical or
biological processes. All mass flow rates into a control volume through its control
surface is equal to all mass flow rates out of the control volume plus the time change
in mass inside the control volume. Written out fully in cartesian coordinates

∂% ∂(%u) ∂(%v) ∂(%w)


+ + + = 0, (2.1)
∂t ∂x ∂y ∂z
or generalizing to three dimensions, it can be written in Einsteinian notation of
repeated indices
∂% ∂(%ui )
+ = 0, (2.2)
∂t ∂xi
where % is the fluid density, t the time, ui the fluid velocity vector and xi is the position
vector. The index i replaces all three spatial dimensions and must be summed over
all directions.

Conservation of Momentum The following Navier-Stokes equation represents the dif-


ferential form of the law of conservation of momentum. It describes the motion of a
flow particle at any time at any given position in the flow field. The equation can
be found in many textbooks [19, 22]. Using the Einsteinian notation and introduc-
ing an additional index j, which again represents all three spatial dimensions, the
Navier-Stokes equation can be written as

∂(%ui ) ∂(%ui uj ) ∂p ∂τij


+ =− + + f i, (2.3)
∂t ∂xi ∂xi ∂xi
where p is the pressure and τij are the viscous stresses. Source and sink terms are
summerized in fi . These are acceleration due to gravity, buoyancy and external forces
by hydraulic structures, wave stresses, etc.

Transport Equation In a general flow, transport of solutes, salinity or heat are due
to advection and diffusion
∂C ∂ ∂2C ∂2C
+ (ui C) = Em,i 2 + Et,i 2 ± KC, (2.4)
∂t ∂xi ∂xi ∂xi

where C is the concentration of the dissolved substance or heat, Em is the molecular,


Et the turbulent diffusion coefficient and KC represents a first-order reaction process.
It should be mentioned that the diffusion coefficients have a constant value in each
2 Fundamentals 3

direction, but are not homogeneous with respect to their spatial direction and are
therefore non-isotropic. A derivation of (2.4) by using Fick’s law is given in Fischer
et al. [7] and Socolofsky and Jirka [24].

Equation of State For sea water containing dissolved salt, the density % is a function
of temperature T , salinity S and the pressure p

% = f (T, S, p). (2.5)

2.2 Assumptions and Approximations


The following assumptions and approximations are applied in this study.

Incompressibility In river, environmental and sedimentation engineering, water flows


at low mach values. The effects of compressibility can be neglected.

Equation of state At a given temperature T and salinity S the fluid density is con-
stant, assuming an incompressible fluid and small variations in topography. Also the
dynamic viscosity µ is of constant value within the flow field.

Newtonian fluid For Newtonian isotropic fluids –like water considered in this study–
the viscous stresses τij in (2.3) depend on the velocity gradients and can be formulated
as [22]  
∂ui ∂uj 2 ∂uk
τij = µ + − δij , (2.6)
∂xj ∂xi 3 ∂xk
where µ is a proportional factor called dynamic viscosity. The kinematic viscosity
of the fluid, which is defined as the ratio of dynamic viscosity and the density, ν =
µ
%
, depends also on the temperature T and pressure p. In the SSIIM software the
kinematic viscosity ν is hard–coded and can not be changed. It is equivalent to water
2
at 20°C, i.e. ν = 1.006 ·10−6 ms .

Reynolds decomposition With regard to turbulent flows, the instantaneous velocity


ui (xi , t) and pressure p(xi , t) can be decomposed in time–averaged mean velocity
plus turbulent fluctuation, respectively in mean pressure plus turbulent fluctuation.
Therefore, the resulting flow has a velocity and pressure such as

ui (xi , t) = ui (xi ) + u0i (xi , t), (2.7a)


p(xi , t) = p(xi ) + p0 (xi , t), (2.7b)

where the overbar indicates the time–averaged mean and the prime the turbulent
fluctuation quantities.
2 Fundamentals 4

The Reynolds-averaged equations are obtained if the Reynolds decomposition is


substituted into the basic equations (2.2) and (2.3). The resulting equations contain
further unknowns, namely the Reynolds stresses. These are new statistical correla-
tions ui uj between different fluctuation velocities. A turbulence closure model de-
scribes these quantities, which is introduced in more detail in chapter 2.4.
Analogously, the other instantaneous quantities are also separated into their mean
parts and a fluctuation from the mean. C(xi , t) = C(xi ) + C 0 (xi , t) is substituted
in (2.4). In natural streams, the magnitude for turbulent diffusion is several orders
greater than molecular diffusion and can safely be removed from (2.4) (cf. Socolofsky
and Jirka [24]). In this thesis, no reaction processes are considered and K is set at
zero.

2.3 Governing Equations in Simplified Form


Invoking the restrictive conditions in chapter 2.2, the equations (2.2), (2.3) and (2.4)
can be reduced to the following:

Conservation of Mass
∂ui
=0 (2.8)
∂xi

Momentum Equations or Reynolds Equations


 
∂ui ∂ui 1 ∂p 1 ∂ ∂ui fi
+ uj =− + µ − %0 u0i u0j + (2.9)
∂t ∂xj %0 ∂xi %0 ∂xj ∂xj %0

where %0 is the reference density.

Transport Equation
∂C ∂ ∂2C
+ (ui C) = Ei 2 (2.10)
∂t ∂xi ∂xi

Equation of state
% = %0 = const. (2.11)
2.4 Turbulence Modelling
2.4.1 Remarks
Most flows which occur in practical river, environmental and sedimentation engineer-
ing are turbulent, which means irregular fluctuation is superimposed on the main
motion. Turbulence involves disorder, is irreproducable in detail, performs efficient
mixing and transport and vorticity is irregularly distributed in all three spatial di-
mensions. The large eddies are the main carriers of kinetic energy in the fluctuations.
They obtain their energy from the mean motion. In a cascade process they decay
and pass on their energy to several small eddies. At these smaller scales of motion,
energy is dissipated by the action of viscosity, i.e. a transfer from mechanical en-
ergy to internal energy. This observation is of great importance for the numerical
modelling of turbulence. To resolve these effects, a very fine grid is required. The
grid represents the investigated natural hydraulic system in a numerical model. For
that purpose the domain must be divided into cells. Because of the huge amount of
data involved, it is only possible today to resolve the small-scale motions by direct
numerical simulation (DNS) and the use of super-computers, which is not feasible for
practical engineering purposes. For this reason the effects of the smaller-scale motion
on the main flow must be modelled.

2.4.2 Reynold Stress Terms


As already indicated in chapter 2.2, in computing a turbulent flow it is useful to
decompose the instantaneous motion into mean and fluctuation velocity. The same
is done with the instantaneous pressure (cf. eqn. 2.7b). Substituting (2.7) into the
Navier-Stokes Equation given in chapter (2.1) leads to the Reynolds-Equations (2.8)
and (2.9). The only difference between these equations and (2.3) is the fluctuation
part; the term −%0 u0i u0j on the right side of (2.9). They cause an enhanced turbulent
momentum transport compared to the laminar flows and thus act like stresses. These
terms are called Reynold stresses and are unknown. In order to be able to calculate
turbulent flow, the system of equation must be supplemented by additional equations
for these occurring unknowns. With regard to Schlichting and Gersten [22], trying to
add new terms for the unknowns will unfortunately lead to further unknowns in form
of higher correlations, which would mean that closure of the system of equations
is never achieved. This is the so–called ’closure problem’ [22]. Therefore, model
equations must be used to connect the Reynold stresses to quantities of the mean
flow.

2.4.3 Boussinesq’s Approximation and Eddy–Viscosity


Boussinesq’s eddy–viscosity concept suggested that in analogy to Newton’s law of
friction, the turbulent stresses are proportional to the mean velocity gradients. For
2 Fundamentals 6

general flow situations, the turbulent stresses may be expressed as [19]


 
∂ui ∂uj 2
τij = −%ui uj = µt + − %kδij (2.12)
∂xj ∂xi 3

It should be pointed out that µt (x, y, z, t) is not a physical property and not of
constant value, but rather a function of position and time, i.e. it depends on the flow
under consideration. Consequently, the distribution of µt across the flow field must
be estimated.
From dimensional analysis [19], the eddy–viscosity is proportional to three param-
eters, namely
µt ∝ %V̂ L, (2.13)
where V̂ is a velocity-scale and L characterizes the large-scale turbulent motion. The
distribution of these scales can be approximated reasonably well in many flows. The
calculation of µt from the given parameters is described in chapter 2.4.5.

2.4.4 Eddy–diffusivity Concept


The relation between the eddy–viscosity and the turbulent diffusion coefficient Et is
described by the following equation
µt
Et = , (2.14)
%σt
where σt is the turbulent Schmidt–number.

2.4.5 The k– Turbulence Model


The k– turbulence model is a two–equation model. As mentioned in chapter 2.4.2,
in order to close the system of equations, additional equations must be introduced.
These equations receive their information from the mean flow. From (2.13), the
parameters V̂ and L are the only unknowns. If these parameters can be expressed in
terms of the main flow, the system of equation will then be closed.
As Rodi [19] writes, the balance of the turbulent kinetic energy is defined as
1 1
k = (u0i u0i ) = (u02 + v 02 + w02 ), (2.15)
2 2
where u0 , v 0 and w0 are the fluctuating velocities in all three dimensions and k is the
kinetic energy of turbulent fluctuations. Obtaining a form of velocity scale, V̂ can be
written as √
V̂ ∝ k (2.16)
2 Fundamentals 7

Table 1: The values of the k– turbulence model [19].


cµ c1 c2 σk σ
0.09 1.44 1.92 1.0 1.3

Dimensional considerations also allow defining a dissipation rate of turbulent en-


ergy given by
3
k2 µ ∂u0i ∂u0i
∝ ∝ . (2.17)
l % ∂xj ∂xj
Substituting (2.15) and (2.17) into (2.13), the k– model calculates the viscosity
µt by
k2
µt = f (k, ) = cµ % , (2.18)

where cµ is an empirical constant.
In order to estimate the eddy–viscosity µt from (2.18), the distribution of k and
 across the flow field must be known. The distribution can be calculated by the
means of transport equations k and . The modelled form according to Rodi [19] for
the turbulent kinetic energy k is
    
∂k ∂k ∂ µt ∂k ∂ui ∂ui ∂uj
% + %ui = + µt + − % (2.19)
∂t ∂xi ∂xi σk ∂xi ∂xj ∂xj ∂xi

and the dissipation of k is denoted  and modelled as

2
    
∂ ∂ ∂ µt ∂  ∂ui ∂ui ∂uj
% + %ui = + c1 % µt + − c2 % . (2.20)
∂t ∂xi ∂xi σ ∂xi k ∂xj ∂xj ∂xi k

The empirical constants given by Rodi [19] are shown in Table 1.


The equations (2.19) and (2.20) in conjunction with (2.12), (2.9) and (2.10) describe
completely the k– model and allow the system of equations to be closed. The
empirical constants in Table 1 are determined by experiments. They are universal
for a wide range of flow situations. For this reason, the k– model is de–facto–
standard in all industrial applications. Furthermore, the k– model shows a very
stabil behaviour in numerical applications.
3 Sediment Transport Mechanism
3.1 Shields Parameter
The river bed experiences shear stresses due to the flow of water. The classical
deviation of the shear stress balances all forces acting on an element of the river.
Taking all forces into account, produces the following equation (Zanke [29])

τ = %w gIe h, (3.1)

where τ is the shear stress, %w the density of water, g is acceleration due to gravity,
Ie represents the energy slope and h is the flow depth.
A particle motion may be initiated by shear stress τ exceeding a critical value of
the bed–shear stress τcrit , which is a function of sediment and hydraulic parameters
[29],
τcrit = f (%s , %w , D, h, ν, g, I0 ), (3.2)
where %s the density of a sediment particle, %w is the density of water and ν its
viscosity coefficient, D the mean particle size, h water depth and I0 represents the
channel slope. Since the flow under consideration is uniform and steady, the energy-
and channel slope are of equal value, Ie =I0 .
Shields [23] investigated the value of τcrit by the means of flume experiments, using
bed material of uniform size. Applying the Buckingham Π–Theorem (cf. Zanke [29]),
these seven basic parameters can be reduced to a set of four dimensionless parameters,
which are:

1. Specific density parameter


%s
s= , (3.3)
%w
2. depth-particle size ratio
h
, (3.4)
D
3. particle Reynolds–number
u∗ D
Re∗ = , (3.5)
ν
q
τ
where u∗ is the bed–shear velocity, defined as u∗ = %w
, and

4. particle Froude–number, defined as the ratio of the frictional load on the grain
to the gravitational force on the grain that resists movement

u2∗ τ
τ∗ = F r ∗ = = . (3.6)
(s − 1)gD (%s − %w )gD
3 Sediment Transport Mechanism 9

The empirical dimensionless parameter F r∗ can be taken from the Shields’ diagram
found in many textbooks [9, 10, 23], where F r∗ is a function of Re∗ . The Shields’
relationship between dimensionless shear stress F r∗ (or Shields parameter τ∗ ) and
particle Reynolds–number Re∗ can be applied for predicting whether or not a particle
will move. The dimensionless value of τ∗,crit = 0.03 to 0.06 is often used as a limit
for bed protection. If the value of bed–shear τ exceeds the critical value τcrit , motion
will be initiated, expressed by following dimensionless expression

τ∗ > τ∗,crit . (3.7)

3.2 Incipient Motion of Sediment Particles on Generalized


Sloping Fluvial Beds
The above chapter describes the forces induced by the flow acting on sediment parti-
cles and its prediction of initial movement. The Shields’ curve has been introduced.
Shields was the pioneer who described the incipient motion of uniform sediment par-
ticles. For his investigations he used two different types of rectangular flow channels.
On the one hand a wooden channel 0.8 m in width, on the other hand a channel made
out of glass and 0.4 m in width. This one could be inclined in a longitudinal direc-
tion [9]. Although Shields’ approach is widely used, it has disadvantages. Shields’
diagram is limited to bed slopes close to zero and can not be used in channels having
substantial bed slope [6]. It is also limited to stream flows without transverse incli-
nation.
In the further study of this thesis a flow in a trapezoidal channel bed is investigated.
The geometry is not rectangular because of the inclination of the inner and outer
bank. A modified form of the Shields’ approach has to be found, in order to take the
effects of a combined transverse and longitudinal sloping bed on a sediment particle
into account.
Dey [5] reports experimental results on incipient motion of non–cohesive, uniform
sediment particles under an unidirectional steady–uniform flow on generalized sloping
beds. He gives a formulation of particle stability by presenting the non–dimensional
ratio τ̃ of critical shear stress on a generalized sloping bed τb compared with the
critical shear stress on a horizontal bed τh , yielding
τb
τ̃ = . (3.8)
τh
From dimensional analysis it can be shown:

τ̃ = f (φ, θ, α), (3.9)

where φ is the angle of repose of sediment particles, θ is the stream–wise bed slope
and α is the transverse bed slope, as depicted in Fig. 1. Conducting experiments,
3 Sediment Transport Mechanism 10

(a) side view (b) cross–section

Figure 1: Definition sketch of angles θ and α

Dey [5] found from his data the following empirical equation for the estimation of τ̃
as  0.745  0.372
θ α
τ̃ = 0.954 1 − 1− . (3.10)
φ φ
The comparison of τ̃ obtained from eqn. (3.10) with the experimental data, a cor-
relation coefficient of value 0.97 can be computed [5]. This indicates an adequate
correspondence between experimental data and given empirical equation. Figure 2
shows variations of τ̃ with α calculated from (3.10) using θ = 5.368 e−04 = 0.03 ◦ and
φ = 0.6 = 34.4 ◦ , as will be used in the further study.
In the software SSIIM, the dimensionless critical shear stress τ∗,crit is computed
by the Shields’ curve, disregarding sloping bed. As in the following an inclined
trapezoidal channel is investigated, longitudinal and transverse sloping bed must be
taken into account. Therefore, the equation (3.10) is implemented to the SSIIM
program by the author. The C++ code for the equation is listed in appendix A.
Rearranging equation (3.8) yields

τh · τ̃ = τb . (3.11)

Since the critical shear stress for horizontal bed τh is the only unknown, the author
approximates this term with sufficient accuracy by the critical Shields’ value for
incipient motion τ∗,crit from the Shields’ curve. From the represented simple empirical
equation (3.10) it can be predicted whether or not a sediment particle on a generalized
sloping bed will move when exposed to flow forces.

3.3 Transport Modes and Particle Motion Modes


Sediment particles can be transported by the flow of water in the form of bed–load
and suspended load, depending on the size of the bed material particles and the flow
conditions [25]. There are 3 modes of particle motion: (1) Rolling and sliding motion
or both, (2) saltation motion and (3) suspended particle motion. van Rijn [25] defines
the transport of particles by rolling and saltating as bed–load transport, i.e. particles
have more or less contact with the bed surface. Furthermore, van Rijn [27] explains
3 Sediment Transport Mechanism 11

Deys empirical equation of critical shear stress ratio

1.4
 0.745  0.372
1.2 τ̃ = 0.954 1 − φθ 1 − αφ
1

τ̃ 0.8
0.6
0.4
0.2 φ = 34.4°, θ = 0.03°
0
-15 -10 -5 0 5 10 15 20 25 30
α in [deg]

Figure 2: Variation of τ̃ with α

suspended load as the part of the total which is moving without continuous contact
to the bed surface as a result of the agitation of the fluid turbulence. Therefore, a
particle only remains in suspension when the turbulent eddies have dominant vertical
velocity components exceeding the particle fall velocity.
The boundary between bed–load and suspended load is defined by van Rijn [25] as a
layer with a maximum thickness of about 10 particle diameters in which the particles
are transported as bed–load.

3.4 Computation of Transport Quantities


3.4.1 Bed–Load Transport
van Rijn [25] proposes an equation to predict a reliable estimate of the bed–load
transport in the particle range 200–2,000µm, given by

qb T 2.1
1.5
= 0.053 , (3.12)
((s − 1)g)2 D50 D∗0.3
where qb is the bed–load transport per unit width, D50 mean particle diameter, D∗
and T are the dimensionless particle diameter and the transport stage parameter,
respectively.
Since the shear velocity u∗ appears both in the Shields parameter (3.6) and in the
particle Reynolds–number (3.5), Julien [10] shows that the dimensionless particle
diameter D∗ can be derived by eliminating the shear velocity u∗ from the particle
3 Sediment Transport Mechanism 12

Figure 3: Dimensionless particle motion diagram, Julien [10]

Reynolds–number by dividing by F r∗ ,

2
! 13
Re∗
D∗ = , (3.13)
F r∗

leading to
  13
(s − 1)g
D∗ = D50 . (3.14)
ν2
Thus, the abscissa of the Shields’ curve can be replaced by the dimensionless particle
diameter, resulting in Fig. 3. The dimensionless transport parameter is obtained by

(u0∗ )2 − (u∗,crit )2
T = , (3.15)
(u∗,crit )2

where u0∗ is the bed–shear velocity related to grains and u∗,crit is the critical bed–
shear velocity according to Shields [23]. This parameter expresses the mobility of the
particles in terms of the stage of movement relative to the critical stage for initiation
of motion.
From the aforementioned concept, the bed–load transport rate qb is calculated by
the help of only two dimensionless parameters.
3 Sediment Transport Mechanism 13

3.4.2 Suspended Load


Sediment particles are transported as suspended load, if the upward turbulent fluid
forces exceed the downward gravitational forces. van Rijn [26] uses the suspension
parameter Z by Rouse [20], expressing the influence of the above–mentioned forces,
yielding
ws
Z= , (3.16)
βκu∗
where Z is the suspension parameter, ws fall velocity of suspended sediment particle,
β a coefficient related to diffusion of sediment particles, κ the constant of von Karman
and u∗ is the bed–shear velocity. A concentration profile can be drawn, using the
Rouse [20] expression,
 Z
c(y) h−y a
= , (3.17)
ca y h−a
where a is the reference level with its reference sediment concentration ca . Fig. 4
depicts some sediment concentration profiles for selected suspension parameters Z.
van Rijn [26] shows that the function for the bed–load concentration as proposed
in [25] can also be used to compute the reference concentration for suspended load.
The reference concentration is expressed as

D50 T 1.5
ca = 0.015 . (3.18)
a D∗0.3

Dividing the water body into cells, the concentration of suspended sediments in
the center of each cell can be computed by eqn. (3.17). In conjunction with the
velocity, the transport of sediments through each cell can be calculated by (2.10) and
is carried out by SSIIM.
3 Sediment Transport Mechanism 14

1
Z = 0.01
0.8
(y − a)/(h − a)

Z = 0.1
0.6

0.4

0.2
Z=1
Z=5
0
0 0.2 0.4 0.6 0.8 1
c/ca
Figure 4: Vertical concentration profile according to (3.17)
4 Flow in Curved Open Channels
4.1 Secondary Flows
Secondary flows are defined as currents that occur in the plane normal to the primary
flow direction. Their velocities are typically one order of magnitude smaller than the
bulk primary velocity [2]. Ciray [4] gives a rather functional definition on secondary
flows as follows: ”When the magnitude of the vector composed by any two compo-
nents [of] the local velocity vector[s] in a three dimensional flow is small compared
with the magnitude of the third component, the latter forms the main flow whereas
the remaining two form the secondary.”
Nevertheless, secondary currents strongly influence the velocity pattern of the
stream. Different velocity patterns in cross–streamwise and vertical direction of the
flow will be found. Secondary currents combined with the primary flow produce a
spiral flow rotating around the primary flow direction and therefore affect the pro-
cesses of flow resistance, sediment transport, bed and bank erosion and development
of channel morphology [2].
Prandtl et al. [18] distinguishes three modes of secondary flows by the releasing
forces. In the present study the secondary flow of the first type will be discussed as
it is of interest to flows under consideration. These secondary flows of the first mode
are provoked by differences in centrifugal forces.

4.2 Formation of Secondary Flows


Rozovskii [21] gives an explanation of the formation of secondary currents. He con-
siders an open stream that moves first in a straight channel and then passes into a
curvature. From currents being curved on a plane, the appearance of a transverse
inclination of the free surface can be observed. To evaluate approximately the extent
of this inclination Ir , centrifugal-, pressure- and friction forces are balanced, yielding

dh v2 τrb
Ir = = βo cm + , (4.1)
dr grc %gh
where h is the water depth, rc is the radius at the center of the bend, τrb the radial
component of the friction stress at the bottom and vcm is the mean velocity. Since
the vertical velocity is unevenly distributed over the depth, an averaging factor β0 is
introduced to estimate the average velocity in streamwise direction.
τrb
As the friction term %gh on the right side of (4.1) is in many cases of small quantity,
it can be omitted. Thus the approximate equation for the inclination of the water
surface is obtained as
2
dh vcm
Ir = = βo . (4.2)
dr grc
4 Flow in Curved Open Channels 16

Figure 5: Definition sketch of transverse inclination of surface and transverse circu-


lation at bend in channel [21], a–section AB; b–plan

Eqn. (4.2) shows that at any bending of a stream on a plane, a transverse inclination
inevitable appears. The level of the free surface at the inner bank, which is nearest to
the center of the turn, will decrease. The reverse at the outer bank, which is farthest
from the center of the turn, will have an increasing water level. To conform to the
technical literature, the inner bank will be termed in the rest of this text the ”convex
bank”, and the outer bank the ”concave bank”.
Another characteristic property of a flow passing a bend is the presence of trans-
verse circulation. Rozovskii [21] considers a volume element at a distance z from the
bottom, which moves along the circular trajectories of an equal radius r as shown in
Figure 5. With the absence of radial velocity components, the friction force is taken
as zero. Again, balancing all forces acting on the volume element leads to
dh v2
Ir = = . (4.3)
dr gr
As (4.2) and (4.3) describe the identical phenomena, Ir must be equal in both cases.
This is achieved only in two cases: first, in a straight channel, where r = ∞, and
second, where v 2 of (4.3) is of constant value and equals β0 vcm
2
of (4.2), i.e. a uniform
distribution of the velocity along the vertical. It is well known that due to friction
at the boundaries of the stream, the velocity at the bottom and at the walls are
minimum. Close to the water surface they are maximum. The velocities therefore
can not be of constant value, but deviate from the mean, depending on their position
in the flow. Owing to variations in vertical velocity, the averaging factor β0 was
introduced in (4.1), so the term v 2 can not be of constant value. For particles moving
near the upper surface with a velocity faster than the mean of the flow, centrifugal
forces will be greater than the pressure forces. These particles will consequently
be displaced in a radial direction towards the concave bank. For the same reason,
particles in the lower part of the stream will move toward the convex bank. Particles
reaching the wall will move in a down- or upward direction, producing a circulation.
4 Flow in Curved Open Channels 17

To conclude, simultaneously with the radial displacement, a vertical velocity com-


ponent also appears, i.e. a flow in the transverse direction, also called transverse
circulation. It is obvious from the above–mentioned that in bending streams, a
transverse circulation is superimposed on the main flow direction and a flow passing
a bend is of a complex spatial character.

4.3 Distribution of Longitudinal Velocity


Owing to basic flow superimposed by a transverse circulation, the velocity structure
of a stream within a bend is subjected to strong changes. A stream which moves
first in a straight, uniform channel will have its maximum longitudinal velocity in
the middle of the channel width. Entering the bend, the maximum velocity moves
towards the convex bank [21]. As aforementioned, by virtue of centrifugal forces
acting on fluid elements in the same way, but varying in its magnitude at different
depths, a transverse slope on the free surface appears. The level at the convex bank
lowers, while rising at the concave bank. The amount of transverse slope can be
determined with sufficient accuracy from (4.2), taking β0 = 1.
By lowering the free surface at the convex bank, the velocity is increased due to a
smaller flow area perpendicular to the main flow direction. This is stated by the law
of conservation of mass in (2.1). If the level of the free surface rises, a simultaneous
decrease in velocity must take place.
If the secondary flow fully develops in the bend, an interchange of momentum
between horizontal currents takes place by reason of transverse circulation. As a
consequence, the maximum velocity moves from the convex to the concave bank [21].

4.4 Distribution of Vertical Velocity


A change in current velocity is not only limited to longitudinal direction. Also in the
vertical direction, a movement of the velocity maximum can be observed. It tends
to compensate the velocity differences in the vertical direction, so that the maximum
velocity in the longitudinal direction not only moves from the convex to the concave
bank, but also from the near-surface to the near-bottom. It is also worth mentioning
that the typical logarithmic velocity profile as found in straight channels does not
hold true in bends [13].

4.5 Effects on Sediment Transport


From a morphological point of view, high bed–shear stresses, high velocities and its
inhomogenity of distribution exert great influence on the water and material transport
properties of a stream.
Increasing velocity near the bed in the direction of primary flow are responsible for
enhanced sediment transport in bends. The direction of sediment motion follows the
4 Flow in Curved Open Channels 18

resulting force of primary and transverse shear stress components. The secondary flow
near the bed leads to erosion at the concave bank and to deposition at the convex
bank, resulting in a scour and a point bar. Owing to these effects, the transverse
bed slope increases and stability of bed material decreases as threshold conditions
for incipient motion changes (cf. chapter 3.2). The morphological processes show
dynamical behaviour. Secondary flows, with their complex structure in time and
space, need three–dimensional models if investigated in more detail.
5 Numerical Model SSIIM 19

5 Numerical Model SSIIM


The study of fluid mechanical problems like hydraulics and sedimentation processes
is carried out by both the use of physical model techniques and numerical models.
Physical model tests or experiments can be run in compliance with laws of similarity
(cf. Kobus [11]). They ensure that dominating forces of the prototype are reproduced
in the model. To meet these laws of similarity, the down-scale factor of the model is
limited. Applying physical model tests is intensive in costs and time.
The advantage of numerical models is that the laws of similarity are met at any time
for any configuration of flow. Also variations in geometry can easily be calculated.
Parameter studies can be realized in a reasonable time at fairly low costs.
Throughout this thesis, the numerical model SSIIM, an acronym for S ediment
S imulation I n I ntakes with M ultiblock option, has been used. It was developed
by Olsen [16] at the Department of Hydraulic Engineering at the Norwegian Uni-
versity of Science and Technology. It solves the Reynolds-averaged Navier-Stokes
equations in three dimensions to compute water flow using the finite volume ap-
proach as discretization method. The control volume is divided into finite volumina
and the governing equations in simplified form (cf. (2.8) and (2.9)) are solved for
each cell and each variable. The k– turbulence model by Rodi [19] computes the
Reynold stress terms in order to close the system of equations. The SIMPLE method
by Patankar [17] computes the pressure term.
Besides the water flow calculation, the software code is supplied with an extra
feature to represent sediment movements and other water quality parameters. The
governing equation to predict the change in concentration is the convection–diffusion
equation (2.10). From (3.12) SSIIM calculates the concentration of all elements
situated closest to the river–bed. The flux of mass through the element surfaces is
estimated by solving (2.10). If sediment continuity is not achieved, i.e. net flux into
the element does not equal net flux out of the element, sediment deposition or erosion
occurs. The geometry change of the bed cell can then be predicted. Consequently,
the grid pattern changes in response to alluvial processes and the flow calculation
has to restart until convergence is achieved.
6 Description of Physical Model
6.1 Geometry and Flow Parameters
The physical model was originally used for experimental studies on longitudinal dis-
persion tests at the University of Sheffield. Besides this main research, at different
locations in the channel, measurements of the geometry have been registered and
repeated at different time intervals, in order to document the evolution of the bed
topography. Collected data [8]are unpublished and have been kindly contributed to
the author. These data are of great interest. It is indispensable to compare results
predicted by SSIIM with experimental data to validate the sediment transport fea-
ture of the program. Because of the availability of experimental data, this case is
chosen to test the sediment transport prediction of SSIIM.
Fig. 7 shows the flow area, which will be termed ”channel” in the further text. It is
of trapezoidal shape with Bsur = 0.3 m in width at the surface. Both, the convex and
concave bank are inclined at an angle of α = 30° against the horizontal. Running the
experiment on bank–full discharge of Q = 0.001 m3/s, the water adjusts to h = 0.05 m
of depth along the whole channel, owing to a uniform and steady flow. Measurements
have been taken after a duration time of ∆t = 4 h.
The modelled case is of a meandering channel in a flume of uniform sand. Initially, the
channel of constant cross–section geometry excavates the sand and forms a sequence of
S–shapes, as depicted in Figure 7. Precisely, two shifted circles of 234° are facing each
other with a distance of 0.250 m from their centers. Both circle parts meet exactly
between their centers, denoted as center line (CL). Figure 6 details the situation seen
from above.
Water passing all bends covers a distance of 27 m, whereas the flume is 12 m in length
and 1.20 m in width. The channel slope amounts to I0 = 5.368 e−04 with respect to
the developed length. In appendix C, Fig. 20 shows the original channel shape before
water discharge is imposed. The flow regime in open channels is defined by the
Reynolds–number:
um DH
Re = , (6.1)
ν
where um is the mean flow velocity, DH is the hydraulic diameter and ν represents
the kinematic viscosity. The hydraulic diameter DH is defined as
A
DH = 4 , (6.2)
Pw
where A is the cross–sectional flow area and Pw is the wetted perimeter. With
6 Description of Physical Model 21

Figure 6: Definition sketch of channel shape seen from above

Figure 7: Definition sketch of channel shape seen from side


6 Description of Physical Model 22

h = 0.05 m, α = 30° and Bb = 0.1268 m,

A= h(Bb + h cot α) = 0.01067 m2 (6.3a)


2h
and Pw = Bb + = 0.3268 m (6.3b)
sin α
resulting in DH = = 0.1306 m, (6.3c)

the mean Reynolds–number is calculated to

Re ≈ 12.000 > 2000, (6.4)

with um = Q/A = 0.093721 m/s. The flow under consideration is therefore turbulent.
Introducing the Froude–number as
um
Fr = p , (6.5)
g (A/Bsur )

where Bsur is the width of the water surface, yields in

F r = 0.16 < 1, (6.6)

low kinetic energy or subcritical flow.

6.2 Sediment Parameters


The geometric standard deviation σg of the particle size distribution given by Dey
[5] r r
d84 0.850 mm
σg = = = 1.2 < 1.4, (6.7)
d16 0.600 mm
is less than 1.4 for uniform sediment, where the variables d84 and d16 denote the
grain size sieve where 84% and 16% of the sediment material is finer. Fig. 8 shows
the partice size distribution of the sediment. A semi-log-scale is used. As the same
sand material is distributed all over the flume, friction due to grain size is of constant
value along the channel.
6 Description of Physical Model 23

1
cumulative passing
retained sieve box

0.84
0.8
fraction finer

0.6

0.4

0.2
0.16

0
0.1 1
particle size d in mm
Figure 8: Particle size distribution d of sand material
7 Numerical Simulation
7.1 Represented Domain
In Computational Fluid Dynamics (CFD) the physical domain of a flow process has
to be described in the numerical model. To represent the complex three–dimensional
geometry of the experiment mentioned in chapter 6, a structured, non-orthogonal grid
is used. Simulating all twelve bends of the experiment, the numerical grid consists of
623 in longitudinal and 16 grid cells in transverse direction. In the vertical direction,
8 layers are chosen, as depicted in Fig. 10. After having carried out first simulation
runs, similar flow and sediment patterns could be observed at different but equivalent
positions. In order to save computation time, simulating only four complete bends
instead of all twelve seems to be sufficient. The number of grid cells in the longitudinal
direction is reduced to 260 cells. Thus, the longitudinal grid size in the middle of the
channel amounts to ∆x = 0.043 m and in the transverse direction to ∆y = 0.01875 m,
while the total water depth of 0.05 m is resolved in 8 evenly spaced vertical layers,
each 0.00625 m high. Fig. 9 shows the mesh layout, seen from above. The flow is
from left to right.

7.2 Presentation of Water Flow Results


The results of the water flow calculation by the model SSIIM are presented. The
basic flow configuration is chosen, as will be introduced in chapter 7.2.1. In this
section, some characteristics of the investigated flow in bends with a rigid bottom
are presented and special attention is given to secondary flow and to distribution of
longitudinal velocity. After having carried out first water flow calculations, two spiral
motions could not be observed, using the mesh as described in chapter 7.1. Refining
the mesh in each spatial direction, it emerges that both spiral motions are resolved.
As a consequence, refining the mesh leads to extended calculation time. In order
to compensate this effect, only the first two bends are represented in the numerical
model, but these are sufficient to characterize the water flow. The second spiral
motion does not strongly affect the sediment transport, and carrying out further
sediment calculations on the original grid as introduced in 7.1, will produce results

Figure 9: Definition sketch of mesh, seen from above


7 Numerical Simulation 25

of adequate accuracy in a reasonable time. From the water flow pattern, qualitative
predictions as to sediment movement can be made.

7.2.1 Basic Flow Configuration


The numerical domain is located in such a way that the center of the circular bends
have a constant parallel offset with respect to the x–axis of alternating ±0.125 m. A
steady discharge of Q = 0.001 m3/s is constrained to the numerical model in order to
be consistent with the experimental setup.
The water velocities and turbulence parameters are prescribed as a Dirichlet bound-
ary condition at the upstream inlet. Zero gradient, or Neumann condition, is used for
all the variables at the downstream outlet. At the water surface, which is modelled
as a rigid lid, Neumann condition is used for all variables except the turbulent kinetic
energy k, which is set to zero. The logarithmic wall law introduced by Schlichting
and Gersten [22] for rough boundaries is applied.
For generating the water surface for the subsequent 3D water flow calculation, a
1D backwater calculation is used. Therefore a roughness coefficient must be passed
to the SSIIM software. The bed and wall roughness is chosen as a constant Strickler
1
value kst = 34.15 m /3/s. In conjunction with uniform water depth of 0.05 m and an
uniform channel slope of I0 = 5.368 e−04 , this configuration leads to a uniform steady
flow as conducted in the experiment.
All numerical calculations are carried out by a second-order upstream scheme. It
is of second-order accuracy and reduces false diffusion. Owing to deflection existing
between flow direction and alignment of grid cells, physically present discontinuities
become ’smeared out’ by the numerical procedure. It has the same effect as molecular
and turbulent diffusion processes and is therefore denoted as false diffusion.

7.2.2 Secondary Flow


If a current deviates from its primary direction due to bends, the fluid elements follow
a curvilinear trajectory. Centrifugal forces occur as described in chapter 4. Compen-
sating the centrifugal forces of different magnitudes, secondary flow is created.
Fig. 11 shows a velocity vector plot at cross-section no. F243 situated between
the the upper part of bend no. 1 and the lower part of bend no. 2. Seen looking
upstream, the left shows the convex bank, while the concave bank is represented by

Figure 10: Definition sketch of cross–section, seen looking upstream


7 Numerical Simulation 26

the right. The vector arrows represent the velocity components v and w in transversal
and vertical direction. The main secondary motion at the water surface is towards
the concave bank. There, the fluid elements move downwards. Reaching the bottom,
the fluid elements are vectored inwards to the convex bank. Consequently, the fluid
exhibits a cross-circulation. The combination of this cross-circulation and the major
flow direction results into fluid spiral motion.
The velocity vector pattern reveals a further phenomena. In subsequent bends,
besides the main spiral motion, a second, less extensive counterrotation appears.
This phenomena has been observed in experimental studies by Mosonyi and Götz
[14], if the ratio of width Bsur to waterdepth h is of lower value than Bsur/h < 10. In
this case the ratio is Bsur/h = 6 < 10.
The second spiral motion could only be resolved using a fine grid as mentioned above.

7.2.3 Distribution of Longitudinal Velocity


In order to present the longitudinal velocity distribution at chosen cross-sections, the
velocities in the vertical layers have been depth–averaged, resulting in one representa-
tive velocity of each water column. Furthermore, these velocities have been divided
by the mean stream–wise velocity um = Q/A, yielding a dimensionless value. The
numerical model uses cartesian velocity components as dependent variables, so the
longitudinal velocity components used in the later comparison must be obtained by
transformation into the corresponding directions at each cross-sections.
Fig. 13 depicts the depth–averaged velocity distribution at seven subsequent cross-
sections, starting with cross-section no. F235 and ending with F313. These are the
apexes of two consecutive bends. F274 represents the cross-over section of these
bends. Fig. 12 gives a more detailed reference to the position of the mentioned
cross-sections. Fig. 13 shows the characteristic radial shift of the tangential velocity
maximum. At the first apex the maximum is located close to the middle of its cross-
section, but still on its convex side. The position of the velocity peak diverges in the
same way as observed by Leschziner and Rodi [12], Meckel [13], Rozovskii [21]. They
found the maximum velocity at the convex bank, accompanied by a lowering of the
surface water level. Reaching the bend exit, the maximum gradually moves over to

Figure 11: Velocity vector plot at cross-section no. F243


7 Numerical Simulation 27

Figure 12: Position sketch of cross–sections


7 Numerical Simulation 28

the concave bank. On entering the subsequent bend, the velocity peak keeps moving
in its prevailing direction and approaches the convex bank, where it reaches its top
maximum velocity within the bank inclination.
It should be pointed out that the velocity can not be equated with the discharge.
Applying the law of conservation of mass (cf. eqn. (2.8)), the discharge Q equals
the product of velocity u and cross-sectional flow area A. Approaching the inclined
banks, the cross-sectional flow area A decreases. In spite of high velocity, the main
discharge Q will not take place at the bank location. As the velocity is important
for the sediment transport mechanism, the author emphasizes the representation of
velocity distribution instead of discharge distribution at the chosen cross-sections.
From the inclined banks the maximum velocity decreases and the peak moves back
to the middle of the cross-section. At the apex of the bend, the velocity distribution
is symmetrical to the center column of the preceding apex.

7.2.4 Distribution of Vertical Velocity


Besides the movement of the longitudinal velocity towards the concave bank, also
a vertical movement of the maximum is observed. The cross-section at the apexes
shows a longitudinal, depth–avaraged velocity maximum close to the middle of the
cross-section, as described in chapter 7.2.3, but having a local peak at the bottom and
the beginning of sloping bed, as shown in Fig. 14(a). In this area the isotaches are
compressed, indicating strong velocity gradients. Considering the subsequent cross-
sections, the core velocity moves towards the concave bank, while in the vertical
direction it drifts towards the center, as can be seen in Fig. 14(a) to Fig. 14(e). The
lines of equal velocity straighten out. Reaching the subsequent bend entrance, the
bulk velocity moves to the convex bank, resulting in compressed isotaches and high
velocity gradients, where the whole bank area is affected (Fig. 14(e)). At the second
apex, the core velocity moves to the bottom again and the high velocity gradients
remain. Meckel [13] observed the same action.

7.2.5 Stream Separation and Formation of Recirculation Zone


No attention has been paid to the reverse velocities occurring at the convex banks.
Rozovskii [21] states that in very acute bends there appears the phenomena of the
stream being drawn away from the walls and of eddy zone formation.
It is known that if the pressure along the flow increases, conditions arise for the
separation of the stream from the wall and for the appearance of eddy currents. On
the other hand, where there is a positive gradient of pressure along the flow, the
phenomena of separation is not observed. From Eulers equation, deceleration of fluid
elements is caused by a negative pressure gradient. Stream separation is expected. As
mentioned in chapter 7.2.3, the bulk velocity moves towards the concave bank, where
the fluid elements are accelerated. At the same time and for reason of conservation of
7 Numerical Simulation 29

bank middle bank


2.5
bend no.

F235
F250 × r r
r
2 F265 b r r
r
r
F274 M
?r ? ?
F283 r ?r ?
?r
dimensionless velocity scaled by um

F298 ?
××× ?r M M M M M
1.5 F313 × b b M M MM r
× ?b b
× b b M
×r ? M
? × b b
M M
× b M
?
r
× b b
b M r × b
? × b
× M b M
1 b ? r ×
M × ?b
× b ? r × b
M ×
b ? r × b
× M
? ×
b r × b
0.5 M
?
? r ?
×b M r ×
?
b ?M r
×
?M
?b M r r
?
0×?rb
M ? ? b× Mr r ? × ?rb
M
? b×r
M ?
?r ? Mr
r rb Mrb ×
Mb Mb M
×
×× ×
−0.5
5 10 15 20 25 30
transverse direction (cell number)
Figure 13: Longitudinal, depth–averaged velocity distribution
7 Numerical Simulation 30

(a) Cross–section No. F235

(b) Cross–section No. F250

(c) Cross–section No. F265

(d) Cross–section No. F274

Figure 14: Horizontal velocity distribution (begin)


7 Numerical Simulation 31

(e) Cross–section No. F283

(f) Cross–section No. F298

(g) Cross–section No. F313

Figure 14: Horizontal velocity distribution (end)


7 Numerical Simulation 32

mass, the fluid elements at the convex bank are decelerated and a negative pressure
gradient established, leading to a separation zone, expanding until the following apex.
Rozovskii [21] found from experimental results that formation of separation zones are
unlikely, even at sharp bends with a small ratio of center line curvature rc to width
B of rc/B = 1. In the investigated case the ratio equals rc/B = 0.916 < 1. To the
extent that stream separation is connected to the negative acceleration effect of fluid
elements at the side wall, Rozovskii [21] asserts that conditions for the appearance
of separation will also depend on the angle of bank slope. He concludes that the
possibility of separation increases, the gentler the bank slope is, or in short, the greater
the influence of friction of the stream against the bank. As in the case considered
the transverse bed slope angle equals α = 30 ◦ , the calculated flow separation zone is
likely, even in the case of a center line curvature radius to a width ratio close to one.

7.2.6 Concluding Remarks


Owing to a lack of measurement data, experimental verification for the water flow
calculation could not be obtained. Therefore, the absolute values of SSIIM results
presented in this chapter can hardly be judged as right or wrong. Rather, the aim
here is limited to identifying a trend. The author must depend on experimental runs
conducted and documented by researchers. The author has reviewed the literature
in order to verify the numerical results.
Since the flow pattern is sensitive to the presence of any preexisting circulation [2],
only experiments with subsequent bends qualify for drawing comparisons.
The numerical calculation of SSIIM seems to qualitatively correspond to the results
obtained in experiments by authors cited in this chapter, even considering differences
in geometry, roughness and flow parameters.
Proceeding from the assumption of physically correct numerical results, the water
flow calculation is the basis for the subsequent sediment transport calculation.

7.3 Presentation of Sediment Transport Results


The sediment transport simulation was carried out after having finished the calcula-
tion of the water flow field as the initial condition. The basic sediment configuration,
which will be referred to as the Base Case, is applied by the SSIIM program accord-
ing to chapter 7.3.1. Roughness parameter, grid size as well as the time step are
changed afterwards. The results are analyzed in order to determine the sensitivity
of the model to the specific changes and the capability of SSIIM to yield universally
valid predictions.
7 Numerical Simulation 33

7.3.1 Basic Sediment Configuration


In the numerical simulation of sediment transport both bed–load and suspended
load transport must be considered as described in chapter 3.4. Applying a rough
calculation of the suspension parameter, Z is estimated to Z ≈ 13. From Fig. 4
one can conclude that almost no sediment particles will be found in the water body
above the bed elements, if Z > 5. Furthermore, absence of suspended sediments
during the experimental run justifies the approximation of simulating the bed–load
transport only. The SSIIM program is assigned to ignore the suspended load, having
the positive effect of reducing computation time.
Several parameters must be passed to the SSIIM software and are kept constant
in all presented numerical simulations, unless otherwise noted.
The density of quartz minerals is typically ρs = 2650 kg/m3 and taken for calculation.
Most natural sediments have densities similar to that of quartz. Another property
of a sediment particle is its characteristic size d or D. The channel is formed in
uniform sand of d50 = 0.7 mm diameter. The fall velocity in vertical direction of a
sediment particle is mostly important for suspended particle motion. van Rijn [25]
assumes the bed–load particles are dominated by gravitational forces, while the effect
of turbulence on the overall trajectory is supposed to be of minor importance. But he
does not rule out particles jumping and leaving the bed surface for a short distance
and height. Therefore, the particle fall velocity must also be included even though
bed–load is the major transport mode. The fall velocity is taken from Chanson [3,
Table 7.3] to ws = 0.09 m/s and added to the second term on the left side of (2.10).
The angle of repose φ must be specified as it is important for the incipient motion of a
particle as discussed in chapter 3.2. It depends on the grain size d, density ρs and its
degree of saturation with water. In the vicinity of water a particle becomes unstable as
compared to a dry environment. Owing to buoyancy force, the density of a sediment
particle is reduced to submerged sediment density. In the basic configuration the
angle of repose is set to φ = 34.4 ◦ .
The bed elevation changes over a time period because of erosion and/or deposition.
The sediment calculation is transient and a time step ∂t has to be chosen. The time
step chosen must not be so large that changes in geometry cause the flow pattern
to change significantly. The shorter the time step ∂t, the more iterations have to
be calculated in order to carry out a simulation time of 4 h, which results in a long
computation time. A time step of ∂t = 20 s is deemed to be a satisfactory compromise
between numerical accuracy and computation time.
No sediment inflow is specified.

7.3.2 Default Case


A sediment transport calculation is carried out using the sediment transport approach
of van Rijn [25] in conjunction with the standard computation of the dimensionless
7 Numerical Simulation 34

Figure 15: Sediment transport results conducting default case

critical shear stress τ∗,crit by Shields [23]. No changes have been made in the software
code of SSIIM. Since no modification has been made running the numerical model,
the configuration is denoted as default case. Water flow and sediment parameters are
used as described in chapter 7.2.1 and 7.3.1, respectively.
The results achieved by conducting the default case is depicted in Fig. 15. SSIIM
predicts almost no changes in bed topography, which is not consistent with the ob-
servations made during the experimental run. Repeating the simulation and testing
increasing roughness height up to ks = 10d50 = 0.007 m, the sediment transport pre-
dictions get closer to measurements as a result of increasing shear stress. But one has
to be aware of an unphysical roughness height, which is not be found in common lit-
erature. Concluding from the conducted tests, SSIIM necessitates modification in the
critical shear stress τcrit computation. An additional algorithm must be implemented
in order to achieve realistic results, when decreasing the roughness height in value
back to physical range. Physically reasonable ranges of the roughness coefficient ks
are introduced in Wilson et al. [28].
As described in chapter 3.2, channel geometry influences the threshold condition
for incipient motion of uniform sediment particles as it reduces or increases the crit-
ical shear stress τcrit . Fig. 10 depicts a representative cross-section of the channel
geometry and shows that half of the river bed is inclined in a transversal direction.
It is assumed that the inclined banks will have a determining influence on incipient
motion and therefore on the sediment transport calculation. Eqn. (3.10) takes a lon-
gitudinal as well as a transversal sloping bed into account. Following calculations are
carried out by using eqn. (3.10), adjusting the critical shear stress τcrit .
7 Numerical Simulation 35

7.3.3 Base Case


Deposition SSIIM predicts deposition or point bar located at the convex apex of
all bends. Fig. 16(a) shows a representative bend. A formation of point bar at the
convex apex is also documented by the measurements of the experimental run, as
seen in Fig. 16(b). The visualized measurements do not exactly take the same shape
as the numerical predictions, because of a not very detailed ground mapping in the
adjacency of the point bar. Fig. 21 in appendix C shows a detailed top view of the
point bar. Its pattern corresponds quite well to the predictions made by SSIIM.
The location of the deposition may be explained by the following: Considering the
preceding water flow calculation, strong velocity gradients are located at the convex
apex. The flow velocity is reduced by the reversed velocity vectors of the eddy zone.
Fig. 13 shows the sudden change in velocity located at the apex. According to Shields
[23], low velocity leads to no conditions for particle motion as critical shear stress is
not exceeded. Sediment particles accumulate at zones of low velocity and have the
effect of increasing bed level. A point bar forms in this case.

Erosion Sediment material erodes from upstream of the point bar. The particles dis-
lodged upstream are the source of the particles deposited at the convex bank of the
next apex, which is described in the paragraph above. Consequently, the position
where particles enter motion is simultaneously the position where the shear stress
applied on the bed material exceeds the critical shear stress. From a water flow
calculation point of view, erosion seems to be predictable at zones of high velocity.
Again, by the use of the longitudinal velocity distribution depicted in Fig. 13, high
velocity coincides with the area of high erosion.
Calculations achieved by SSIIM and measurements show the same characteristic
trend. Interpreting the measurements, erosion takes place upstream of the bend
apex, but with a smaller expansion than predicted by SSIIM, which overpredicts
the erosion. The erosion zone at the concave bank downstream of the apex can not
be observed in Fig. 21. It is assumed that the appearance of the erosion zone in
the measurement data arises from imprecisions of the gauging device. The gauging
equipment is not able to reproduce elevation differences in the required resolution.
Nevertheless, the longitudinal position of the deposition and erosion zone is identical
in both cases.

7.3.4 Parameter Variation


Bed roughness The equivalent sand roughness is determined to ks = 0.0021 m. As
the roughness height influences the shear stress τ , the threshold condition of incipient
particle motion is affected. It is assumed that increasing the roughness height will
lead to increasing erosion and vice versa. Both simulations of ks = 0.0007 m and
ks = 0.0035 m are carried out. Fig. 17(a) and Fig. 17(b) show the results. Besides
7 Numerical Simulation 36

(a) Computed by SSIIM, base case (b) Experiment

Figure 16: Comparison of bed elevation changes

the variation of absolute values in deposition and erosion as assumed by the author,
movement of the point bar position is observed. Applying a reduced ks –value com-
pared to the base case, the point bar moves upstream in the direction of the erosion
zone. Increasing the hydraulic roughness, the opposite effect is observed. The point
bar travels in flow direction downstream and becomes stretched. A second scour
point next to the deposition forms.
The enlargement of the deposition zone is explained by the simultaneous quantitative
increment of erosion. The sediment particles try to deposit themselves at the apex for
the reason mentioned in paragraph Base Case above. During the bed changes, the
bed slope may increase to a level above the angle of repose φ. The sediment particles
are not able to drop in this cell and are forced to move sideways to a neighbouring
cell. A stretching of the deposition zone occurs, as long as no state of equilibrium is
achieved. This effect is supported by the bed sloping algorithm implemented by the
author (cf. chapter 3.2).
Variations of roughness height have no effect on the position of erosion area.

Time step As the sediment transport calculations is transient, an explicit calculation


scheme method is used. Owing to time step sensitivity using an explict scheme,
additionally tests are carried out by choosing ∂t = 5 s and ∂t = 60 s. The comparison
shows no variations among the simulations. All three chosen time steps give identical
results and are therefore not presented in more detail.
7 Numerical Simulation 37

(a) ks = 0.0007m (b) ks = 0.0035m

Figure 17: Bed elevation changes with regard to different bed roughness heights

Grid size The grid independence of the solution is checked using a finer mesh, where
the number of gridlines is doubled in all directions. It is exactly the same mesh
as used for the water flow calculation resolving the two spiral motions as described
in chapter 7.2. The amount of cells increases by a factor 23 , resulting in very long
computation time for time–dependent sediment transport. Investigating the time
step as discussed above, extending the time step to ∂t = 60 s has no influence on the
solution and fewer iterations have to be calculated, for the same simulation time of
∆t = 4 h. Still, the computation time is very long and is not suitable for practical
engineering purposes. The computation time for carrying out a simulation time of
∆t = 4 t is shown in Table 2.
Doubling the grid in all three spatial directions is a reasonable approach in order
to test the grid independence. The aspect ratio remains unchanged, so influences of
a distorted grid are eliminated. Fig. 18 shows the predicted bed formation in a bend,
using a time step of ∂t = 60 s. The scour point is still at the same position. The
deposition process tends to separate into two individual pool bars. Between them,

Table 2: Computation time for a simulation time of ∆t = 4 h


conducted case time step computation time
base case ∂t = 20 s 13.5 h
number of grid lines doubled ∂t = 60 s 240.0 h
simulations were conducted on a Pentium IV, 2.7 MHz, 1 GB RAM
7 Numerical Simulation 38

Figure 18: Sediment calculation with refined mesh and ∂t = 60s

a scour point develops. Comparing the numerical result to another photo taken of
a bend during the experimental run, similarities are observed. At a different bend
a different bed formation pattern appears as presented in Fig. 22. To gather from
this observation, neither the experimental run nor the numerical simulation seems
to tend to a prefered solution. The predictions show variations in detail, but the
main characteristic, which is the formation of a pool bar at the convex apex, remains
unchanged.
Wilson et al. [28] conducts numerical water flow simulations of a pseudo–natural
meandering flow, using a distorted computation mesh in a longitudinal direction.
To check if the result is mesh independent, a simulation is conducted using a mesh
with double the number of gridlines in a longitudinal direction, making the mesh less
distorted. The computed velocities from the two grids were found to be reasonably in
agreement. Mesh independence is confirmed and even a distorted mesh gives accurate
results.
Also a distorted mesh was tested for sediment transport calculations in this study.
Therefore, the number of grid cells in only the transversal and vertical direction was
doubled. Simulations having a time step of ∂t = 20 s and ∂t = 60 s were carried out.
The sediment calculation results are presented in Fig. 19. Considering a time step of
∂t = 60 s of the fine mesh calculation and comparing the results achieved by a fine
but distorted mesh, the bed formation pattern is quite similar. This observation leads
to the conclusion that the accuracy of the solution is not affected by a refinement of
the grid in a longitudinal direction.
7 Numerical Simulation 39

(a) ∂t = 20s (b) ∂t = 60s

Figure 19: Sediment calculations with distorted mesh at different time steps

7.4 Uncertainties in Numerical Modelling


In the numerical simulations there are three categories of error: (1) errors in the de-
scription of the physical process, (2) errors arising from the numerical implementation
of the physical process, and (3) errors concerning uncertainties of physical proper-
ties. The first mode of error is resulting from assumptions and approximations which
are applied in the simplified governing equations (cf. chapter 2.3). The convection–
diffusion equation (2.10) calculates the turbulent diffusion parameter Et based on the
k– model. The k– model calculates the turbulent diffusion coefficient as an average
of the three spatial directions [15]. In experiments it is found that the eddy–viscosity
coefficient depends on the spatial direction and is therefore non-isotropic. For some
cases the non-isotropy of the eddy–viscosity may influence the final results.
van Rijn [25] developed the equation (3.12) for bed concentration based on mea-
surements of concentration and experimental data from a huge number of flume
studies with sand particles. Most of them have been conducted and documented by
researchers other than van Rijn [25]. It should be pointed out that the derivation of
eqn. (3.12) is of an empirical nature and contains approximations and assumptions.
The equation is tested for straight channels with logarithmic velocity profile; in other
words for flows of mainly 1D character. It is not verified if the prediction of bed
concentration holds for complex three–dimensional flows with sharply curved bends
without having a logarithmic velocity profile as they are investigated in this thesis.
Owing to a lack of different approaches, the mentioned equation of van Rijn [25] must
be used.
The second category comprises the errors arising from the numerical implementa-
7 Numerical Simulation 40

tion of the physical processes, e.g. false diffusion generated by the numerical grid or
non-orthogonal diffusive terms, which are neglected for the calculation of the sedi-
ment concentrations. An injudicious parameter choice can produce highly erroneous
results, even if remaining parameters are set correctly. On the other hand, a reason-
able diligence in the model establishment by the user may minimize these errors.
The remedy to prevent false diffusion is to align the grid with the fluid direction
and/or increase the number of elements in the grid [15]. The author has tested the
second approach. As described in chapter 7.3.4, doubling the number of cells in each
direction does not change the accuracy of the solution. Consequently, inaccuracy is
not caused by false diffusion. Another alternative is to use a second-order method as
applied throughout all simulations.
The third category of error mentioned concerns the uncertainty of physical prop-
erties, e.g. roughness height and angle of repose φ. The latter is set to a constant
value as recommended by the literature for submerged sand particles. During the
experimental runs, φ varies with time. At the beginning of the experiment the sand
is dry and becomes saturated with water gradually. The variation of φ during the
experiment is time-dependent and is not reproduceable in SSIIM.
Special attention is paid to the equivalent sand roughness coefficient ks as it influences
the bed friction and therefore the bed–load transport. In the numerical simulation,
the ks –value is adjusted until the model output matches the measured conditions
by the experiment. This is achieved using an equivalent sand roughness height of
ks = 0.0021 m, which corresponds to ks = 2.3d90 , if d90 = 0.0009 m as seen in Fig. 8.
The value found from this calibration is in the range of empirically found values pre-
sented by Wilson et al. [28].
Fig. 16(a) shows an overprediction in erosion by the SSIIM software. The sediment
material located at the convex bank is carried away by current force. The photos
shown in Fig. 21 and Fig. 22 taken from the bend do not exhibit such strong erosion,
but rather demonstrate resistance to the acting force. Speculating about this observa-
tion, the sediment particles could exercise cohesion forces and stick together, having
the effect of variation in threshold condition for incipient sediment motion. Besides
this explanation, there could also be a compaction of the sediment material during
the formation of the trapezoidal shape. The mentioned influences are very uncertain
and almost impossible to quantify or to consider in the numerical simulation.
8 Conclusions and Outlook
This thesis presents a comparison of simulations performed in the numerical model
SSIIM and a physical model. The objective is to examine the modelling results
with respect to the sediment processes observed in the experimental run in order to
determine whether or not the numerical model SSIIM is able to predict sediment
transport correctly.
The physical process of sediment transport and its implementation in the numerical
model was described. Furthermore, the software code was extended by a subroutine
taking generalized sloping bed for incipient particle motion into account [5]. The sub-
routine emerges as the key for achieving improved results, reproducing the sediment
processes as closely as possible to the experiment.
Influencing factors are the calibration coefficients, i.e. the equivalent sand rough-
ness height. Suggestions from the literature were adopted, but further simulations
have shown that small variations in roughness will strongly effect the bed formation
pattern. The time step variation shows minor change, so that this parameter should
not be taken as a calibration coefficient of the numerical model. The degree of its
influence on the results in combination with changes in grid size could be an interest-
ing topic for further studies. But due to long computation times, such investigations
can not be carried out in a reasonable time with present computer capacity.
The numerical model results have proved to be grid size independent in combination
with the time steps chosen by the author. A refined grid shows separation of the point
bar, reflecting observations documented by photos but not by measurements.
The software SSIIM performs capably in reproducing sediment transport processes
in a realistic way, even with regard to sharply curved bends. The calculation can be
carried out if the user establishes reasonable model assumptions and approximations.
In this thesis, results in good agreement to the experiment have been achieved even
with the coarsest grid tested and a time step of ∂t = 60 s for a simulation time
of ∆t = 4 t. The computation time is as long as it is appropriate for practical
engineering purpose. Refining the computational mesh or decreasing the time step
means extending computation time, which is not feasible with the power of present–
day personal computers (cf. Table 2). Conducted calculations concerning variation
of the mentioned numerical parameters do not yield any improvement of the result
accuracy. They show variations of the bed formation pattern as also observed in
the experiment. However, without having field data, running reliable simulations is
hardly not possible. But even if applying unphysical parameters, SSIIM is able to
indicate the trend of pool- resp. point bar formation.
SSIIM’s extension by the implemented subroutine shows that the software is a re-
liable instrument in river, environmental and sedimentation engineering. The ability
to run the software on standard personal computers will promote its distribution
among practical working hydraulic engineers and will cause it to be accepted as a
design tool for water and environmental related engineering.
A Listing C++ Subroutine 42

A Listing C++ Subroutine

/************************************************************/
/**** correction for sloping bed ****************************/
/**** only called if F 7 B is used***************************/
/************************************************************/

// variables:
// vectorX: component in x direction of unit vector normal
to the bed surface
// vectorY: component in y direction of unit vector normal
to the bed surface
// vectorZ: component in z direction of unit vector normal
to the bed surface
// velocityX: velocity in x direction in bed cell
// velocityY: velocity in y direction in bed cell
// param1: parameter in formula
// param2: parameter in formula

double computeBedSlopeCorrection(double vectorX, double vectorY,


double vectorZ, double velocityX,
double velocityY, double param1,
double param2) {

double alpha, phi, teta, eta, tau, AT, PT;

eta = 0.954;
phi = param1;
teta = param2;

alpha = fabs(acos(vectorZ));
tau = 0;
PT = phi/teta;

if (alpha >= teta) AT = 1; //inclination of transverse cannot


else AT = alpha/teta; //be larger than angle-of-repose

//function//
tau = max(0.333,
eta * pow(1-(PT),0.745) * pow(1-(AT),0.372)
A Listing C++ Subroutine 43

); //max, barrier for alpha-->teta


return tau;}

//************ from Joern ****************//


B Listing of Sediment Calculation Input File 44

B Listing of Sediment Calculation Input File

T sed_shf_.21mm_20s_NNN_9degree title field


F 1 CD
F 2 RIS run option (result, initialize, sediment)
F 4 0.5 50 0.001 concentration calculation, relaxation,
iterations, convergence
F 7 B correction sloping bed
F 11 2.65 -0.03 density and shield coefficient
F 16 0.0021 roughness coef. of walls and bed
F 33 20.0 100 time step 20sec, inner iterations
F 37 1 tcs algorith
F 48 9 tecplot option
F 54 0.001 residual during transient calculation convergence criteria
F 56 10 0.5 number iterations, angle of response
f 78 1 0.0 sediments not follow flow vectors
F 84 1 v rijn bed load formular

F 109 5.368e-04 0.6 0.0 bed slope phi teta 0.0


F 111 1 beddll.dll

F 155 2.0 0 different sediment distribution at banks

G 1 261 17 9 1 grid and array sizes


G 3 0.000 12.5 25.00 37.5 50.0 62.50 75.00 87.50 100.00
vertical grid distribution
G 7 0 1 2 17 2 9 0 0 0.001 1 0 0.0 inflow
G 7 1 -1 2 17 2 9 0 0 0.001 1 0 0 outflow

P 10 1 number of iterations between printing results tecplot

S 1 0.0007 0.09 sediments_size and fall velocity


N 0 1 1 composition of sieve line
I 1 0.00 sediment inflow kg/s
B 0 0 0 0 0 spatial distribution of sediment

W 1 34.15 0.001 0.04397 kst, discharge, downstream water level


W 2 8 1 40 80 120 160 200 240 261 backwater calculation

K 1 720 60000 iteration options 4hours


B Listing of Sediment Calculation Input File 45

K 2 0 1 sidewall/surface 0=wall law/1=free surface


K 3 0.8 0.8 0.8 0.3 0.5 0.5 relaxation factors
K 6 1 1 1 0 0 0 1-->2nd order up-w.scheme, u,v,w,p,k,e
C Photos 46

C Photos

Figure 20: Channel before experimental run, Guymer and Dutton [8]
C Photos 47

Figure 21: Photo of channel bend showing one point bar, flow is from left to right,
Guymer and Dutton [8]

Figure 22: Photo of channel bend showing separated point bars, flow is from left to
right, Guymer and Dutton [8]
References 48

References
[1] Wasser- und Schifffahrtsamt Freiburg. Information Flyer, 2004.

[2] Bathurst, James C., Thorne, Colin R., and Hey, Richard D. Secondary flow and
shear stress at river bends. Journal of the Hydraulics Division, 10:1277–1295,
October 1979.

[3] Chanson, Hubert. The Hydraulics of Open Channel Flow: An Introduction. El-
sevier Butterworth–Heinemann, Amsterdam, Bosten, Heidelberg, London, New
York, Oxford, 2 edition, 2004. ISBN 0–7506–5978–5.

[4] Ciray, Cahit. On secondary currents. The International Association for Hy-
draulic Research, 1:408–413, September 1967. 12th IAHR–Congress, Fort
Collins, 11–14. September 1967.

[5] Dey, Subhasish. Experimental study on incipient motion of sediment particles


on generalized sloping fluvial beds. International Journal of Sediment Research,
16:391–398, 2001. No. 3.

[6] Dey, Subhasish. Threshold of sediment motion on combined transverse and


longitudinal sloping beds. Journal of Hydraulic Research, 41:405–415, 2003.
No. 4.

[7] Fischer, Hugo B., List, E. John, Koh, Robert C. Y., Imberger, Joerg, and Brooks,
Norman H. Mixing in Inland and Coastal Waters. Academic Press, New York,
London, Toronto, Sydney, San Francisco, 1979. ISBN 0-12-258150-4.

[8] Guymer, Ian and Dutton, Richard. Measurement data and photographs of phys-
ical experiment. unpublished, 2003. kindly left to the author.

[9] Imiela, Manfred. Langfristige mehrdimensionale Feststofftransportmodellierung


zum morphologischen Gleichgewichtszustand anhand von Modellversuchen und
einer Renaturierungsmaßnahme an der Enz. Master’s thesis, Universität Karls-
ruhe, 2004. Institut für Hydromechanik.

[10] Julien, Pierre Y. Erosion and Sedimentation. Cambridge University Press, Cam-
bridge, United Kingdom, 1989. ISBN 0–521–63639–6.

[11] Kobus, Helmut. Wasserbauliches Versuchswesen, volume 4. DVWW, Essen,


1978. Mitteilungsheft.

[12] Leschziner, Michael A. and Rodi, Wolfgang. Calculation of strongly curved open
channel flow. Journal of the Hydraulics Division, 10:1297–1314, October 1979.
References 49

[13] Meckel, Hermann. Spiralströmung und Sedimentbewegung in Fluss– und


Kanalkrümmungen. Wasserwirtschaft, 10:287–294, Oktober 1978. Heft 10.

[14] Mosonyi, Emil and Götz, Werner. Secondary currents in subsequent model
bends. International Symposium on River Mechanics, 1:191–201, January 1973.
Proceedings of the International Association for Hydraulic Research, 9–12 Jan-
uary 1973, Asian Institute of Technology, Bangkok, Thailand.

[15] Olsen, Nils Reidar B. A three-dimensional numerical model for simulation of


sediment movements in water intakes. PhD thesis, Division of Hydraulic En-
gineering, University of Trondheim, Norwegian Institute of Technology, 1991.
59.

[16] Olsen, Nils Reidar B. A Three–Dimensional Numerical Model for Simulation


of Sediment Movements in Water Intakes with Multiblock Option. Department
of Hydraulic and Environmental Engineering, The Norwegian University of Sci-
ence and Technology, 2002. URL http://www.ntnu.no/~nilsol/ssiimwin/
manual3.pdf. user’s manual.

[17] Patankar, Suhas V. Numerical heat transfer and fluid flow. Hemisphere
Publ. Corp., New York, 1980. ISBN 0-89116-522-3.

[18] Prandtl, Ludwig, Oswatitsch, Klaus, and Wieghardt, Karl. Führer durch die
Strömungslehre. Friedr. Vieweg & Sohn, Braunschweig, Wiesbaden, 1984. ISBN
3–528-18209–5. 8. vollständig überarbeitete Auflage.

[19] Rodi, Wolfgang. Turbulence Models and Their Application in Hydraulics. Inter-
national Association for Hydraulics Research, Balkema, Rotterdam, 3rd edition,
1993. ISBN 90 5410 150 4. A state–of–the–art review.

[20] Rouse, Hunter. Hydraulics, Mechanics of Fluids, Engineering Education, pages


57–61. Selected Writings of Hunter Rouse. Dover Publications, Inc., New York.
Institute of Hydraulic Research, University of Iowa.

[21] Rozovskii, I. L. Flow of Water in Bends of Open Channels, volume PST Cat.
No. 363. Academy of Sciences of the Ukrainian SSR, Kiev, 1957. translation by
Y. Prushansky, Israel Program for Scientific Translation, 1961.

[22] Schlichting, Herrmann and Gersten, Klaus. Boundary Layer Theory. Springer
Verlag, Berlin, Heidelberg, New York, 8th revised and enlarged edition, 2000.
ISBN 3-540-66270-7.

[23] Shields, Albert. Anwendung der Ähnlichkeitsmechanik und der Turbulenz-


forschung auf die Geschiebebewegung. Mitteilungen der Preußischen Versuch-
sanstalt für Wasserbau und Schiffbau, 26:26, 1936. (in German).
[24] Socolofsky, Scott A. and Jirka, Gerhard H. Environmental Fluid Mechanics
Part I, Mass Transfer and Diffusion. Class notes, 2002. Institut for Hydrome-
chanics, University of Karlsruhe.

[25] van Rijn, Leo C. Sediment Transport, Part I: Bed Load Transport. Journal
of Hydraulic Engineering, American Society of Civil Engineers, 110:1431–1456,
October 1984. No. 10.

[26] van Rijn, Leo C. Sediment Transport, Part II: Suspended Load Transport.
Journal of Hydraulic Engineering, American Society of Civil Engineers, 110:
1613–1641, November 1984. No. 11.

[27] van Rijn, Leo C. Sediment Transport, Part III: Bed Forms and Alluvial Rough-
ness. Journal of Hydraulic Engineering, American Society of Civil Engineers,
110:1733–1755, December 1984. No. 12.

[28] Wilson, C.A.M.E., Boxall, J.B., Guymer, I., and Olsen, N.R.B. Validation of a
three–dimensional numerical code in the simulation of pseudo–natural meander-
ing flows. Journal of the Hydraulic Engineering, 10:758–768, October 2003.

[29] Zanke, Ulrich. Grundlagen der Sedimentbewegung. Springer Verlag, Berlin, Hei-
delberg, New York, 1982. ISBN 3–540–11672–9.

Vous aimerez peut-être aussi